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.
Grove et al. have demonstrated L-type Ca2+ channel activity of a synthetic channel peptide (SCP) composed of four helices (sequence: DPWNVFDFLI10VIGSIIDVIL20SE) tethered by their C-termini to a nanopeptide template. We sought to obtain the optimal conformations of SCP and locate the binding sites for Ca2+ and for the dihydropyridine ligand nifedipine. Eight Ca2+ ions were added to neutralize the 16 acidic residues in the helices. Eight patterns of the salt bridges between Ca2+ ions and pairs of the acidic residues were calculated by the Monte Carlo-with-energy-minimization (MCM) protocol. In the energetically optimal conformation, two Ca2+ ions were bound to Asp-1 residues at the intracellular side of SCP, and six Ca2+ ions were arrayed in two files at the diametrically opposite sides of the pore, implying a Ca2+ relay mechanism. Nine modes of nifedipine binding to SCP were simulated by the MCM calculations. In the energetically optimal mode, the ligand fits snugly in the pore. The complex is stabilized by Ca2+ bound between two Asp-17 residues and hydrophilic groups of the ligand. The latter substitute water molecules adjacent to Ca2+ in the ligand-free pore and thus do not obstruct Ca2+ relay. The ligand-binding site is proximal to a hydrophobic bracelet of Ile-10 residues whose rotation is sterically hindered. In some conformations, the bracelet is narrow enough to block the permeation of the hydrated Ca2+ ions. The bracelet may thus act as a "gate" in SCP. Nifedipine and (R)-Bay K 8644, which act as blockers of the SCP, extend a side-chain hydrophobic moiety toward the Ile-10 residues. This would stabilize the pore-closing conformation of the gate. In contrast, the channel activator (S)-Bay K 8644 exposes a hydrophilic moiety toward the Ile-10 residues, thus destabilizing the pore-closing conformation of the gate.  相似文献   

2.
The dihydropyridine (DHP)-binding site has been identified within L-type Ca(2+) channel alpha(1C) subunit. However, the molecular mechanism underlying modulation of Ca(2+) channel gating by DHPs has not been clarified. To search for novel determinants of high affinity DHP binding, we introduced point mutations in the rat brain Ca(2+) channel alpha(1C) subunit (rbCII or Ca(v)1.2c) based on the comparison of amino acid sequences between rbCII and the ascidian L-type Ca(2+) channel alpha(1) subunit, which is insensitive to DHPs. The alpha(1C) mutants (S1115A, S1146A, and A1420S) and rbCII were transiently expressed in BHK6 cells with beta(1a) and alpha(2)/delta subunits. The mutation did not affect the electrophysiological properties of the Ca(2+) channel, or the voltage- and concentration-dependent block of Ca(2+) channel currents produced by diltiazem and verapamil. However, the S1115A channel was significantly less sensitive to DHP antagonists. Interestingly, in the S1115A channel, DHP agonists failed to enhance whole-cell Ca(2+) channel currents and the prolongation of mean open time, as well as the increment of NP(o). Responsiveness to the non-DHP agonist FPL-64176 was also markedly reduced in the S1115A channel. When S1115 was replaced by other amino acids (S1115D, S1115T, or S1115V), only S1115T was slightly sensitive to S-(-)-Bay K 8644. These results indicate that the hydroxyl group of Ser(1115) in IIIS5-S6 linker of the L-type Ca(2+) channel alpha(1C) subunit plays a critical role in DHP binding and in the action of DHP Ca(2+) channel agonists.  相似文献   

3.
Invertebrate L-type calcium channel, LCa(v) 1, isolated from the pond snail Lymnaea stagnalis is nearly indistinguishable from mammalian Ca(v) 1.2 (α1C) calcium channel in biophysical characteristics observed in vitro. These L-type channels are likely constrained within a narrow range of biophysical parameters to perform similar functions in the snail and mammalian cardiovascular systems. What distinguishes snail and mammalian L-type channels is a difference in dihydropyridine sensitivity: 100 nM isradipine exhibits a significant block of mammalian Ca(v) 1.2 currents without effect on snail LCa(v)1 currents. The native snail channel serves as a valuable surrogate for validating key residue differences identified from previous experimental and molecular modeling work. As predicted, three residue changes in LCa(v)1 (N_3o18, F_3i10, and I_4i12) replaced with DHP-sensing residues in respective positions of Ca(v) 1.2, (Q_3o18, Y_3i10, and M_4i12) raises the potency of isradipine block of LCa(v)1 channels to that of mammalian Ca(v) 1.2. Interestingly, the single N_3o18_Q mutation in LCa(v) 1 channels lowers DHP sensitivity even further and the triple mutation bearing enhanced isradipine sensitivity, still retains a reduced potency of agonist, (S)-Bay K8644.  相似文献   

4.
Prior observations have raised the possibility that dihydropyridine (DHP) agonists directly affect the sarcoplasmic reticulum (SR) cardiac Ca(2+) release channel [i.e., ryanodine receptor (RyR)]. In single-channel recordings of purified canine cardiac RyR, both DHP agonists (-)-BAY K 8644 and (+)-SDZ202-791 increased the open probability of the RyR when added to the cytoplasmic face of the channel. Importantly, the DHP antagonists nifedipine and (-)-SDZ202-791 had no competitive blocking effects either alone or after channel activation with agonist. Thus there is a stereospecific effect of SDZ202-791, such that the agonist activates the channel, whereas the antagonist has little effect on channel activity. Further experiments showed that DHP agonists changed RyR activation by suppressing Ca(2+)-induced inactivation of the channel. We concluded that DHP agonists can also influence RyR single-channel activity directly at a unique allosteric site located on the cytoplasmic face of the channel. Similar results were obtained in human purified cardiac RyR. An implication of these data is that RyR activation by DHP agonists is likely to cause a loss of Ca(2+) from the SR and to contribute to the negative inotropic effects of these agents reported by other investigators. Our results support this notion that the negative inotropic effects of DHP agonists result in part from direct alteration in the activity of RyRs.  相似文献   

5.
Two fragments of the C-terminal tail of the alpha(1) subunit (CT1, amino acids 1538-1692 and CT2, amino acids 1596-1692) of human cardiac L-type calcium channel (Ca(V)1.2) have been expressed, refolded, and purified. A single Ca(2+)-calmodulin binds to each fragment, and this interaction with Ca(2+)-calmodulin is required for proper folding of the fragment. Ca(2+)-calmodulin, bound to these fragments, is in a more extended conformation than calmodulin bound to a synthetic peptide representing the IQ motif, suggesting that either the conformation of the IQ sequence is different in the context of the longer fragment, or other sequences within CT2 contribute to the binding of calmodulin. NMR amide chemical shift perturbation mapping shows the backbone conformation of calmodulin is nearly identical when bound to CT1 and CT2, suggesting that amino acids 1538-1595 do not contribute to or alter calmodulin binding to amino acids 1596-1692 of Ca(V)1.2. The interaction with CT2 produces the greatest changes in the backbone amides of hydrophobic residues in the N-lobe and hydrophilic residues in the C-lobe of calmodulin and has a greater effect on residues located in Ca(2+) binding loops I and II in the N-lobe relative to loops III and IV in the C-lobe. In conclusion, Ca(2+)-calmodulin assumes a novel conformation when part of a complex with the C-terminal tail of the Ca(V)1.2 alpha(1) subunit that is not duplicated by synthetic peptides corresponding to the putative binding motifs.  相似文献   

6.
1,4-Dihydropyridines (DHPs) constitute a major class of ligands for L-type Ca2+ channels (LTCC). The DHPs have a boat-like, six-membered ring with an NH group at the stern, an aromatic moiety at the bow, and substituents at the port and starboard sides. Various DHPs exhibit antagonistic or agonistic activities, which were previously explained as stabilization or destabilization, respectively, of the closed activation gate by the portside substituents. Here we report a novel structural model in which agonist and antagonist activities are determined by different parts of the DHP molecule and have different mechanisms. In our model, which is based on Monte Carlo minimizations of DHP-LTCC complexes, the DHP moieties at the stern, bow, and starboard form H-bonds with side chains of the key DHP-sensing residues Tyr_IIIS6, Tyr_IVS6, and Gln_IIIS5, respectively. We propose that these H-bonds, which are common for agonists and antagonists, stabilize the LTCC conformation with the open activation gate. This explains why both agonists and antagonists increase probability of the long lasting channel openings and why even partial disruption of the contacts eliminates the agonistic action. In our model, the portside approaches the selectivity filter. Hydrophobic portside of antagonists may induce long lasting channel closings by destabilizing Ca2+ binding to the selectivity filter glutamates. Agonists have either hydrophilic substituents or a hydrogen atom at their portside, and thus lack this destabilizing effect. The predicted orientation of the DHP core allows accommodation of long substituents in the domain interface or in the inner pore. Our model may be useful for developing novel clinically relevant LTCC blockers.1,4-Dihydropyridines (DHPs)2 form a major class of L-type Ca2+ channel (LTCC) ligands. DHPs can operate as agonists or antagonists depending on their chemical structure. Importantly, activity of some DHP derivatives may shift from agonism to antagonism (and vice versa) upon site-specific mutation of the channel or modified experimental conditions (for reviews, see Refs. 1 and 2). It has been the dual nature of DHP activity (agonism and antagonism), which has made it challenging for interpretation in structural terms. The other major classes of LTCC ligands, the phenylalkylamines and benzothiazepines, are strictly antagonists.Despite a number of studies, the molecular mechanism for the activity of DHPs on LTCC remains unclear. In the present work we have addressed this problem by combining a molecular modeling approach with analyses of relevant published data. We employed a method of multiple Monte Carlo energy minimizations for docking various DHPs in our earlier reported homology model of LTCC (3), which is based on the crystal structure of the KvAP K+ channel (4). We constrained our analyses to potential ligand-binding modes that would be consistent with the features of the ligand-channel activity relationship described in published experiments. These include the structure-activity relationship of the many derivatives of DHPs and the results of mutational analyses of the DHP binding site. These experiments were mostly patch clamp electrophysiology and in vitro binding studies with LTCC heterologously expressed in Xenopus oocytes or human cell lines.The biophysical features of Ca2+ channels are important for understanding the action of DHPs. Ca2+ channels operate in three gating modes: (i) an inactivated mode from which channels do not open upon depolarization, (ii) a depolarization-elicited mode with multiple short openings, and (iii) a naturally occurring, but usually infrequent, long openings mode. Hess and coauthors (5) reported that DHP agonists stabilized long openings, whereas DHP antagonists promoted inactivated modes. In many respects, the separation of DHP agonists and antagonists is not so clear. DHP antagonists can exhibit agonist-like features, by increasing the percentage of available channels in the long opening mode (5). DHPs can also act as either agonists or antagonists under different experimental conditions (68). The effect of DHPs is Ca2+-dependent. It has been proposed that DHP antagonists bind to and stabilize a non-conducting channel state in which the selectivity filter is occupied by a single Ca2+ ion. Binding of a second Ca2+ ion is considered to destabilize DHP binding (911).The structure of DHPs can be described as a flattened-boat six-membered ring with the NH group at the stern, an aromatic moiety at the bow, and various substituents at the port and starboard sides (Fig. 1). Experimental data reveals that the agonistic or antagonistic action is primarily determined by the nature of the portside group in the ortho position of the DHP ring relative to the bowsprit. Hydrophobic groups such as COOMe promote an antagonistic effect, whereas hydrophilic groups like NO2 promote an agonistic effect (12). Intriguingly, enantiomers of some DHPs, e.g. (R)- and (S)-Bay k 8644 demonstrate opposite, antagonistic and agonistic effects on LTCC (13, 14).Open in a separate windowFIGURE 1.Chemical structures of DHPs. Some compounds are named with the prefix GS that stands for Goldmann and Stoltefuss, the authors of a fundamental review on structure-activity of DHPs (12), and the compound number as it appears in the review. Abbreviation BK-10 includes initials of the first and last authors of Ref. 18, and the length of an oligomethylene linker between the DHP core and trimethylammonium group.The access of DHPs to LTCCs has been studied by several groups (1517) and the consensus is that DHPs reach their binding site within the pore-forming α1-subunit from the extracellular side. A series of DHP derivatives with a permanently charged ammonium group was used to estimate the distance of the DHP binding site from the extracellular surface of the membrane (18). The optimal potency was found with an ammonium group linked to the dihydropyridine ring via a decamethylene chain.The DHP binding site has been outlined to the interface between repeats III and IV of the pore-forming α1-subunit of LTCC using antibody mapping of proteolytically labeled channel fragments and a series of subsequent studies with chimeras and site-specific mutations (1929). Key DHP-sensing residues were defined in transmembrane segments IIIS5, IIIS6, and IVS6 and in the pore helix IIIP. Using the x-ray structure of K+ channels as a guide, the corresponding DHP residues appear tightly spaced at the interface of the III/IV domain in the expected three-dimensional structure of LTCC.The available experimental data are insufficient to elaborate structural models of DHP-LTCC complexes. An experimental structure of such a complex would be of great importance, but x-ray structures of voltage-gated Ca2+ channels are unavailable. Crystallographic studies of DHP-LTCC complexes are even more challenging. Indeed, many small-molecule ligands of K+ channels are known, but only a few ligands have so far been co-crystallized with KcsA. In these circumstances, molecular modeling remains the only feasible approach to provide insights into atomic details of ligand-channel interactions such as positions and orientations of DHP ligands and determinants for their agonistic and antagonistic actions. Predictions from a modeling study remain hypothetical until they are experimentally confirmed, but hypotheses that explain numerous experimental observations stimulate further experimental studies.To date three models of DHP-bound LTCC have been published (3032). Despite differences in the proposed location of ligands and patterns of the ligand-receptor interactions, all three models suggest a similar molecular mechanism for DHP action. The portside group in these models is oriented toward the ring of hydrophobic residues in the C-terminal halves of S6 helices, close to the proposed activation gate in LTCC. Hydrophobic groups at the portside of the DHP antagonists are proposed to stabilize the closed state of the activation gate, whereas hydrophilic groups of agonists are considered to destabilize the closed state (or stabilize the open state). A primary weakness of these models is that agonistic and antagonistic activities of DHPs are considered to be caused by opposite (stabilizing and destabilizing) effects on the same molecular target (near the activation gate). This “single target” model does not adequately account for the wide variety of observed behaviors of DHP derivatives in published experiments.We have elaborated an alternative model in our present work that suggests different molecular targets and different mechanisms for agonistic and antagonistic activities of DHPs. As published experiments have revealed, the agonistic or antagonistic activity of a DHP ligand not only depends on its chemical structure, but also is sensitive to the experimental conditions and the structure of the drug target. Small changes to the structure of LTCC (as revealed by the behaviors of DHPs in chimeric and mutagenized LTCC) can shift a DHP agonist into an antagonist. The model that explains these behaviors is one where atomic determinants for both agonist and antagonist capacities are present within a single DHP molecule. Manifestation of these capacities depends on structural peculiarities of the DHP ligand and the DHP receptor, as well as on the ligand-receptor orientation that can be sensitive to experimental conditions.  相似文献   

7.
Syntaxin 1A binds to and inhibits epithelial cystic fibrosis transmembrane conductance regulator (CFTR) Cl(-) channels and synaptic Ca(2+) channels in addition to participating in SNARE complex assembly and membrane fusion. We exploited the isoform-specific nature of the interaction between syntaxin 1A and CFTR to identify residues in the H3 domain of this SNARE (SNARE motif) that influence CFTR binding and regulation. Mutating isoform-specific residues that map to the surface of syntaxin 1A in the SNARE complex led to the identification of two sets of hydrophilic residues that are important for binding to and regulating CFTR channels or for binding to the syntaxin regulatory protein Munc-18a. None of these mutations affected syntaxin 1A binding to other SNAREs or the assembly and stability of SNARE complexes in vitro. Conversely, the syntaxin 1A-CFTR interaction was unaffected by mutating hydrophobic residues in the H3 domain that influence SNARE complex stability and Ca(2+) channel regulation. Thus, CFTR channel regulation by syntaxin 1A involves hydrophilic interactions that are mechanistically distinct from the hydrophobic interactions that mediate SNARE complex formation and Ca(2+) channel regulation by this t-SNARE.  相似文献   

8.
KcsA is a homotetrameric 68-kDa membrane-associated potassium channel which selectively gates the flux of potassium ions across the membrane. The channel is known to undergo a pH-dependent open-to-closed transition. Here we describe an NMR study of the monomeric subunit of the channel (KcsAM), solubilized in SDS micelles. Chemical shift, solvent exchange, backbone 15N relaxation and residual dipolar coupling (RDC) data show the TM1 helix to remain intact, but the TM2 helix contains a distinct kink, which is subject to concentration-independent but pH-dependent conformational exchange on a microsecond time scale. The kink region, centered at G99, was previously implicated in the gating of the tetrameric KcsA channel. An RDC-based model of KcsAM at acidic pH orients TM1 and the two helical segments of the kinked TM2 in a configuration reminiscent of the open conformation of the channel. Thus, the transition between states appears to be an inherent capability of the monomer, with the tetrameric assembly exerting a modulatory effect upon the transition which gives the channel its physiological gating profile.  相似文献   

9.
Potassium channels fluctuate between closed and open states. The detailed mechanism of the conformational changes opening the intracellular pore in the K+ channel from Streptomyces lividans (KcsA) is unknown. Applying Monte Carlo normal mode following, we find that gating involves rotation and unwinding of the TM2 bundle, lateral movement of the TM2 helices away from the channel axis, and disappearance of the TM2 bundle. The open-state conformation of KcsA exhibits a very wide inner vestibule, with a radius approximately 5-7 A and inner helices bent at the A98-G99 hinge. Computed conformational changes demonstrate that spin labeling and X-ray experiments illuminate different stages in gating: transition begins with clockwise rotation of the TM2 helices ending at a final state with the TM2 bend hinged near residues A98-G99. The concordance between the computational and experimental results provides atomic-level insights into the structural rearrangements of the channel's inner pore.  相似文献   

10.
KcsA is a homotetrameric 68-kDa membrane-associated potassium channel which selectively gates the flux of potassium ions across the membrane. The channel is known to undergo a pH-dependent open-to-closed transition. Here we describe an NMR study of the monomeric subunit of the channel (KcsAM), solubilized in SDS micelles. Chemical shift, solvent exchange, backbone 15N relaxation and residual dipolar coupling (RDC) data show the TM1 helix to remain intact, but the TM2 helix contains a distinct kink, which is subject to concentration-independent but pH-dependent conformational exchange on a microsecond time scale. The kink region, centered at G99, was previously implicated in the gating of the tetrameric KcsA channel. An RDC-based model of KcsAM at acidic pH orients TM1 and the two helical segments of the kinked TM2 in a configuration reminiscent of the open conformation of the channel. Thus, the transition between states appears to be an inherent capability of the monomer, with the tetrameric assembly exerting a modulatory effect upon the transition which gives the channel its physiological gating profile.  相似文献   

11.
Using the bacterial K+ channel KcsA as a template, we constructed models of the pore region of the cardiac ryanodine receptor channel (RyR2) monomer and tetramer. Physicochemical characteristics of the RyR2 model monomer were compared with the template, including homology, predicted secondary structure, surface area, hydrophobicity, and electrostatic potential. Values were comparable with those of KcsA. Monomers of the RyR2 model were minimized and assembled into a tetramer that was, in turn, minimized. The assembled tetramer adopts a structure equivalent to that of KcsA with a central pore. Characteristics of the RyR2 model tetramer were compared with the KcsA template, including average empirical energy, strain energy, solvation free energy, solvent accessibility, and hydrophobic, polar, acid, and base moments. Again, values for the model and template were comparable. The pores of KcsA and RyR2 have a common motif with a hydrophobic channel that becomes polar at both entrances. Quantitative comparisons indicate that the assembled structure provides a plausible model for the pore of RyR2. Movement of Ca2+, K+, and tetraethylammonium (TEA+) through the model RyR2 pore were simulated with explicit solvation. These simulations suggest that the model RyR2 pore is permeable to Ca2+ and K+ with rates of translocation greater for K+. In contrast, simulations indicate that tetraethylammonium blocks movement of metal cations.  相似文献   

12.
Conformations of proline residues in membrane environments   总被引:3,自引:0,他引:3  
Although noted as hydrophilic residues with helix-breaking potential, proline residues are observed in putatively alpha-helical transmembrane (TM) segments of many channel-forming integral membrane proteins. In addition to the recognized property of X-Pro peptide bonds (where X = any amino acid) to occur in cis as well as trans isomeric states, the tertiary amide character of the X-Pro bond confers increased propensity for involvement of its carbonyl group in specific H-bonded structures (e.g., beta- and gamma-turns) and/or liganding interactions with positively charged species. To examine this latter situation in further detail, we identified Leu-Pro-Phe as a consensus sequence triad based on actual occurrences of intramembranous Pro residues in transport protein TM segments. Accordingly, we have undertaken the synthesis of hydrophobic peptides with potential membrane affinity, of which t-butyloxycarbonyl-L-Ala-L-Ala-L-Ala-L-Leu-L-Pro-L-Phe-OH (t-Boc-AAALPF-OH) is an initial compound. Partitioning of this peptide into model membrane environments composed of lipid micelles induces specific conformation(s) for the membrane-bound hexapeptide, as monitored by 75-MHz 13C-nmr spectral behavior of 13C-enriched Leu and Pro carbonyl carbons, and by 300-MHz 1H-nmr spectra of peptide alpha, beta, and aromatic protons. Data are interpreted in terms of an intramolecularly H-bonded inverse gamma-turn conformation in the membrane environment involving the Leu-Pro-Phe triad. The inherent structural instability of a Pro-containing segment in a TM helix due to the multiplicity of possible local conformations is discussed as a functional aspect of membrane-buried prolines in transport proteins.  相似文献   

13.
Missense mutations in presenilin 1 (PS1) and presenilin 2 (PS2) proteins are a major cause of familial Alzheimer disease. Presenilins are proteins with nine transmembrane (TM) domains that function as catalytic subunits of the γ-secretase complex responsible for the cleavage of the amyloid precursor protein and other type I transmembrane proteins. The water-filled cavity within presenilin is necessary to mediate the intramembrane proteolysis reaction. Consistent with this idea, cysteine-scanning mutagenesis and NMR studies revealed a number of water-accessible residues within TM7 and TM9 of mouse PS1. In addition to γ-secretase function, presenilins also demonstrate a low conductance endoplasmic reticulum Ca(2+) leak function, and many familial Alzheimer disease presenilin mutations impair this function. To map the potential Ca(2+) conductance pore in PS1, we systematically evaluated endoplasmic reticulum Ca(2+) leak activity supported by a series of cysteine point mutants in TM6, TM7, and TM9 of mouse PS1. The results indicate that TM7 and TM9, but not TM6, could play an important role in forming the conductance pore of PS1. These results are consistent with previous cysteine-scanning mutagenesis and NMR analyses of PS1 and provide further support for our hypothesis that the hydrophilic catalytic cavity of presenilins may also constitute a Ca(2+) conductance pore.  相似文献   

14.
The role of the inactivated channel conformation in the molecular mechanism of Ca(2+) channel block by the 1,4-dihydropyridine (DHP) (+)-isradipine was analyzed in L-type channel constructs (alpha(1Lc); Berjukow, S., Gapp, F., Aczel, S., Sinnegger, M. J., Mitterdorfer, J., Glossmann, H., and Hering, S. (1999) J. Biol. Chem. 274, 6154-6160) and a DHP-sensitive class A Ca(2+) channel mutant (alpha(1A-DHP); Sinnegger, M. J., Wang, Z., Grabner, M., Hering, S., Striessnig, J., Glossmann, H., and Mitterdorfer, J. (1997) J. Biol. Chem. 272, 27686-27693) carrying the high affinity determinants of the DHP receptor site but inactivating at different rates. Ca(2+) channel inactivation was modulated by coexpressing the alpha(1A-DHP)- or alpha(1Lc)-subunits in Xenopus oocytes with either the beta(2a)- or the beta(1a)-subunit and amino acid substitutions in L-type segment IVS6 (I1497A, I1498A, and V1504A). Contrary to a modulated receptor mechanism assuming high affinity DHP binding to the inactivated state we observed no clear correlation between steady state inactivation and Ca(2+) channel block by (+)-isradipine: (i) a 3-fold larger fraction of alpha(1A-DHP)/beta(1a) channels in steady state inactivation at -80 mV (compared with alpha(1A-DHP)/beta(2a)) did not enhance the block by (+)-isradipine; (ii) different steady state inactivation of alpha(1Lc) mutants at -30 mV did not correlate with voltage-dependent channel block; and (iii) the midpoint-voltages of the inactivation curves of slowly inactivating L-type constructs and more rapidly inactivating alpha(1Lc)/beta(1a) channels were shifted to a comparable extent to more hyperpolarized voltages. A kinetic analysis of (+)-isradipine interaction with different L-type channel constructs revealed a drug-induced inactivated state. Entry and recovery from drug-induced inactivation are modulated by intrinsic inactivation determinants, suggesting a synergism between intrinsic inactivation and DHP block.  相似文献   

15.
Dihydropyridine (DHP) Ca2+ channel modulators were used to explore the relationship between voltage-gated Ca2+ channels and PRL secretion, synthesis, and mRNA in PRL-secreting pituitary cells. Optical isomers of the Ca2+ channel agonist Bay K 8644 produced stereospecific and opposing effects on L-type Ca2+ current, PRL release, and synthesis in GH3 and GH4C1 cells. (-)-Bay K 8644 (R5417) behaved as a pure agonist, enhancing Ca2+ current several-fold while shifting the current-voltage curve 10-15 mV in the hyperpolarizing direction. The agonist effect was independent of holding potential, but decreased during prolonged Ba2+ or Ca2+ entry. R5417 produced a concentration-dependent increase in acute PRL release and enhanced PRL production by GH cells several-fold during a 72-h period. (+)-Bay K 8644 (R4407) behaved as a weak Ca2+ channel antagonist, inhibiting L-type Ca2+ current, KCl-stimulated PRL secretion, and PRL production at concentrations of 0.5-5 microM. These two isomers produced similar effects on PRL production by normal rat pituitary cells in dispersed culture. R5417 (500 nM) increased PRL produced in 72 h to 233 +/- 8% of the control value. R4407 reduced this quantity by 36 +/- 9%. The effects of the DHPs on PRL mRNA levels were consistent with the effects observed for acute secretion and hormone production. The agonist R5417 increased PRL mRNA 147 +/- 5% over a 30-h period, and the potent DHP Ca2+ channel blocker nimodipine inhibited PRL mRNA production 2-fold. These results demonstrate that racemic Bay K 8644 interacts with L-type Ca2+ channels in normal and transformed pituitary cells as a mixed agonist-antagonist.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

16.
Cysteine-scanning mutagenesis (SCAM) and computer-based modeling were used to investigate key structural features of the S6 transmembrane segment of the calcium-activated K(+) channel of intermediate conductance IKCa. Our SCAM results show that the interaction of [2-(trimethylammonium)ethyl] methanethiosulfonate bromide (MTSET) with cysteines engineered at positions 275, 278, and 282 leads to current inhibition. This effect was state dependent as MTSET appeared less effective at inhibiting IKCa in the closed (zero Ca(2+) conditions) than open state configuration. Our results also indicate that the last four residues in S6, from A283 to A286, are entirely exposed to water in open IKCa channels, whereas MTSET can still reach the 283C and 286C residues with IKCa maintained in a closed state configuration. Notably, the internal application of MTSET or sodium (2-sulfonatoethyl) methanethiosulfonate (MTSES) caused a strong Ca(2+)-dependent stimulation of the A283C, V285C, and A286C currents. However, in contrast to the wild-type IKCa, the MTSET-stimulated A283C and A286C currents appeared to be TEA insensitive, indicating that the MTSET binding at positions 283 and 286 impaired the access of TEA to the channel pore. Three-dimensional structural data were next generated through homology modeling using the KcsA structure as template. In accordance with the SCAM results, the three-dimensional models predict that the V275, T278, and V282 residues should be lining the channel pore. However, the pore dimensions derived for the A283-A286 region cannot account for the MTSET effect on the closed A283C and A286 mutants. Our results suggest that the S6 domain extending from V275 to V282 possesses features corresponding to the inner cavity region of KcsA, and that the COOH terminus end of S6, from A283 to A286, is more flexible than predicted on the basis of the closed KcsA crystallographic structure alone. According to this model, closure by the gate should occur at a point located between the T278 and V282 residues.  相似文献   

17.
Invertebrate L-type calcium channel, LCav1, isolated from the pond snail Lymnaea stagnalis is nearly indistinguishable from mammalian Cav1.2 (α1C) calcium channel in biophysical characteristics observed in vitro. These L-type channels are likely constrained within a narrow range of biophysical parameters to perform similar functions in the snail and mammalian cardiovascular systems. What distinguishes snail and mammalian L-type channels is a difference in dihydropyridine sensitivity: 100 nM isradipine exhibits a significant block of mammalian Cav1.2 currents without effect on snail LCav1 currents. The native snail channel serves as a valuable surrogate for validating key residue differences identified from previous experimental and molecular modeling work. As predicted, three residue changes in LCav1 (N_3o18, F_3i10, and I_4i12) replaced with DHP-sensing residues in respective positions of Cav1.2, (Q_3o18, Y_3i10, and M_4i12) raises the potency of isradipine block of LCaV1 channels to that of mammalian Cav1.2. Interestingly, the single N_3o18_Q mutation in LCav1 channels lowers DHP sensitivity even further and the triple mutation bearing enhanced isradipine sensitivity, still retains a reduced potency of agonist, (S)-Bay K8644.  相似文献   

18.
Receptor-mediated Ca(2+) release from the endoplasmic reticulum (ER) is often followed by Ca(2+) entry through Ca(2+)-release-activated Ca(2+) (CRAC) channels in the plasma membrane . RNAi screens have identified STIM1 as the putative ER Ca(2+) sensor and CRACM1 (Orai1; ) as the putative store-operated Ca(2+) channel. Overexpression of both proteins is required to reconstitute CRAC currents (I(CRAC); ). We show here that CRACM1 forms multimeric assemblies that bind STIM1 and that acidic residues in the transmembrane (TM) and extracellular domains of CRACM1 contribute to the ionic selectivity of the CRAC-channel pore. Replacement of the conserved glutamate in position 106 of the first TM domain of CRACM1 with glutamine (E106Q) acts as a dominant-negative protein, and substitution with aspartate (E106D) enhances Na(+), Ba(2+), and Sr(2+) permeation relative to Ca(2+). Mutating E190Q in TM3 also affects channel selectivity, suggesting that glutamate residues in both TM1 and TM3 face the lumen of the pore. Furthermore, mutating a putative Ca(2+) binding site in the first extracellular loop of CRACM1 (D110/112A) enhances monovalent cation permeation, suggesting that these residues too contribute to the coordination of Ca(2+) ions to the pore. Our data provide unequivocal evidence that CRACM1 multimers form the Ca(2+)-selective CRAC-channel pore.  相似文献   

19.
Structure of the Ca channel open pore is unlikely to be the same as that of the K channel because Ca channels do not contain the hinge residues Gly or Pro. The Ca channel does not have a wide entry into the inner pore, as is found in K channels. First we sought to simulate the open state of the Ca channel by modeling forced opening of the KcsA channel using a procedure of restrained minimization with distance constraints at the level of the α-helical bundle, corresponding to segments Thr-107-Val-115. This produced an intermediate open state, which was populated by amino acid residues of Ca channels and then successively optimized until the opening of the pore reached a diameter of about 10 Å, large enough to allow verapamil to enter and block the Ca channel from inside. Although this approach produced a sterically plausible structure, it was in significant disagreement with the MTSET accessibility data for single cysteine mutations of S6 segments of the P/Q channel1 that do not fit with an α-helical pattern. Last we explored the idea that the four S6 segments of Ca channels may contain intra-molecular deformations that lead to reorientation of its side chains. After introduction of ≠-bulges, the model agreed with the MTSET accessibility data. MTSET modification of a cysteine at the C-end of only one S6 could produce physical occlusion and block of the inner pore of the open Ca channel, as observed experimentally, and as expected if the pore opening is narrower than that of K channels.  相似文献   

20.
The determination of the crystal structure of the Ca(2+)-ATPase of sarcoplasmic reticulum (SR) in its Ca(2+)-bound [Nature 405 (2000) 647] and Ca(2+)-free forms [Nature 418 (2002) 605] gives the opportunity for an analysis of conformational changes on the Ca(2+)-ATPase and of helix-helix and helix-lipid interactions in the transmembrane (TM) region of the ATPase. The locations of the ends of the TM alpha-helices on the cytoplasmic side of the membrane are reasonably well defined by the location of Trp residues and by the location of Lys-262 that snorkels up to the surface. The locations of the lumenal ends of the helices are less clear. The position of Lys-972 on the lumenal side of helix M9 suggests that the hydrophobic thickness of the protein is only about 21 A, rather than the normal 30 A. The experimentally determined TM alpha-helices do not agree well with those predicted theoretically. Charged headgroups are required for strong interaction of lipids with the ATPase, consistent with the large number of charged residues located close to the lipid-water interface. Helix packing appears to be rather irregular. Packing of helices M8 and M10 is of the 3-4 ridges-into-grooves or knobs-into-holes types. Packing of helices M5 and M7 involves two Gly residues in M7 and one Gly residue in M5. Packing of the other helices generally involves just one or two residues on each helix at the crossing point. The irregular packing of the TM alpha-helices in the Ca(2+)-ATPase, combined with the diffuse structure of the ATPase on the lumenal side of the membrane, is suggested to lead to a relative low activation energy for changing the packing of the TM alpha-helices, with changes in TM alpha-helical packing being important in the process of transfer of Ca(2+) ions across the membrane. The inhibitor thapsigargin binds in a cleft between TM alpha-helices M3, M5 and M7. It is suggested that this and other similar clefts provide binding sites for a variety of hydrophobic molecules affecting the activity of the Ca(2+)-ATPase.  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