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 inside-out patch-clamp technique was applied to the plasmolyzedplasmalemma of inter-nodes of Chara corallina without enzymatictreatment. We found two different types of channel activitythat were CP-sensitive. Both types of channel were Ca2+-dependent.However, the one that exhibited greater dependence on Ca2+ ionswas the focus of our studies, and we named it the Ca2+-dependentCP-sensitive anion channel. When the concentration ofCa2+ ions on the cyto-plasmic side was 1.0 µM, the Ca2+-dependentCP-sensitive channel opened most frequently between approximately–80 and –100 mV. At 10 µM Ca2+, it openedless frequently, and at 0.1 µM Ca2+ it scarcely openedat all. These observations indicate that the anion channel ofinterest is voltage-dependent over a restricted range of concentrationsof Ca2+ ions. The dependence on Ca2+ and voltage of the channelcan explain the behavior of the excitable Ca2+-activated Clchannel in the Chara plasmalemma. The channel activity was blockedby several antagonists of calmodulin. 4 Present Address: Department of Biology, College of GeneralEducation, Osaka University, Toyonaka, 560 Osaka, Japan (Received October 8, 1990; Accepted April 4, 1991)  相似文献   

2.
Internodal cells of Chara australis were subjected to two consecutiveintracellular perfusions with a Ca2+-free EGTA medium whichdisintegrated the tonoplast within about 10 minutes and thenwith a Ca2+-buffered medium. All perfusion media usually contained1 mM ATP. To stop the electrogenic pump, the internode was depletedof intracellular ATP. The excitability of the plasmalemma wasnot significantly influenced by intracellular free Ca2+ concentrationsup to 10–4 M. To trigger action potentials, minimum currentdensities of 1 to 2 µA cm–2 had to be applied atall tested Ca2+ concentrations. In the absence of cytoplasmicATP, excitability was completely lost at all Ca2+ concentrations. 1 Present address: Botanisches Institut der Universit?t Bonn,Venusbergweg 22, D-5300 Bonn, FRG. (Received September 22, 1984; Accepted March 6, 1985)  相似文献   

3.
The mechanism of the Ca2+-dependent Cl efflux was studiedin tonoplast-free cells, in which the intracellular chemicalcomposition can be freely controlled. Tonoplast-free cells wereprepared by perfusing the cell interior of internodal cellsof Chara corallina with a medium that contained EGTA. The Ca2+-inducedCl efflux was measured together with the membrane potentialduring continuous intracellular perfusion. The dependenciesof Cl efflux and the membrane potential on the intracellularCa2+ or Cl concentrations were analyzed. When perfusionwas started with medium that contained Ca2+ ions, Clefflux and membrane depolarization were induced. The amountof Cl efflux varied considerably among individual cells.The rate of efflux decreased exponentially but a residual effluxremained detectable. The Cl efflux was induced at concentrationsof Ca2+ ions above 1 µM and reached a maximum at 1 mM.By contrast, the membrane depolarization reached a maximum atabout 10 µM Ca2+. The rate of Cl efflux increasedlinearly with logarithmic increases in the intracellular Clconcentrations. These findings suggest that more than two kindsof Ca2+-dependent Cl channel might be present in theplasma membrane. Addition of ATP or its removal from the perfusion medium didnot affect the Ca2+-dependent Cl efflux. Calmodulin antagonistsslightly inhibited the Ca2+-dependent Cl efflux. 1Present address: Biological Laboratory, Hitotsubashi University,Naka 2-1, Kunitachi, Tokyo, 186 Japan.  相似文献   

4.
Effects of cytoplasmic Ca2+ on the electrical properties ofthe plasma membrane were investigated in tonoplast-free cellsof Chara australis that had been internally perfused with media,containing either 1 mM ATP to fuel the electrogenic pump orhexokinase and glucose to deplete the ATP and stop the pump. In the presence of ATP, cytoplasmic Ca2+ up to 2.5?10–5M did not affect the membrane potential (about -190 mV), butmembrane resistance decreased uniformly with increasing [Ca2+]i.In the absence of ATP, the membrane potential, which was onlyabout -110 mV, was depolarized further by raising [Ca2+]i from1.4?10–6 to 2.5?10–5 M. Membrane resistance, whichwas nearly the twofold that of ATP-provided cells, decreasedmarkedly with an increase in [Ca2+]i from zero to 1.38?10–6M, but showed no change for further increases. Internodal cellsof Nitellopsis obtusa were more sensitive to intracellular Ca2+with respect to membrane potential than were those of Charaaustralis, reconfirming the results obtained by Mimura and Tazawa(1983). The effect of cytoplasmic Ca2+ on the ATP-dependent H+ effluxwas measured. No marked difference in H+ effluxes was detectedbetween zero and 2.5?10–5 M [Ca2+]i; but, at 10–4M the ATP-dependent H+ efflux was almost zero. Ca2+ efflux experimentswere done to investigate dependencies on [Ca2+]i and [ATP]i.The efflux was about 1 pmol cm–2 s–1 at all [Ca2+]iconcentrations tested (1.38?10–6, 2.5?10–5, 10–4M).This value is much higher than the influx reported by Hayamaet al. (1979), and this efflux was independent of [ATP]i. Thepossibility of a Ca2+-extruding pump is discussed. 1 Present address: Botanisches Institut der Universit?t Bonn,Venusbergweg 22, 5300 Bonn, F.R.G. (Received September 22, 1984; Accepted February 19, 1985)  相似文献   

5.
Calmodulin (CaM) is a versatile Ca2+-binding protein that regulates the activity of numerous effector proteins in response to Ca2+ signals. Several CaM-dependent regulatory mechanisms have been identified, including autoinhibitory domain displacement, sequestration of a ligand-binding site, active site reorganization, and target protein dimerization. We recently showed that the N- and C-lobes of animal and plant CaM isoforms could independently and sequentially bind to target peptides derived from the CaM-binding domain of Nicotiana tabacum mitogen-activated protein kinase phosphatase (NtMKP1), to form a 2:1 peptide:CaM complex. This suggests that CaM might facilitate the dimerization of NtMKP1, although the dimerization mechanism is distinct from the previously described simultaneous binding of other target peptides to CaM. The independent and sequential binding of the NtMKP1 peptides to CaM also suggests an alternative plausible scenario in which the C-lobe of CaM remains tethered to NtMKP1, and the N-lobe is free to recruit a second target protein to the complex, such as an NtMKP1 target. Thus, we hypothesize that CaM may be capable of functioning as a Ca2+-dependent adaptor or recruiter protein.Key Words: calmodulin, calcium, EF-hand, adaptor protein, mitogen-activated protein kinase phosphataseCalcium (Ca2+) is a dynamic secondary messenger that regulates many signaling events in both plant and animal cells. Intracellular Ca2+ transients and oscillations (Ca2+ signals) are decoded by a large superfamily of calcium-binding proteins, the most important of which is calmodulin (CaM).13 The prototypical CaM protein consists of four tandem helix-loop-helix “EF-hand” Ca2+-binding motifs that are divided into distinct N- and C-terminal globular lobes connected by a flexible linker. CaM proteins from all species including the single mammalian CaM and the many different plant CaM isoforms each undergo similar Ca2+-induced conformational changes involving a rearrangement of the position of its α-helices that opens distinct hydrophobic target protein-binding patches on the surface of each lobe; known as the “open” conformation (Fig. 1B). These hydrophobic patches can interact with numerous different target proteins including protein kinases, protein phosphatases, cytoskeletal proteins and other cell signaling enzymes, to regulate their activity. The closed or semi-open conformations adopted by the N- and C-lobes of Ca2+-free CaM (apo-CaM) (Fig. 1A) can also interact with another subset of proteins, to target CaM to certain cellular locations or facilitate Ca2+-independent regulatory events.13Open in a separate windowFigure 1Structures of CaM and CaM-target complexes. (A) apo-CaM (PDB:1DMo), (B) Ca2+-CaM (PDB:1CLL). Complexes of CaM bound to (C) CaMBD of smooth muscle myosin light chain kinase (PDB:1CDL), (D) partial CaMBD of plasma membrane Ca2+-pump C20W (PDB:1CFF), (E) the adenylyl cyclase protein from Bacillus anthracis (PDB:1K93), (F) 2 glutamate decarboxylase CaMBD''s (PDB:1NWD), (G) 2 CaM proteins bound to 2 small conductance Ca2+-activated potassium channel (SK channel) CaMBD''s (PDB:1G4Y), (H) 2 apo-CaM proteins bound to 2 tandem IQ motifs from murine myosin V (PDB:2IX7). In each panel CaM is shown in ivory, the target molecule is shown in blue and the Ca2+ ions bound to the N- and/or C-lobes of CaM are represented by red spheres.The CaM-dependent regulation of target proteins can occur through numerous different mechanisms. For example, Ca2+-CaM can relieve autoinhibition by binding to a short (20–25 residue) calmodulin-binding domain (CaMBD) sequence that is adjacent to or within an autoinhibitory region of the enzyme (Fig. 2A).3 Numerous structures of these Ca2+-CaM-CaMBD complexes have been reported, which reveal a characteristic “wrap-around” binding mode (Fig. 1C). Typically the CaM C-lobe binds with high affinity to a Trp residue within the N-terminal part of the target sequence, and the flexible central linker allows the N-lobe to pivot and bind to a second bulky hydrophobic “anchor” residue within the C-terminal part of the target sequence.3 Truncation of this second anchor residue can lead to binding of only one CaM domain and an extended CaM conformation (Fig. 1D).4,5 Studies with plant CaM isoforms having mutations to non-CaMBD-coordinating residues have also suggested that a secondary binding interface exists on the opposite surface of the CaM protein which also contributes to the activation of some of these target enzymes.6,7Open in a separate windowFigure 2Schematic model for the various mechanisms of CaM-dependent target regulation. (A) autoinhibitory domain displacement, (B) sequestering of a ligand binding site, (C) active-site reorganization, (D) CaM-induced target protein dimerization (1:2 complex), (E) CaM-induced target protein dimerization (2:2 complex), (F) hypothesized model for CaM acting as an adaptor/recruiter protein. In each panel CaM is shown as a red dumbbell shaped molecule with Ca2+ ions represented by yellow circles, and the target proteins are shown in various colors. See the text for details on each model.Another regulatory mechanism involving Ca2+-CaM-binding to a single contiguous CaMBD sequence may occur with the potato kinesin-like CaM-binding protein (KCBP)8 as well as some plant cyclic-nucleotide gated channels (CNGC''s).9 In both cases the Ca2+-CaM binding site on the target protein overlaps with the respective ligand binding site, and thus the binding of KCBP to microtubules or the binding of cyclic nucleotide monophosphates to CNGC''s may be prevented by interaction with Ca2+-CaM (Fig. 2B). In a variation on this mechanism, CaM can bind to the cytoplasmic juxtamembrane region of the human epidermal growth factor receptor and sequester a threonine residue which is a specific phosphorylation target of protein kinase C (PKC). CaM-binding inhibits PKC phosphorylation of this threonine, and PKC phosphorylation inhibits CaM-binding.10There are also several examples of CaM-target interactions where the N- and C-lobes bind to noncontiguous target protein regions, and play distinct roles in target regulation. The structures of a CaM-activated adenylyl cyclase from Bacillus anthracis with and without bound CaM shows how the N- and C-lobes of CaM can bind two distant regions of the adenylyl cyclase enzyme and induce a conformation reorganization that creates the enzyme''s active site (Figs. 1E and and2C2C).11 An interesting feature of this interaction is that the CaM N-lobe remains Ca2+-free and in a closed conformation, while the C-lobe is in a canonical Ca2+-bound open conformation. Indeed, Ca2+-binding to the C-lobe but not N-lobe is required for activation of the adenylyl cyclase.12The N- and C-lobes of Ca2+-CaM can also each simultaneously bind to identical peptides derived from the petunia glutamate decarboxylase (GAD) enzyme to form a 1:2 Ca2+-CaM:GAD complex (Fig. 1F).13,14 This suggests that Ca2+-CaM-induced target protein dimerization may be another way in which CaM can regulate target proteins (Fig. 2D). CaM-dependent dimerization has also been shown to regulate the activity of small conductance Ca2+-activated K+ channels (SK channel), although in this case a novel 2:2 CaM:SK channel complex is formed (Figs. 1G and and2E2E).15 This structure is also unique because Ca2+ is bound to the “lower affinity” N-lobe EF-hands, but not to the “higher affinity” C-lobe EF-hands of CaM.In addition to the SK channel, CaM can regulate voltage-gated sodium channels, voltage-gated calcium channels, as well as ryanodine-sensitive calcium release channels.16 With these channels CaM typically binds in complex Ca2+-dependent and Ca2+-independent ways to several noncontiguous target sequences in the same protein, and often to so-called IQ motifs (IQXXXRGXXXR). IQ motifs are generally thought to be constitutive apo-CaM binding sites which retain CaM under resting (low [Ca2+]) cellular conditions to ensure a rapid response to Ca2+-stimuli.17 However many IQ motifs can also bind specifically to Ca2+-CaM or to both apo-CaM and Ca2+-CaM. Structures of some Ca2+-CaM-IQ domain complexes have revealed wrap-around binding modes, albeit with differences in lobe and peptide orientation compared to other complexes.1820 For a discussion about the mechanisms of CaM-dependent ion channel regulation (see ref. 16). A very recent crystal structure of apo-CaM bound to an IQ domain from myosin V (Fig. 1H) has also revealed yet another variation on the wrap-around binding mode, where the apo-C-lobe of CaM adopts a semi-open conformation and forms numerous interactions with the target sequence, while the apo-N-lobe adopts a closed conformation and forms weaker interactions with the IQ domain.21Using several biophysical techniques we recently characterized the interaction between CaM isoforms (mammalian CaM, soybean CaM isoforms SCaM-1 and SCaM-4) and a novel CaMBD derived from the Nicotiana tabacum mitogen-activated protein kinase phosphatase (NtMKP1).22 The NtMKP1 protein was initially identified as a CaM-binding protein by Ohashi and coworkers,23 and the same group recently showed that CaM-binding NtMKP1 homologs are also present in other plant species as well.24 We found that each CaM isoform was capable of binding to the NtMKP1 CaMBD in the absence of Ca2+ using only the apo-C-lobe, with the primary binding site consisting of NtMKP1 residues N438 - S449, and additional C-terminal residues G450 - K460 enhancing the overall binding affinity (Kd ∼10−5 M). In the presence of Ca2+, a 1:1 complex could be formed with the CaM C-lobe having significantly increased affinity for the N438 - S449 region of NtMKP1 (Kd 10−7 − 10−10 M). However, the Ca2+-loaded CaM N-lobe interacted only very weakly with the C-terminal NtMKP1 sequence in this 1:1 complex, despite an abundance of seemingly suitable hydrophobic “anchor” residues in this region. Interestingly, the addition of more peptide triggered the independent binding of a second NtMKP1 peptide to the Ca2+-CaM N-lobe (Kd 10−5 − 10−6 M) to form a 1:2 Ca2+-CaM:NtMKP1 complex. As with GAD, these results suggest that CaM is capable of facilitating the dimerization of NtMKP1, although the independent and sequential NtMKP1 peptide binding to the C- and N-lobes markedly distinguishes the CaM-NtMKP1 interaction from the simultaneous high-affinity binding of 2 GAD CaMBD''s to CaM.While our NtMKP1 study was ongoing, Ohashi and coworkers reported that CaM is incapable of stimulating the phosphatase activity of the NtMKP1 enzyme, thereby implying that the CaM-NtMKP1 interaction is necessary for something other than direct enzyme regulation.25 The independent and sequential binding of the NtMKP1 fragments to the Ca2+-saturated C- and then N-lobes of CaM observed in our study suggests a plausible situation in which the C-lobe of CaM is tightly bound to NtMKP1, leaving the N-lobe free to recruit a different target protein to the complex, for example, a NtMKP1 protein substrate. Therefore, CaM may be capable of acting as an adaptor or recruiter protein, which would add yet another mechanism of target regulation to CaM''s repertoire (Fig. 2F). In addition to NtMKP1 peptides, the isolated N-lobe of CaM is capable of binding to other CaMBD peptides26,27 as well as intact target proteins,28 increasing the likelihood that the N-lobe could serve as a recruiter domain. The pre-association of the apo-C-lobe of CaM with NtMKP1 under resting conditions would also ensure a rapid response response to Ca2+-stimuli, since CaM would only need to recruit one rather than both protein targets.Although the ability of CaM to act as an adaptor protein in vivo has not yet been demonstrated, there are examples of related EF-hand proteins acting as adaptor proteins, including centrin29 and calcium- and integrin-binding protein 1.30 With the abundance of poorly characterized CaM-binding proteins in plants, many of which have CaMBD''s with little sequence resemblance to the better characterized motifs in animals1 it seems likely that sequences will be identified which bind preferentially to the CaM N-lobe. Considering the incredible assortment of known CaM interaction modes and regulatory mechanisms, many of which have only been identified within the last decade, it is likely only a matter of time before CaM is proven to function as an adaptor protein in vivo.  相似文献   

6.
Cav1.4 channels are unique among the high voltage-activated Ca2+ channel family because they completely lack Ca2+-dependent inactivation and display very slow voltage-dependent inactivation. Both properties are of crucial importance in ribbon synapses of retinal photoreceptors and bipolar cells, where sustained Ca2+ influx through Cav1.4 channels is required to couple slow graded changes of the membrane potential with tonic glutamate release. Loss of Cav1.4 function causes severe impairment of retinal circuitry function and has been linked to night blindness in humans and mice. Recently, an inhibitory domain (ICDI: inhibitor of Ca2+-dependent inactivation) in the C-terminal tail of Cav1.4 has been discovered that eliminates Ca2+-dependent inactivation by binding to upstream regulatory motifs within the proximal C terminus. The mechanism underlying the action of ICDI is unclear. It was proposed that ICDI competitively displaces the Ca2+ sensor calmodulin. Alternatively, the ICDI domain and calmodulin may bind to different portions of the C terminus and act independently of each other. In the present study, we used fluorescence resonance energy transfer experiments with genetically engineered cyan fluorescent protein variants to address this issue. Our data indicate that calmodulin is preassociated with the C terminus of Cav1.4 but may be tethered in a different steric orientation as compared with other Ca2+ channels. We also find that calmodulin is important for Cav1.4 function because it increases current density and slows down voltage-dependent inactivation. Our data show that the ICDI domain selectively abolishes Ca2+-dependent inactivation, whereas it does not interfere with other calmodulin effects.Retinal photoreceptors and bipolar cells contain a highly specialized type of synapse designated ribbon synapses. Glutamate release in these synapses is controlled via graded and sustained changes in membrane potential that are maintained throughout the duration of a light stimulus (1, 2). In recent years, it became clear that Cav1.4 L-type Ca2+ channels are the main channel subtype converting these analog input signals into corresponding permanent glutamate release (1, 35). In support of this mechanism, mutations in the Cav1.4 gene have been identified in patients suffering from congenital stationary night blindness type 2 and X-linked cone rod dystrophy (68). Individuals displaying congenital stationary night blindness type 2 as well as mice deficient in Cav1.4 typically have abnormal electroretinograms that indicate a loss of neurotransmission from the rods to second order bipolar cells, which is attributable to a loss of Cav1.4 (3).Retinal Cav1.4 channels are set apart from other high voltage-activated (HVA)3 Ca2+ channels by their total lack of Ca2+-dependent inactivation (CDI) and their very slow voltage-dependent inactivation (VDI). Recently, we and others discovered an inhibitory domain (ICDI: inhibitor of CDI) in the C-terminal tail of the Cav1.4 channel that eliminates Ca2+-dependent inactivation in this channel by binding to upstream regulatory motifs (9, 10). Importantly, introducing the ICDI into the backbone of Cav1.2 or Cav1.3 almost completely abolishes the CDI of these channels. Contrasting with the clear cut function, the underlying mechanism by which ICDI abolishes CDI remains controversial. It was suggested that ICDI displaces the Ca2+ sensor calmodulin (CaM) from binding to the proximal C terminus (10), suggesting that the binding sites of CaM and ICDI are largely overlapping or allosterically coupled to each other. Alternatively, our own data rather suggested that CaM and the ICDI domain bind to different portions of the proximal C terminus (9). We proposed that the interaction between the ICDI domain and the EF-hand, a motif with a central role for transducing CDI (1116), switches off CDI without impairing binding of CaM to the channel. In this study, we designed experiments to differentiate between these two models. Here, using FRET in HEK293 cells, we provide evidence that in living cells, CaM is bound to the full-length C terminus of Cav1.4 (i.e. in the presence of ICDI). Furthermore, our data suggest that the steric orientation of the CaM/Cav channel complex differs between Cav1.2 and Cav1.4 channels. We show that CaM preassociation with Cav1.4 controls current density and also affects VDI. Thus, although CaM does not trigger CDI in Cav1.4 as it does in other HVA Ca2+ channels, it is still an important regulator of this channel.  相似文献   

7.
8.
9.
Potassium channels allow the selective flux of K+ excluding the smaller, and more abundant in the extracellular solution, Na+ ions. Here we show that Shab is a typical K+ channel that excludes Na+ under bi-ionic, Nao/Ki or Nao/Rbi, conditions. However, when internal K+ is replaced by Cs+ (Nao/Csi), stable inward Na+ and outward Cs+ currents are observed. These currents show that Shab selectivity is not accounted for by protein structural elements alone, as implicit in the snug-fit model of selectivity. Additionally, here we report the block of Shab channels by external Ca2+ ions, and compare the effect that internal K+ replacement exerts on both Ca2+ and TEA block. Our observations indicate that Ca2+ blocks the channels at a site located near the external TEA binding site, and that this pore region changes conformation under conditions that allow Na+ permeation. In contrast, the latter ion conditions do not significantly affect the binding of quinidine to the pore central cavity. Based on our observations and the structural information derived from the NaK bacterial channel, we hypothesize that Ca2+ is probably coordinated by main chain carbonyls of the pore´s first K+-binding site.  相似文献   

10.
Calmodulin binds to IQ motifs in the α1 subunit of CaV1.1 and CaV1.2, but the affinities of calmodulin for the motif and for Ca2+ are higher when bound to CaV1.2 IQ. The CaV1.1 IQ and CaV1.2 IQ sequences differ by four amino acids. We determined the structure of calmodulin bound to CaV1.1 IQ and compared it with that of calmodulin bound to CaV1.2 IQ. Four methionines in Ca2+-calmodulin form a hydrophobic binding pocket for the peptide, but only one of the four nonconserved amino acids (His-1532 of CaV1.1 and Tyr-1675 of CaV1.2) contacts this calmodulin pocket. However, Tyr-1675 in CaV1.2 contributes only modestly to the higher affinity of this peptide for calmodulin; the other three amino acids in CaV1.2 contribute significantly to the difference in the Ca2+ affinity of the bound calmodulin despite having no direct contact with calmodulin. Those residues appear to allow an interaction with calmodulin with one lobe Ca2+-bound and one lobe Ca2+-free. Our data also provide evidence for lobe-lobe interactions in calmodulin bound to CaV1.2.The complexity of eukaryotic Ca2+ signaling arises from the ability of cells to respond differently to Ca2+ signals that vary in amplitude, duration, and location. A variety of mechanisms decode these signals to drive the appropriate physiological responses. The Ca2+ sensor for many of these physiological responses is the Ca2+-binding protein calmodulin (CaM).2 The primary sequence of CaM is tightly conserved in all eukaryotes, yet it binds and regulates a broad set of target proteins in response to Ca2+ binding. CaM has two domains that bind Ca2+ as follows: an amino-terminal domain (N-lobe) and a carboxyl-terminal domain (C-lobe) joined via a flexible α-helix. Each lobe of CaM binds two Ca2+ ions, and binding within each lobe is highly cooperative. The two lobes of CaM, however, have distinct Ca2+ binding properties; the C-lobe has higher Ca2+ affinity because of a slower rate of dissociation, whereas the N-lobe has weaker Ca2+ affinity and faster kinetics (1). CaM can also bind to some target proteins in both the presence and absence of Ca2+, and the preassociation of CaM in low Ca2+ modulates the apparent Ca2+ affinity of both the amino-terminal and carboxyl-terminal lobes. Differences in the Ca2+ binding properties of the lobes and in the interaction sites of the amino- and carboxyl-terminal lobes enable CaM to decode local versus global Ca2+ signals (2).Even though CaM is highly conserved, CaM target (or recognition) sites are quite heterogeneous. The ability of CaM to bind to very different targets is at least partially due to its flexibility, which allows it to assume different conformations when bound to different targets. CaM also binds to various targets in distinct Ca2+ saturation states as follows: Ca2+-free (3), Ca2+ bound to only one of the two lobes, or fully Ca2+-bound (47). In addition, CaM may bind with both lobes bound to a target (5, 6) or with only a single lobe engaged (8). If a target site can bind multiple conformers of CaM, CaM may undergo several transitions that depend on Ca2+ concentration, thereby tuning the functional response. Identification of stable intermediate states of CaM bound to individual targets will help to elucidate the steps involved in this fine-tuned control.Both CaV1.1 and CaV1.2 belong to the L-type family of voltage-dependent Ca2+ channels, which bind apoCaM and Ca2+-CaM at carboxyl-terminal recognition sites in their α1 subunits (914). Ca2+ binding to CaM, bound to CaV1.2 produces Ca2+-dependent facilitation (CDF) (14). Whether CaV1.1 undergoes CDF is not known. However, both CaV1.2 and CaV1.1 undergo Ca2+- and CaM-dependent inactivation (CDI) (14, 15). CaV1.1 CDI is slower and more sensitive to buffering by 1,2-bis(o-aminophenoxy)ethane-N,N,N′,N′-tetraacetic acid than CaV1.2 CDI (15). Ca2+ buffers are thought to influence CDI and/or CDF in voltage-dependent Ca2+ channels by competing with CaM for Ca2+ (16).The conformation of the carboxyl terminus of the α1 subunit is critical for channel function and has been proposed to regulate the gating machinery of the channel (17, 18). Several interactions of this region include intramolecular contacts with the pore inactivation machinery and intermolecular contacts with CaM kinase II and ryanodine receptors (17, 1922). Ca2+ regulation of CaV1.2 may involve several motifs within this highly conserved region, including an EF hand motif and three contiguous CaM-binding sequences (10, 12). ApoCaM and Ca2+-CaM-binding sites appear to overlap at the site designated as the “IQ motif” (9, 12, 13), which are critical for channel function at the molecular and cellular level (14, 23).Differences in the rate at which 1,2-bis(o-aminophenoxy)ethane-N,N,N′,N′-tetraacetic acid affects CDI of CaV1.1 and CaV1.2 could reflect differences in their interactions with CaM. In this study we describe the differences in CaM interactions with the IQ motifs of the CaV1.1 and the CaV1.2 channels in terms of crystal structure, CaM affinity, and Ca2+ binding to CaM. We find the structures of Ca2+-CaM-IQ complexes are similar except for a single amino acid change in the peptide that contributes to its affinity for CaM. We also find that the other three amino acids that differ in CaV1.2 and CaV1.1 contribute to the ability of CaV1.2 to bind a partially Ca2+-saturated form of CaM.  相似文献   

11.
Cytoplasmic drops were prepared from internodal cells of thebrackish Characeae Lamprothamnium succinctum. Applying the patch-clamptechnique to single drops covered with tonoplast, we demonstratedthe presence of Ca2+-regulated K+ channels in the tonoplast.In a cell-attached mode, the selectivity of such channels forK+ was about 50 times that for Na+. This channel showed a tendencyto rectify in an outward direction. In the negative region ofthe pipette voltage, the conductance of this channel was 50pS, while it was 100 pS in the positive voltage region. Whenthe pipette voltage was increased above 50 mV, two conductancelevels were found in the cell-attached mode as well as in theexcised patch (cytoplasmic-side-out patch), which was obtainedby pulling the patch pipette from the cytoplasmic drop underconditions of low levels of Ca2+. Using the excised patch, wecontrolled the level of Ca2+ on the cytoplasmic side of thechannels. At a low level of Ca2+ (pCa=8) on the cytoplasmicside, the open frequency was very low and the opening time wasshort. An increase in Ca2+ on the cytoplasmic side (pCa = 5)increased both the frequency and the duration of opening. However,the conductance of the channels did not change. This regulationby Ca2+ of the K+ channels was reversible, that is, additionof EGTA on the cytoplasmic side inactivated the channels. Thepresent study demonstrates a direct action of Ca2+ on the K+channels. The physiological role of the K+ channel in the regulationof turgor in Lamprothamnium is discussed. (Received January 9, 1989; Accepted March 8, 1989)  相似文献   

12.
Tonoplast K+ channels of Chara corallina are well characterized but only a few reports mention anion channels, which are likely to play an important role in the tonoplast action potential and osmoregulation of this plant. For experiments internodal cells were isolated. Cytoplasmic droplets were formed in an iso-osmotic bath solution according to a modified procedure. Ion channels with conductances of 48 pS and 170 pS were detected by the patch-clamp technique. In the absence of K+ in the bath solution the 170 pS channel was not observed at negative pipette potential values. When Cl on either the vacuolar side or the cytoplasmic side was partly replaced with F, the reversal potential of the 48 pS channel shifted conform to the Cl equilibrium potential with similar behavior in droplet-attached and excised patch mode. These results showed that the 48 pS channel was a Cl channel. In droplet-attached mode the channel rectified outward current flow, and the slope conductance was smaller. When Chara droplets were formed in a bath solution containing low (10−8 m) Ca2+, then no Cl channels could be detected either in droplet-attached or in inside-out patch mode. Channel activity was restored if Ca2+ was applied to the cytoplasmic side of inside-out patches. Rectification properties in the inside-out patch configuration could be controlled by the holding pipette potential. Holding potential values negative or positive to the calculated reversal potential for Cl ions induced opposite rectification properties. Our results show Ca2+-activated Cl channels in the tonoplast of Chara with holding potential dependent rectification. Received: 30 March 1999/Revised: 10 August 1999  相似文献   

13.
14.
Cytoplasmic drops covered with the tonoplast were prepared frominternodal cells of Nitellopsis grown in fresh water. Applyingthe patch-clamp technique and the microinjection technique tosuch drops, we characterized the ion channels in the tonoplast.Both in cell-free patches and in the cytoplasmic-drop-attachedpatches, the tonoplast K+ channel was identified. The permeabilityratio between Na+ and K+ was calculated to be 0.2. This channelwould provide a molecular basis for the Na+/K+ exchange at thetonoplast. In cell-free patches, the K+channel was not activatedby Ca2+. However, in the case of attached patches, microinjectionof Ca2+ into a drop activated the K+ channel with a lag of afew seconds, suggesting that some cytoplasmic factor(s) maymediate the activation of the K+ channel by Ca2+. The conductanceof this channel was not changed by cytoplasmic Ca2+, but theprobability of opening increased markedly. In addition to theK+ channel, a second type of channel was also identified incell-free patches. This channel may be the Cl channel. 3 Present address: Department of Insect Physiology and Behavior,National Institute of Sericultural and Entomological Science,Tsukuba, Ibaraki, 305 Japan (Received August 6, 1990; Accepted December 6, 1990)  相似文献   

15.
Control of Passive Permeability in the Chara Plasmalemma   总被引:2,自引:0,他引:2  
Conductance to K+ alters as a function of membrane potential(m). Conductance to H+ (or OH) changes with externalpH (pHo) This conductance change can be modulated by alteringcytoplasmic pH or external K+ concentration, both of which alsoalter m. We suggest a role for H+ conductance in regulatingcytoplasmic pH above pHo 7.0.  相似文献   

16.
Using a patch-clamp technique in the whole-cell configuration, we studied the effect of a nitric oxide (NO) donor, nitroglycerin (NG), on outward transmembrane ion current in isolated smooth muscle cells (SMC) of the main pulmonary artery of the rabbit. We also studied the characteristics of unitary high-conductance Ca2+-dependent K+ channels (KCa channels) in the SMC membrane in the cell-attached and outside-out configurations. Nitroglycerin in a 10 M concentration increased the amplitude and intensified oscillations of outward transmembrane current induced by step depolarization. In this case, the threshold of activation of the current (–40 mV) did not change. If the potential was +70 mV, the transmembrane current in the presence of NG increased, as compared with the control, by 32.6 ± 19.4% (n = 6), on average. Simultaneous addition of 10 M NG and 1 mM tetraethylammonium chloride (TEA), a blocker of KCa channels, to the external solution at the potential of +70 mV decreased the amplitude of outward transmembrane current with respect to the control by 25.2 ± 11% (n = 6) and suppressed oscillations of this current. In the series of experiments carried out in the outside-out configuration (concentration of K+ ions in the external solution was 5.9 mM), we calculated the conductance of a single KCa channel, which was approximately 150 pS. In the case where the potential was equal to +40 mV, 1 mM TEA suppressed completely the current through unitary KCa channels. In the series of experiments performed in the cell-attached configuration, 100 M NG to a considerable extent intensified the activity of unitary high-conductance KCa channels by increasing the probability of the channel open state (P 0), on average, by 80 ± 1%, as compared with the control. In this case, NG did not influence the conductance of single KCa channels. We concluded that the NO donor NG increases the amplitude of outward transmembrane current in SMC of the rabbit main pulmonary artery by stimulation of the activity of TEA-sensitive high-conductance KCa channels. Our experiments carried out on single KCa channels demonstrated that the activating effect of NG on KCa channels is realized at the expense of an increase in the P 0 of these channels, but not of a change in the conductance of single channels.  相似文献   

17.
An internode of Chara was permeabilized as described by Shimmenand Tazawa [(1983) Protoplasma 117:93]. The Cl effluxof the permeabilized cell increased when the extracellular Ca2+concentration was increased, and the degree of the increasewas dependent on the Ca2+ concentration. This suggests thatthe Cl channel in the tonoplast was activated by Ca2+. (Received May 22, 1987; Accepted October 21, 1987)  相似文献   

18.
Large conductance Ca2+-activated potassium channels (BK) are targets for research that explores therapeutic means to various diseases, owing to the roles of the channels in mediating multiple physiological processes in various cells and tissues. We investigated the pharmacological effects of curcumin, a compound isolated from the herb Curcuma longa, on BK channels. As recorded by whole-cell patch-clamp, curcumin increased BK (α) and BK (α+β1) currents in transfected HEK293 cells as well as the current density of BK in A7r5 smooth muscle cells in a dose-dependent manner. By incubating with curcumin for 24 hours, the current density of exogenous BK (α) in HEK293 cells and the endogenous BK in A7r5 cells were both enhanced notably, though the steady-state activation of the channels did not shift significantly, except for BK (α+β1). Curcumin up-regulated the BK protein expression without changing its mRNA level in A7r5 cells. The surface expression and the half-life of BK channels were also increased by curcumin in HEK293 cells. These effects of curcumin were abolished by MG-132, a proteasome inhibitor. Curcumin also increased ERK 1/2 phosphorylation, while inhibiting ERK by U0126 attenuated the curcumin-induced up-regulation of BK protein expression. We also observed that the curcumin-induced relaxation in the isolated rat aortic rings was significantly attenuated by paxilline, a BK channel specific blocker. These results show that curcumin enhances the activity of the BK channels by interacting with BK directly as well as enhancing BK protein expression through inhibiting proteasomal degradation and activating ERK signaling pathway. The findings suggest that curcumin is a potential BK channel activator and provide novel insight into its complicated pharmacological effects and the underlying mechanisms.  相似文献   

19.
20.
The electrical conductance of the plasmalemma of cells of Charainflata, due to the diffusion of ions, consists predominantlyof K+, Cl and leak components. When the membrane electricalpotential difference is stepped in a negative direction witha voltage-clamp, the resulting inward current has componentsIK, ICl and IL (leak). During such voltage-clamp steps IK isinactivated, and Ic activated with voltage-dependent half-times.Increases in the external NaCl concentration reduce the magnitudeof IK and increase the magnitude of Ic, but reduce the half-timeof inactivation or activation. The NaCl-induced changes in Ikand ICl and their kinetics were more pronounced at pH0 =6.5than at pH0 =9.5. When the concentration of external CaCl2 wasincreased, Ik, ICl and the half-time of inactivation, (T1/2),of Ik were all reduced. The half-time of activation of ICl wasincreased. The NaCI-induced changes could result from increases in bothexternal ion concentration and osmotic pressure. Previous experimentshave shown that an increase in external osmotic pressure alonealters the properties of the conductances. In this paper weattempt to separate the purely ionic effects from the osmoticones. Key words: Chara inflata, ionic effects, K+ and Cl currents  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