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 I(Ks) potassium channel, critical to control of heart electrical activity, requires assembly of α (KCNQ1) and β (KCNE1) subunits. Inherited mutations in either I(Ks) channel subunit are associated with cardiac arrhythmia syndromes. Two mutations (S140G and V141M) that cause familial atrial fibrillation (AF) are located on adjacent residues in the first membrane-spanning domain of KCNQ1, S1. These mutations impair the deactivation process, causing channels to appear constitutively open. Previous studies suggest that both mutant phenotypes require the presence of KCNE1. Here we found that despite the proximity of these two mutations in the primary protein structure, they display different functional dependence in the presence of KCNE1. In the absence of KCNE1, the S140G mutation, but not V141M, confers a pronounced slowing of channel deactivation and a hyperpolarizing shift in voltage-dependent activation. When coexpressed with KCNE1, both mutants deactivate significantly slower than wild-type KCNQ1/KCNE1 channels. The differential dependence on KCNE1 can be correlated with the physical proximity between these positions and KCNE1 as shown by disulfide cross-linking studies: V141C forms disulfide bonds with cysteine-substituted KCNE1 residues, whereas S140C does not. These results further our understanding of the structural relationship between KCNE1 and KCNQ1 subunits in the I(Ks) channel, and provide mechanisms for understanding the effects on channel deactivation underlying these two atrial fibrillation mutations.  相似文献   

2.
Long QT syndrome is an inherited disorder of cardiac repolarization caused by mutations in cardiac ion channel genes, including KVLQT1. In this study, the functional consequences of three long QT-associated missense mutations in KvLQT1 (R243C, W248R, E261K) were characterized using the Xenopus oocyte heterologous expression system and two-microelectrode voltage clamp techniques. These mutations are located in or near the intracellular linker between the S4 and S5 transmembrane domains, a region implicated in activation gating of potassium channels. The E261K mutation caused loss of function and did not interact with wild-type KvLQT1 subunits. R243C or W248R KvLQT1 subunits formed functional channels, but compared with wild-type KvLQT1 current, the rate of activation was slower, and the voltage dependence of activation and inactivation was shifted to more positive potentials. Co expression of minK and KvLQT1 channel subunits induces a slow delayed rectifier K(+) current, I(Ks), characterized by slow activation and a markedly increased magnitude compared with current induced by KvLQT1 subunits alone. Coexpression of minK with R243C or W248R KvLQT1 subunits suppressed current, suggesting that coassembly of mutant subunits with minK prevented normal channel gating. The decrease in I(Ks) caused by loss of function or altered gating properties explains the prolonged QT interval and increased risk of arrhythmia and sudden death associated with these mutations in KVLQT1.  相似文献   

3.
Cardiac slow delayed rectifier (IKs) channel complex consists of KCNQ1 channel and KCNE1 auxiliary subunits. The extracellular juxtamembranous region of KCNE1 is an unstructured loop that contacts multiple KCNQ1 positions in a gating-state-dependent manner. Congenital arrhythmia-related mutations have been identified in the extracellular S1–S2 linker of KCNQ1. These mutations manifest abnormal phenotypes only when coexpressed with KCNE1, pointing to the importance of proper KCNQ1/KCNE1 interactions here in IKs channel function. We investigate the interactions between the KCNE1 loop (positions 36–47) and KCNQ1 S1–S2 linker (positions 140–148) by means of disulfide trapping and voltage clamp techniques. During transitions among the resting-state conformations, KCNE1 positions 36–43 make contacts with KCNQ1 positions 144, 145, and 147 in a parallel fashion. During conformational changes in the activated state, KCNE1 position 40 can make contacts with all three KCNQ1 positions, while the neighboring KCNE1 positions (36, 38, 39, and 41) can make contact with KCNQ1 position 147. Furthermore, KCNQ1 positions 143 and 146 are high-impact positions that cannot tolerate cysteine substitution. To maintain the proper IKs channel function, position 143 requires a small side chain with a hydroxyl group, and position 146 requires a negatively charged side chain. These data and the proposed molecular motions provide insights into the mechanisms by which mutations in the extracellular juxtamembranous region of the IKs channel impair its function.  相似文献   

4.
Structural effects of yeast cofilin on skeletal muscle and yeast actin were examined in solution. Cofilin binding to native actin was non-cooperative and saturated at a 1:1 molar ratio, with K(d)相似文献   

5.
ClC-1 is a dimeric, double-pored chloride channel that is present in skeletal muscle. Mutations of this channel can result in the condition myotonia, a muscle disorder involving increased muscle stiffness. It has been shown that the dominant form of myotonia often results from mutations that affect the so-called slow, or common, gating process of the ClC-1 channel. Mutations causing dominant myotonia are seen to cluster at the interface of the ClC-1 channel monomers. This study has investigated the role of the H, I, P, and Q helices, which lie on this interface, as well as the G helix, which is situated immediately behind the H and I helices, on ClC-1 gating. 11 mutant ClC-1 channels (T268M, C277S, C278S, S289A, T310M, S312A, V321S, T539A, S541A, M559T, and S572V) were produced using site-directed mutagenesis, and gating properties of these channels were investigated using electrophysiological techniques. Six of the seven mutations in G, H, and I, and two of the four mutations in P and Q, caused shifts of the ClC-1 open probability. In the majority of cases this was due to alterations in the common gating process, with only three of the mutants displaying any change in fast gating. Many of the mutant channels also showed alterations in the kinetics of the common gating process, particularly at positive potentials. The changes observed in common gating were caused by changes in the opening rate (e.g. T310M), the closing rate (e.g. C277S), or both rates. These results indicate that mutations in the helices forming the dimer interface are able to alter the ClC-1 common gating process by changing the energy of the open and/or closed channel states, and hence altering transition rates between these states.  相似文献   

6.
Multiple transmembrane (TM) segments line the pore of the cystic fibrosis transmembrane conductance regulator Cl(-) channel; however, the relative alignment of these TMs and their relative movements during channel gating are unknown. To gain three-dimensional structural information on the outer pore, we have used patch clamp recording to study the proximity of pairs of cysteine side chains introduced into TMs 6 and 11, using both disulfide cross-linking and Cd(2+) coordination. Following channel activation, disulfide bonds could apparently be formed between three cysteine pairs (of 15 studied): R334C/T1122C, R334C/G1127C, and T338C/S1118C. To examine the state dependence of cross-linking, we combined these cysteine mutations with a nucleotide-binding domain mutation (E1371Q) that stabilizes the channel open state. Investigation of the effects of the E1371Q mutation on disulfide bond formation and Cd(2+) coordination suggests that although R334C/T1122C and T338C/S1118C are closer together in the channel open state, R334C/G1127C are close together and can form disulfide bonds only when the channel is closed. These results provide important new information on the three-dimensional structure of the outer mouth of the cystic fibrosis transmembrane conductance regulator channel pore: TMs 6 and 11 are close enough together to form disulfide bonds in both open and closed channels. Moreover, the altered relative locations of residues in open and in closed channels that we infer allow us to propose that channel opening and closing may be associated with a relative translational movement of TMs 6 and 11, with TM6 moving "down" (toward the cytoplasm) during channel opening.  相似文献   

7.
NM23-H1 is a member of the NM23/NDP kinase gene family and a putative metastasis suppressor. Previously, a screen for NM23-H1-interacting proteins that could potentially modulate its activity identified serine-threonine kinase receptor-associated protein (STRAP), a transforming growth factor (TGF)-beta receptor-interacting protein. Through the use of cysteine to serine amino acid substitution mutants of NM23-H1 (C4S, C109S, and C145S) and STRAP (C152S, C270S, and C152S/C270S), we demonstrated that the association between these two proteins is dependent on Cys(145) of NM23-H1 and Cys(152) and Cys(270) of STRAP but did not appear to involve Cys(4) and Cys(109) of NM23-H1, suggesting that a disulfide linkage involving Cys(145) of NM23-H1 and Cys(152) or Cys(270) of STRAP mediates complex formation. The interaction was dependent on the presence of dithiothreitol or beta-mercaptoethanol but not H(2)O(2). Ectopic expression of wild-type NM23-H1, but not NM23-H1(C145S), negatively regulated TGF-beta signaling in a dose-dependent manner, enhanced stable association between the TGF-beta receptor and Smad7, and prevented nuclear translocation of Smad3. Similarly, wild-type NM23-H1 inhibited TGF-beta-induced apoptosis and growth inhibition, whereas NM23-H1(C145S) had no effect. Knockdown of NM23-H1 by small interfering RNA stimulated TGF-beta signaling. Coexpression of wild-type STRAP, but not STRAP(C152S/C270S), significantly stimulated NM23-H1-induced growth of HaCaT cells. These results suggest that the direct interaction of NM23-H1 and STRAP is important for the regulation of TGF-beta-dependent biological activity as well as NM23-H1 activity.  相似文献   

8.
Dynamic properties of F-actin structure prompted suggestions (Squire, J. M., and Morris, E. P. (1998) FASEB J. 12, 761-771) that actin subdomain 2 movements play a role in thin-filament regulation. Using fluorescently labeled yeast actin mutants Q41C, Q41C/C374S, and D51C/C374S and azidonitrophenyl putrescine (ANP) Gln(41)-labeled alpha-actin, we monitored regulation-linked changes in subdomain 2. These actins had fully regulated acto-S1 ATPase activities, and emission spectra of regulated Q41C(AEDANS)/C374S and D51C(AEDANS)/C374S filaments did not reveal any calcium-dependent changes. Fluorescence energy transfer in these F-actins mostly occurred from Trp(340) and Trp(356) to 5-(2((acetyl)amino)ethyl)amino-naphthalene-1-sulfonate (AEDANS)-labeled Cys(41) or Cys(51) of adjacent same strand protomers. Our results show that fluorescence energy transfer between these residues is similar in the mostly blocked (-Ca(2+)) and closed (+Ca(2+)) states. Ca(2+) also had no effect on the excimer band in the pyrene-labeled Q41C-regulated actin, indicating virtually no change in the overlap of pyrenes on Cys(41) and Cys(374). ANP quenching of rhodamine phalloidin fluorescence showed that neither Ca(2+) nor S1 binding to regulated alpha-actin affects the phalloidin-probe distance. Taken together, our results indicate that transitions between the blocked, closed, and open regulatory states involve no significant subdomain 2 movements, and, since the cross-linked alpha-actin remains fully regulated, that subdomain 2 motions are not essential for actin regulation.  相似文献   

9.
The E462R mutation in the fifth position of the AID (alpha1 subunit interaction domain) region in the I-II linker is known to significantly accelerate voltage-dependent inactivation (VDI) kinetics of the L-type CaV1.2 channel, suggesting that the AID region could participate in a hinged-lid type inactivation mechanism in these channels. The recently solved crystal structures of the AID-CaVbeta regions in L-type CaV1.1 and CaV1.2 channels have shown that in addition to E462, positions occupied by Q458, Q459, E461, K465, L468, D469, and T472 in the rabbit CaV1.2 channel could also potentially contribute to a hinged-lid type mechanism. A mutational analysis of these residues shows that Q458A, Q459A, K465N, L468R, D469A, and T472D did not significantly alter VDI gating. In contrast, mutations of the negatively charged E461, E462, and D463 to neutral or positively charged residues increased VDI gating, suggesting that the cluster of negatively charged residues in the N-terminal end of the AID helix could account for the slower VDI kinetics of CaV1.2. A mutational analysis at position 462 (R, K, A, G, D, N, Q) further confirmed that E462R yielded faster VDI kinetics at +10 mV than any other residue with E462R > E462K approximately E462A > E462N > wild-type approximately E462Q approximately E462G > E462D (from the fastest to the slowest). E462R was also found to increase the VDI gating of the slow CEEE chimera that includes the I-II linker from CaV1.2 into a CaV2.3 background. The fast VDI kinetics of the CaV1.2 E462R and the CEEE + E462R mutants were abolished by the CaVbeta2a subunit and reinstated when using the nonpalmitoylated form of CaVbeta2a C3S + C4S (CaVbeta2a CS), confirming that CaVbeta2a and E462R modulate VDI through a common pathway, albeit in opposite directions. Altogether, these results highlight the unique role of E461, E462, and D463 in the I-II linker in the VDI gating of high-voltage activated CaV1.2 channels.  相似文献   

10.
Voltage-dependent calcium channels consist of a pore-forming subunit (Ca(V)alpha(1)) that includes all the molecular determinants of a voltage-gated channel, and several accessory subunits. The ancillary beta-subunit (Ca(V)beta) is a potent activator of voltage-dependent calcium channels, but the mechanisms and structural bases of this regulation remain elusive. Ca(V)beta binds reversibly to a conserved consensus sequence in Ca(V)alpha(1), the alpha(1)-interaction domain (AID), which forms an alpha-helix when complexed with Ca(V)beta. Conserved aromatic residues face to one side of the helix and strongly interact with a hydrophobic pocket on Ca(V)beta. Here, we studied the effect of mutating residues located opposite to the AID-Ca(V)beta contact surface in Ca(V)1.2. Substitution of AID-exposed residues by the corresponding amino acids present in other Ca(V)alpha(1) subunits (E462R, K465N, D469S, and Q473K) hinders Ca(V)beta's ability to increase ionic-current to charge-movement ratio (I/Q) without changing the apparent affinity for Ca(V)beta. At the single channel level, these Ca(V)1.2 mutants coexpressed with Ca(V)beta(2a) visit high open probability mode less frequently than wild-type channels. On the other hand, Ca(V)1.2 carrying either a mutation in the conserved tryptophan residue (W470S, which impairs Ca(V)beta binding), or a deletion of the whole AID sequence, does not exhibit Ca(V)beta-induced increase in I/Q. In addition, we observed a shift in the voltage dependence of activation by +12 mV in the AID-deleted channel in the absence of Ca(V)beta, suggesting a direct participation of these residues in the modulation of channel activation. Our results show that Ca(V)beta-dependent potentiation arises primarily from changes in the modal gating behavior. We envision that Ca(V)beta spatially reorients AID residues that influence the channel gate. These findings provide a new framework for understanding modulation of VDCC gating by Ca(V)beta.  相似文献   

11.
We have developed a new nonoverlapping infectious viral genome (NO-SV40) in order to facilitate structure-based analysis of the simian virus 40 (SV40) life cycle. We first tested the role of cysteine residues in the formation of infectious virions by individually mutating the seven cysteines in the major capsid protein, Vp1. All seven cysteine mutants-C9A, C49A, C87A, C104A, C207S, C254A, and C267L-retained viability. In the crystal structure of SV40, disulfide bridges are formed between certain Cys104 residues on neighboring pentamers. However, our results show that none of these disulfide bonds are required for virion infectivity in culture. We also introduced five different mutations into Cys254, the most strictly conserved cysteine across the polyomavirus family. We found that C254L, C254S, C254G, C254Q, and C254R mutants all showed greatly reduced (around 100,000-fold) plaque-forming ability. These mutants had no apparent defect in viral DNA replication. Mutant Vp1's, as well as wild-type Vp2/3, were mostly localized in the nucleus. Further analysis of the C254L mutant revealed that the mutant Vp1 was able to form pentamers in vitro. DNase I-resistant virion-like particles were present in NO-SV40-C254L-transfected cell lysate, but at about 1/18 the amount in wild-type-transfected lysate. An examination of the three-dimensional structure reveals that Cys254 is buried near the surface of Vp1, so that it cannot form disulfide bonds, and is not involved in intrapentamer interactions, consistent with the normal pentamer formation by the C254L mutant. It is, however, located at a critical junction between three pentamers, on a conserved loop (G2H) that packs against the dual interpentamer Ca(2+)-binding sites and the invading C-terminal helix of an adjacent pentamer. The substitution by the larger side chains is predicted to cause a localized shift in the G2H loop, which may disrupt Ca(2+) ion coordination and the packing of the invading helix, consistent with the defect in virion assembly. Our experimental system thus allows dissection of structure-function relationships during the distinct steps of the SV40 life cycle.  相似文献   

12.
In contrast to fast inactivation, the molecular basis of sodium (Na) channel slow inactivation is poorly understood. It has been suggested that structural rearrangements in the outer pore mediate slow inactivation of Na channels similar to C-type inactivation in potassium (K) channels. We probed the role of the outer ring of charge in inactivation gating by paired cysteine mutagenesis in the rat skeletal muscle Na channel (rNav1.4). The outer charged ring residues were substituted with cysteine, paired with cysteine mutants at other positions in the external pore, and coexpressed with rat brain beta1 in Xenopus oocytes. Dithiolthreitol (DTT) markedly increased the current in E403C+E758C double mutant, indicating the spontaneous formation of a disulfide bond and proximity of the alpha carbons of these residues of no more than 7 A. The redox catalyst Cu(II) (1,10-phenanthroline)3 (Cu(phe)3) reduced the peak current of double mutants (E403C+E758C, E403C+D1241C, E403C+D1532C, and D1241C+D1532C) at a rate proportional to the stimulation frequency. Voltage protocols that favored occupancy of slow inactivation states completely prevented Cu(phe)3 modification of outer charged ring paired mutants E403C+E758C, E403C+D1241C, and E403C+D1532C. In contrast, voltage protocols that favored slow inactivation did not prevent Cu(phe)3 modification of other double mutants such as E403C+W756C, E403C+W1239C, and E403C+W1531C. Our data suggest that slow inactivation of the Na channel is associated with a structural rearrangement of the outer ring of charge.  相似文献   

13.
Opening and closing of the cystic fibrosis transmembrane conductance regulator (CFTR) chloride channel are controlled by ATP binding and hydrolysis by its nucleotide binding domains (NBDs). This is presumed to control opening of a single "gate" within the permeation pathway, however, the location of such a gate has not been described. We used patch clamp recording to monitor access of cytosolic cysteine reactive reagents to cysteines introduced into different transmembrane (TM) regions in a cysteine-less form of CFTR. The rate of modification of Q98C (TM1) and I344C (TM6) by both [2-sulfonatoethyl] methanethiosulfonate (MTSES) and permeant Au(CN)(2)(-) ions was reduced when ATP concentration was reduced from 1mM to 10μM, and modification by MTSES was accelerated when 2mM pyrophosphate was applied to prevent channel closure. Modification of K95C (TM1) and V345C (TM6) was not affected by these manoeuvres. We also manipulated gating by introducing the mutations K464A (in NBD1) and E1371Q (in NBD2). The rate of modification of Q98C and I344C by both MTSES and Au(CN)(2)(-) was decreased by K464A and increased by E1371Q, whereas modification of K95C and V345C was not affected. These results suggest that access from the cytoplasm to K95 and V345 is similar in open and closed channels. In contrast, modifying ATP-dependent channel gating alters access to Q98 and I344, located further into the pore. We propose that ATP-dependent gating of CFTR is associated with the opening and closing of a gate within the permeation pathway at the level of these pore-lining amino acids.  相似文献   

14.
Site-3 toxins have been shown to inhibit a component of gating charge (33% of maximum gating charge, Q(max)) in native cardiac Na channels that has been identified with the open-to-inactivated state kinetic transition. To investigate the role of the three outermost arginine amino acid residues in segment 4 domain IV (R1, R2, R3) in gating charge inhibited by site-3 toxins, we recorded ionic and gating currents from human heart Na channels with mutations of the outermost arginines (R1C, R1Q, R2C, and R3C) expressed in fused, mammalian tsA201 cells. All four mutations had ionic currents that activated over the same voltage range with slope factors of their peak conductance-voltage (G-V) relationships similar to those of wild-type channels, although decay of I(Na) was slowest for R1C and R1Q mutant channels and fastest for R3C mutant channels. After Na channel modification by Ap-A toxin, decays of I(Na) were slowed to similar values for all four channel mutants. Toxin modification produced a graded effect on gating charge (Q) of mutant channels, reducing Q(max) by 12% for the R1C and R1Q mutants, by 22% for the R2C mutant, and by 27% for the R3C mutant, only slightly less than the 31% reduction seen for wild-type currents. Consistent with these findings, the relationship of Q(max) to G(max) was significantly shallower for R1 mutants than for R2C and R3C mutant Na channels. These data suggest that site-3 toxins primarily inhibit gating charge associated with movement of the S4 in domain IV, and that the outermost arginine contributes the largest amount to channel gating, with other arginines contributing less.  相似文献   

15.
Bian Y  Liang X  Fang N  Tang XF  Tang B  Shen P  Peng Z 《FEBS letters》2006,580(25):6007-6014
Thermophilic WF146 protease possesses four surface loop insertions and a disulfide bond, resembling its psychrophilic (subtilisins S41 and S39) and mesophilic (subtilisins SSII and sphericase) homologs. Deletion of the insertion 3 (positions 193-197) or insertion 4 (positions 210-221) of WF146 protease resulted in a significant decrease of the enzyme stability. In addition, substitution of the residues Pro211 and Ala212 or residue Glu221 which localized in the vicinity of a Ca(2+) binding site of the enzyme by the corresponding residues in subtilisin S41 remarkably reduced the half-life of the enzyme at 70 degrees C, suggesting that the three residues contributed to the thermostability of the enzyme, probably by enhancing the affinity of enzyme to Ca(2+). In the presence of dithiothreitol, the WF146 protease suffered excessive autolysis, indicating that the Cys52-Cys65 disulfide bond played a critical role in stabilizing the WF146 protease against autolysis. The autolytic cleavage sites of the WF146 protease were identified to locate between residues Asn63-Gly64 and Cys65-Ala66 by N-terminal amino acid analysis of the autolytic product. It was noticed that the effect of the autolytic cleavage at Asn63-Gly64 could be compensated by the disulfide bond Cys52-Cys65 under non-reducing condition, and the disulfide bond cross-linked autolytic product remained active. The apparent stabilization effect of the disulfide bond Cys52-Cys65 in the WF146 protease might provide a rational basis for improving the stability of subtilase against autolysis by protein engineering.  相似文献   

16.
Transient receptor potential vanilloid subtype I (TRPV1) is a thermosensory ion channel that is also gated by chemical substances such as vanilloids. Adjacent to the channel gate, this polymodal thermoTRP channel displays a TRP domain, referred to as AD1, that plays a role in subunit association and channel gating. Previous studies have shown that swapping the AD1 in TRPV1 with the cognate from the TRPV2 channel (AD2) reduces protein expression and produces a nonfunctional chimeric channel (TRPV1-AD2). Here, we used a stepwise, sequential, cumulative site-directed mutagenesis approach, based on rebuilding the AD1 domain in the TRPV1-AD2 chimera, to unveil the minimum number of amino acids needed to restore protein expression and polymodal channel activity. Unexpectedly, we found that virtually full restitution of the AD1 sequence is required to reinstate channel expression and responses to capsaicin, temperature, and voltage. This strategy identified E692, R701, and T704 in the TRP domain as important for TRPV1 activity. Even conservative mutagenesis at these sites (E692D/R701K/T704S) impaired channel expression and abolished TRPV1 activity. However, the sole mutation of these positions in the TRPV1-AD2 chimera (D692E/K701R/S704T) was not sufficient to rescue channel gating, implying that other residues in the TRP domain are necessary to endow activity to TRPV1-AD2. A biophysical analysis of a functional chimera suggested that mutations in the TRP domain raised the energetics of channel gating by altering the coupling of stimuli sensing and pore opening. These findings indicate that inter- and/or intrasubunit interactions in the TRP domain are essential for correct TRPV1 gating.  相似文献   

17.
The production of class A beta-lactamases is a major cause of clinical resistance to beta-lactam antibiotics. Some of class A beta-lactamases are known to have a disulfide bridge. Both narrow spectrum and extended spectrum beta-lactamases of TEM and the SHV enzymes possess a disulfide bond between Cys77 and Cys123, and the enzymes with carbapenem-hydrolyzing activity have a well-conserved disulfide bridge between Cys69 and Cys238. We produced A77C/G123C mutant of the extended-spectrum beta-lactamase Toho-1 in order to introduce a disulfide bond between the cysteine residues at positions 77 and 123. The result of 5,5'-dithiobis-2-nitrobenzoic acid (DTNB) titrations confirmed formation of a new disulfide bridge in the mutant. The results of irreversible heat inactivation and circular dichroism (CD) melting experiments indicated that the disulfide bridge stabilized the enzyme significantly. Though kinetic analysis indicated that the catalytic properties of the mutant were quite similar to those of the wild-type enzyme, E. coli producing this mutant showed drug resistance significantly higher than E. coli producing the wild-type enzyme. We speculate that the stability of the enzymes provided by the disulfide bond may explain the wide distribution of TEM and SHV derivatives and explain how various mutations can cause broadened substrate specificity without loss of stability.  相似文献   

18.
The human voltage-gated sodium channel Nav1.7 plays a crucial role in transmission of noxious stimuli. The inherited pain disorder erythromelalgia (IEM) has been linked to Nav1.7 gain-of-function mutations. Here we show that the IEM-associated Q875E mutation located on the pore module of Nav1.7 produces a large hyperpolarizing shift (−18 mV) in the voltage dependence of activation. Three-dimensional homology modeling indicates that the side chains of Gln-875 and the gating charge Arg-214 of the domain I voltage sensor are spatially close in the activated conformation of the channel. We verified this proximity by using an engineered disulfide bridge approach. The Q875E mutation introduces a negative charge that may modify the local electrical field experienced by the voltage sensor and, upon activation, interact directly via a salt bridge with the Arg-214 gating charge residue. Together these processes could promote transition to, and stabilization of, the domain I voltage sensor in the activated conformation and thus produce the observed gain of function. In support of this hypothesis, an increase in the extracellular concentration of Ca2+ or Mg2+ reverted the voltage dependence of activation of the IEM mutant to near WT values, suggesting a cation-mediated electrostatic screening of the proposed interaction between Q875E and Arg-214.  相似文献   

19.
The crystal structures of two thermally stabilized subtilisin BPN' variants, S63 and S88, are reported here at 1.8 and 1.9 A resolution, respectively. The micromolar affinity calcium binding site (site A) has been deleted (Delta75-83) in these variants, enabling the activity and thermostability measurements in chelating conditions. Each of the variants includes mutations known previously to increase the thermostability of calcium-independent subtilisin in addition to new stabilizing mutations. S63 has eight amino acid replacements: D41A, M50F, A73L, Q206W, Y217K, N218S, S221C, and Q271E. S63 has 75-fold greater stability than wild type subtilisin in chelating conditions (10 mm EDTA). The other variant, S88, has ten site-specific changes: Q2K, S3C, P5S, K43N, M50F, A73L, Q206C, Y217K, N218S, and Q271E. The two new cysteines form a disulfide bond, and S88 has 1000 times greater stability than wild type subtilisin in chelating conditions. Comparisons of the two new crystal structures (S63 in space group P2(1) with A cell constants 41.2, 78.1, 36.7, and beta = 114.6 degrees and S88 in space group P2(1)2(1)2(1) with cell constants 54.2, 60.4, and 82.7) with previous structures of subtilisin BPN' reveal that the principal changes are in the N-terminal region. The structural bases of the stabilization effects of the new mutations Q2K, S3C, P5S, D41A, Q206C, and Q206W are generally apparent. The effects are attributed to the new disulfide cross-link and to improved hydrophobic packing, new hydrogen bonds, and other rearrangements in the N-terminal region.  相似文献   

20.
A thermophilic serine protease, Aqualysin I, from Thermus aquaticus YT-1 has two disulphide bonds, which are also found in a psychrophilic serine protease from Vibrio sp. PA-44 and a proteinase K-like enzyme from Serratia sp. at corresponding positions. To understand the significance of these disulphide bonds in aqualysin I, we prepared mutants C99S, C194S and C99S/C194S (WSS), in which Cys69-Cys99, Cys163-Cys194 and both of these disulphide bonds, respectively, were disrupted by replacing Cys residues with Ser residues. All mutants were expressed stably in Escherichia coli. The C99S mutant was 68% as active as the wild-type enzyme at 40 degrees C in terms of k(cat) value, while C194S and WSS were only 6 and 3%, respectively, as active, indicating that disulphide bond Cys163-Cys194 is critically important for maintaining proper catalytic site conformation. Mutants C194S and WSS were less thermostable than wild-type enzyme, with a half-life at 90 degrees C of 10 min as compared to 45 min of the latter and with transition temperatures on differential scanning calorimetry of 86.7 degrees C and 86.9 degrees C, respectively. Mutant C99S was almost as stable as the wild-type aqualysin I. These results indicate that the disulphide bond Cys163-Cys194 is more important for catalytic activity and conformational stability of aqualysin I than Cys67-Cys99.  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