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 500 毫秒
1.
Allosteric antagonism of insect odorant receptor ion channels   总被引:1,自引:0,他引:1  

Background

At a molecular level, insects utilize members of several highly divergent and unrelated families of cell-surface chemosensory receptors for detection of volatile odorants. Most odors are detected via a family of odorant receptors (ORs), which form heteromeric complexes consisting of a well-conserved OR co-receptor (Orco) ion channel and a non-conserved tuning OR that provides coding specificity to each complex. Orco functions as a non-selective cation channel and is expressed in the majority of olfactory receptor neurons (ORNs). As the destructive behaviors of many insects are principally driven by olfaction, Orco represents a novel target for behavior-based control strategies. While many natural and synthetic odorants have been shown to agonize Orco/Or complexes, only a single direct Orco modulator, VUAA1, has been described. In an effort to identify additional Orco modulators, we have investigated the structure/activity relationships around VUAA1.

Results

A search of our compound library identified several VUAA1 analogs that were selected for evaluation against HEK cells expressing Orco from the malaria vector Anopheles gambiae (AgOrco). While the majority of compounds displayed no activity, many of these analogs possess no intrinsic efficacy, but instead, act as competitive VUAA1 antagonists. Using calcium mobilization assays, patch clamp electrophysiology, and single sensillum in vivo recording, we demonstrate that one such candidate, VU0183254, is a specific allosteric modulator of OR signaling, capable of broadly inhibiting odor-mediated OR complex activation.

Conclusions

We have described and characterized the first Orco antagonist, that is capable of non-competitively inhibiting odorant-evoked activation of OR complexes, thereby providing additional insight into the structure/function of this unique family of ligand-gated ion channels. While Orco antagonists are likely to have limited utility in insect control programs, they represent important pharmacological tools that will facilitate the investigation of the molecular mechanisms underlying insect olfactory signal transduction.  相似文献   

2.
Regulation of neuronal ion channels via P2Y receptors   总被引:1,自引:0,他引:1  
Within the last 15 years, at least 8 different G protein-coupled P2Y receptors have been characterized. These mediate slow metabotropic effects of nucleotides in neurons as well as non-neural cells, as opposed to the fast ionotropic effects which are mediated by P2X receptors. One class of effector systems regulated by various G protein-coupled receptors are voltage-gated and ligand-gated ion channels. This review summarizes the current knowledge about the modulation of such neuronal ion channels via P2Y receptors. The regulated proteins include voltage-gated Ca2+ and K+ channels, as well as N-methyl-d-aspartate, vanilloid, and P2X receptors, and the regulating entities include most of the known P2Y receptor subtypes. The functional consequences of the modulation of ion channels by nucleotides acting at pre- or postsynaptic P2Y receptors are changes in the strength of synaptic transmission. Accordingly, ATP and related nucleotides may act not only as fast transmitters (via P2X receptors) in the nervous system, but also as neuromodulators (via P2Y receptors). Hence, nucleotides are as universal transmitters as, for instance, acetylcholine, glutamate, or -aminobutyric acid.  相似文献   

3.
As a liverwort Conocephalum conicum belongs to the oldest terrestrial plants1 and is phylogenetically located between green algae and higher plants. Recent patch-clamp recordings on Conocephalum vacuoles2,3 demonstrate ion channels very similar to higher plants and clearly different from vacuolar ion channels described in green algae. Here we summarize the features of a vacuolar cation channel and a vacuolar anion channel that both are common in terrestrial plants but are not detected in green algae, and we speculate about the molecular identity of these channels in the liverwort Conocephalum.Key words: vacuole, SV channel, anion channel, Conocephalum conicum, Embryophyta  相似文献   

4.
IRBIT (also called AHCYL1) was originally identified as a binding protein of the intracellular Ca2 + channel inositol 1,4,5-trisphosphate (IP3) receptor and functions as an inhibitory regulator of this receptor. Unexpectedly, many functions have subsequently been identified for IRBIT including the activation of multiple ion channels and ion transporters, such as the Na+/HCO3 co-transporter NBCe1-B, the Na+/H+ exchanger NHE3, the Cl channel cystic fibrosis transmembrane conductance regulator (CFTR), and the Cl/HCO3 exchanger Slc26a6. The characteristic serine-rich region in IRBIT plays a critical role in the functions of this protein. In this review, we describe the evolution, domain structure, expression pattern, and physiological roles of IRBIT and discuss the potential molecular mechanisms underlying the coordinated regulation of these diverse ion channels/transporters through IRBIT. This article is part of a Special Issue entitled: Calcium signaling in health and disease. Guest Editors: Geert Bultynck, Jacques Haiech, Claus W. Heizmann, Joachim Krebs, and Marc Moreau.  相似文献   

5.
6.

Background

Uridine 5''-triphosphate (UTP) and uridine 5''-diphosphate (UDP) act via P2Y receptors to evoke contraction of rat pulmonary arteries, whilst adenosine 5''-triphosphate (ATP) acts via P2X and P2Y receptors. Pharmacological characterisation of these receptors in intact arteries is complicated by release and extracellular metabolism of nucleotides, so the aim of this study was to characterise the P2Y receptors under conditions that minimise these problems.

Methods

The perforated-patch clamp technique was used to record the Ca2+-dependent, Cl- current (ICl,Ca) activated by P2Y receptor agonists in acutely dissociated smooth muscle cells of rat small (SPA) and large (LPA) intrapulmonary arteries, held at -50 mV. Contractions to ATP were measured in isolated muscle rings. Data were compared by Student''s t test or one way ANOVA.

Results

ATP, UTP and UDP (10-4M) evoked oscillating, inward currents (peak = 13–727 pA) in 71–93% of cells. The first current was usually the largest and in the SPA the response to ATP was significantly greater than those to UTP or UDP (P < 0.05). Subsequent currents tended to decrease in amplitude, with a variable time-course, to a level that was significantly smaller for ATP (P < 0.05), UTP (P < 0.001) and UDP (P < 0.05) in the SPA. The frequency of oscillations was similar for each agonist (mean≈6–11.min-1) and changed little during agonist application. The non-selective P2 receptor antagonist suramin (10-4M) abolished currents evoked by ATP in SPA (n = 4) and LPA (n = 4), but pyridoxalphosphate-6-azophenyl-2'',4''-disulphonic acid (PPADS) (10-4M), also a non-selective P2 antagonist, had no effect (n = 4, 5 respectively). Currents elicited by UTP (n = 37) or UDP (n = 14) were unaffected by either antagonist. Contractions of SPA evoked by ATP were partially inhibited by PPADS (n = 4) and abolished by suramin (n = 5). Both antagonists abolished the contractions in LPA.

Conclusion

At least two P2Y subtypes couple to ICl,Ca in smooth muscle cells of rat SPA and LPA, with no apparent regional variation in their distribution. The suramin-sensitive, PPADS-resistant site activated by ATP most resembles the P2Y11 receptor. However, the suramin- and PPADS-insensitive receptor activated by UTP and UDP does not correspond to any of the known P2Y subtypes. These receptors likely play a significant role in nucleotide-induced vasoconstriction.  相似文献   

7.
Large conductance calcium- and voltage-dependent BK potassium channels (aka BKCa, MaxiK, Slo1, KCa1.1, and KCNMA1) are expressed in a wide variety of tissues throughout the body and are activated by both intracellular Ca2+ and membrane depolarization. Owing to these properties, BK channels participate in diverse physiological processes from electrical excitability in neurons and secretory cells, and regulation of smooth muscle tone to tuning of auditory hair cells (Vergara et al., 1998; Ghatta et al., 2006). The response to voltage and Ca2+ allows BK channels to integrate electrical and calcium signaling, which is central to their physiological role. Understanding how BK and other multimodal channels are regulated by and integrate diverse stimuli is not only physiologically important but also relevant to the topic of conformational coupling. As a voltage- and ligand-dependent channel, BK channels contain both voltage-sensor and ligand-binding domains as well as a gate to regulate the flow of K+ through the pore. Coupling of conformational changes in one domain to another provides the basis for transducing voltage and ligand binding into channel opening and, therefore, defines, together with the functional properties of the gate and sensors, the signal transduction properties of the channel. The goal of this perspective is to provide an overview on the role and molecular basis of conformational coupling between functional domains in BK channels and outline some of the questions that remain to be answered.

BK channel structure

The BK channel is a member of the superfamily of voltage-gated channels that assembles as a homotetramer of pore-forming Slo1 (α) subunits. Each subunit contains seven transmembrane segments (S0–S6), including an S5–S6 pore gate domain (PGD) and voltage-sensor domain (VSD) that is likely to include S0–S4 segments (Liu et al., 2008) with charged voltage-sensing residues in S2, S3, and S4 (Fig. 1 A; Ma et al., 2006). In addition, a large C-terminal cytoplasmic domain (CTD) consisting of two homologous regulator of K+ conductance domains (RCK1 and RCK2) contains binding sites for Ca2+ and other ligands (Hou et al., 2009). The CTDs form a tetrameric gating-ring structure whose conformation changes upon Ca2+ binding. Crystal structures of the isolated BK channel gating ring and related prokaryotic Ca2+-activated K+ channel MthK have been solved in the presence and absence of Ca2+ (Jiang et al., 2002; Ye et al., 2006; Wu et al., 2010; Yuan et al., 2010, 2012). The atomic structure of the transmembrane domain has yet to be determined but is assumed to be homologous with that of voltage-gated potassium (KV) channels, consistent with a low resolution cryo-EM structure of the entire channel(Wang and Sigworth, 2009). Although BK channels formed from Slo1 alone are fully functional, each channel may also coassemble with up to four regulatory subunits of which several subtypes exist (β1–4 and γ1–4; Brenner et al., 2000; Yan and Aldrich, 2012). Regulatory subunits tune BK channel function in different tissues, contain one or two transmembrane segments, occupy sites adjacent to the VSD (Liu et al., 2010), and act in part to regulate voltage-dependent gating (Bao and Cox, 2005; Yan and Aldrich, 2010).Open in a separate windowFigure 1.BK channel structure. (A) Topology of the pore-forming Slo1 subunit, including VSD, PGD, and CTD. Charged residues in the VSD that are important for voltage sensing are highlighted in yellow. The CTD contains binding sites for Ca2+, Mg2+, and heme. (B) Schematic organization of functional domains in the tetrameric channel. S6 segments in the PDG are connected to the CTD gating ring through S6-RCK1 linkers.

BK channel function

The response of BK channels to voltage and Ca2+ is illustrated in Fig. 2 by plotting steady-state open probability (PO; Fig. 2 A) and Log(PO) (Fig. 2 B) versus voltage at different [Ca2+]i (0–100 µM) for heterologously expressed Slo1 channels (Horrigan and Aldrich, 2002). The Po-V relation in 0 Ca2+ (∼0.5 nM) shows channels can be fully activated in the absence of Ca2+ binding, but only at voltages approaching +300 mV. Calcium, to a first approximation, shifts Po-V to more negative voltages (Fig. 2 A), allowing the channel to activate in a physiological voltage range. However, plotting the data on a log scale reveals that Ca2+ does not simply shift the curve but rather increases Log(PO) in a nearly voltage-independent manner until PO saturates (Fig. 2 B). This response indicates that BK channels are activated independently by voltage- and Ca2+-sensors. Furthermore, Log(PO) becomes almost voltage independent at extreme negative voltages indicating that channels can open in the absence of voltage-sensor activation, a conclusion supported by gating current measurements (Horrigan and Aldrich, 1999).Open in a separate windowFigure 2.Voltage- and Ca2+-gating of BK channels. (A) PO-V relations for mSlo1 estimated as GK/GKmax from macroscopic tail currents after 30 ms voltage pulses in 0–100 µM Ca2+. (B) Log(Po)-V relations extend PO to <10−2 using steady-state unitary current recordings from macropatches. A and B represent mean ± SEM and are fit (solid curves) by the HA model (Horrigan and Aldrich, 2002). The increase in Log(Po) from 0 Ca2+ to saturating 100 µM Ca2+ at −120 mV (where voltage-sensors are in the resting state) reflects Ca2+-sensor/gate coupling energy (ΔΔGCOCa). The increase in Log(Po) in 0 Ca2+ from −120 mV to ∼+300 mV (where voltage sensors are fully activated) reflects voltage-sensor/gate coupling (ΔΔGCOV). The values of ΔΔGCOCa and ΔΔGCOV are determined from the change in the free energy of the gate (e.g., ΔΔGCOCa=ΔGCO[100Ca]ΔGCO[0Ca], where ΔGCOkTln[PO/(1 ? PO)]), and, in the case of ΔΔGCOV, the measurement at −120 mV must be extrapolated to positive voltages (dashed line) to correct for the weak voltage dependence of the C-O transition (zL = 0.3 e). (C) Allosteric model indicates the possible conformations of the gate (C and O), voltage sensors (R and A), and Ca2+ sensors (X and X-Ca2+) in each of four subunits and allosteric factors (C, D, and E) that describe the energetic coupling between these three parts of the channel. J and L are voltage-dependent equilibrium constants with zero-voltage values J0 and L0 and partial charges (zJ, zL), K = [Ca2+]/KD, where KD is the elementary Ca2+ dissociation constant for the closed channel.

An allosteric mechanism of BK channel gating

That BK channels exhibit basal activity in the absence of voltage-sensor activation and Ca2+ binding implies that sensor/gate coupling is not an obligatory process. That is, sensor activation promotes but is not required for channel opening. The ability of voltage or Ca2+ sensors in different subunits to influence a concerted conformational change (opening) in a nonobligatory fashion is well described in terms of allosteric mechanisms (Monod et al., 1965). A dual allosteric model (Fig. 2 C, HA model) was used to fit the steady-state data in Fig. 2 (A and B, curves), accounts for many other features of Slo1 gating (Horrigan and Aldrich, 1999, 2002; Horrigan et al., 1999), and provides a useful framework for analyzing BK channel gating in terms of domain/domain interactions. The HA (Horrigan-Aldrich) model asserts that the channel gate can undergo a closed to open (C-O) conformational change that is regulated by four independent and identical voltage- and Ca2+-sensors. Voltage sensors can be in a resting (R) or activated (A) conformation, whereas Ca2+ sensors can be Ca2+ free (X) or Ca2+ bound (X-Ca2+). The function of each domain is defined by equilibrium constants for gate opening (L), voltage-sensor activation (J), and Ca2+ binding (K). The coupling, or energy transfer, between domains is represented by allosteric factors (C, D, and E) which define the ability of a transition in one domain to affect the equilibrium constant in another.

The energetics of conformational coupling

The structures of BK and homologous channels provide important clues concerning the molecular basis of conformational coupling, as discussed below. However, many features of coupling can only be resolved through structure-function analysis using site-directed mutagenesis. A prerequisite to such analysis is to quantify coupling interactions represented by allosteric factors C, D, and E in the HA model. Mutations that alter channel activity may perturb sensors, the gate, or their coupling. Therefore, it is crucial to distinguish changes in coupling from changes in the function of sensor or gate. One way to do this is by fitting steady-state data over a wide range of voltage and [Ca2+] as in Fig. 2 (A and B) to determine all parameters in the HA model. However, a more direct and model-independent approach to measure coupling is to determine the energetic effect on one domain of forcing coupled domains into defined conformations (e.g., all activated or deactivated) under extreme stimulus conditions (see also Chowdhury and Chanda in this issue). For example, the total coupling between all Ca2+ sensors and the gate (ΔΔGCOCa=5.0kcal mol1) can be determined by comparing Log(Po) in 0 Ca2+ and saturating 100 µM Ca2+ at extreme negative voltages where voltage sensors are in the resting state, and voltage-sensor/gate coupling (ΔΔGCOV=7.6kcal mol1) can be determined by comparing Log(Po) at extreme negative and positive voltages in 0 Ca2+ (Fig. 2 B; Horrigan and Aldrich, 2002). In general, the HA model predicts that channels can occupy many different open and closed states, defined by the number of voltage- and Ca2+-sensors activated in each channel. But under extreme stimulus conditions, where all voltage- or Ca2+-sensors are either activated or deactivated, gating reduces to a single closed and open state. For a two-state process, the free energy difference between closed and open (ΔGCO) can be defined in terms of the equilibrium constant (Po/1-Po) for the C-O transition (i.e., ΔGCOkTln[PO/(1 ? PO)]; Chowdhury and Chanda, 2010, 2012). Thus, changes in Log(Po) in Fig. 2 B reflect the energetic effects of voltage- and Ca2+-sensor activation on the gate. Similarly, a weak coupling between voltage- and Ca2+-sensors (ΔΔGRACa=0.5kcal mol1) has been measured by comparing the effects of 0 Ca2+ and saturating Ca2+ on voltage-sensor activation measured with gating currents while the gate is closed (Horrigan and Aldrich, 2002).Although measurement of Ca2+-sensor/gate coupling as in Fig. 2 B is relatively straightforward, measurement of voltage-sensor/gate coupling is subject to several challenges (Fig. 3). The relationship between gating and voltage-sensor/gate coupling is illustrated in Fig. 3 A by expanding the HA model to show the four combinations of states that the gate (C and O) and voltage-sensor (R and A) can assume in a single subunit (RC, RO, AC, and AO) together with the equilibrium constants between them. The equilibrium constant for the C-O transition increases from L when the voltage sensor is in the resting state (Fig. 3 A, #1) to LD when a voltage sensor is activated (Fig. 3 A, #2). Thus the voltage-sensor/gate coupling factor D can be determined by comparing Po at extreme voltages (Fig. 3 B, #1 and 2), with the expectation that Po/(1-Po) will increase by a factor of D4 when all four voltage sensors are activated. One challenge is the large dynamic range of these measurements, which span a seven-order-of-magnitude change in Po and require both macroscopic and unitary current recordings in the same patch and a high level of expression that cannot always be achieved with mutant channels. A related challenge is that Po in 0 Ca2+ approaches saturation near unity at +300 mV as voltage sensors become fully activated (Fig. 2 A). As Po approaches unity, determination of the equilibrium constant Po/(1-Po) becomes problematic and coupling energy can therefore be underestimated, especially in the presence of Ca2+ or with mutants that are easier to open than the WT because Po may saturate before voltage sensors are fully activated. Finally, the C-O conformational change, as in many ion channels, has a weak intrinsic voltage dependence (L(V)) that must be taken into account when comparing Po at extreme voltages (Figs. 2 B and 3 B, dashed lines).Open in a separate windowFigure 3.The energetics of voltage sensor/gate coupling. (A) Voltage sensor/gate states in a single subunit. Numbers refer to data in B, C, and D used to determine the indicated equilibrium constants. (B) PO-V determines L (1) and LD4 (2). Dashed lines indicate the predicted voltage dependence of PO with all voltage sensors either in the resting or activated state and zL = 0.3 e. (C) Normalized QC-V for closed channels determined from ON gating currents (inset) is fit by a Boltzmann function with zJ = 0.58 e and defines J (3). (D) QO-V relation for open channels defines JD (4). QO is estimated by fitting the foot of the qa-V relation (qa = kTdln(PO)/dV) with a Boltzmann function. The shift between QO and QC relations defines the coupling factor D.A complementary approach to measure voltage-sensor/gate coupling is to determine the effect of channel opening on voltage-sensor activation. The equilibrium constant for the R-A transition increases from J when channels are closed to JD when channels are open (Fig. 3 A, #3 and 4). The equilibrium constant J is determined from the charge distribution for closed channels (QC-V; Fig. 3 C). Qc can be measured by integrating ON gating currents (Fig. 3 B, inset) because voltage-sensor activation in BK channels is rapid and occurs while most channels remain closed. The equilibrium constant JD is determined from the charge distribution for open channels (QO-V; Fig. 3 D). QO cannot be determined from gating currents because BK channels close rapidly. However, QO can be estimated in a model-independent fashion from the log-slope of the Po-V relation (qa = kTdln(PO)/dV), as shown in Fig. 3 D (Horrigan and Aldrich, 2002). The QO-V and QC-V relations have the same shape but are shifted relative to each other by a voltage directly proportional to coupling energy for a single voltage sensor (ΔV=kT ln(D)/zJ=ΔΔGCOV/4zJ; Ma et al., 2006). This method yields similar results as in Fig. 3 B and avoids challenges relating to the voltage dependence of L(V), or PO saturation. However, gating currents are difficult to measure in BK channels and require an even higher level of expression than in Fig. 3 B.

Molecular mechanisms of conformational coupling

Although the energetics and biophysical mechanisms of sensor/gate and sensor/sensor coupling in BK channels are well defined, many fundamental questions remain concerning the molecular basis of these interactions. Where in the channel do they occur? What are the identity and nature of amino acid interactions involved? When do the interactions occur during channel activation? These questions are broad because many potential sites of domain/domain interaction exist, the conformational changes that these domains and their interfaces undergo during gating are not all understood in detail, and insight provided by the structure of BK or homologous channels is in many cases limited. For example, although voltage-sensor/gate coupling is likely to be mediated in large part by interfaces between the VSD and PGD as in Kv channels, there is also a unique interface between the VSD and CTD in BK channels that may contribute to this process. Intracellular Mg2+ is a BK channel activator that is coordinated by residues in both VSD and CTD and interacts electrostatically with the voltage sensor, indicating (together with cross-linking experiments) that these domains come in contact (Yang et al., 2007, 2008). The primary functional effect of Mg2+ is to enhance voltage-sensor/gate coupling, suggesting that VSD/CTD interaction could provide a basis for this process (Horrigan and Ma, 2008). VSD/CTD interaction must also mediate coupling between voltage and Ca2+ sensors. However information about the VSD/CTD interface is incomplete. Structures of BK and MthK channel CTDs and the Kv channel VSD help identifying residues that may lie at this interface. But the complete BK channel structure including the VSD/CTD interface has yet to be determined, and neither MthK nor Kv channels contain a homologous interface.Because most domain/domain interactions involved in conformational coupling remain to be identified, it is useful to consider what properties such interactions must have. In general, coupling may occur through linkers that directly connect two domains or through noncovalent interactions at domain/domain interfaces. In BK channels, Ca2+-sensor/gate coupling is thought to be mediated by the S6-RCK1 linker connecting CTD to PGD (Jiang et al., 2002; Niu et al., 2004). However multiple domain/domain interfaces are also involved in coupling, as noted above. Structural information has helped to identify where these interfaces are and to locate potential interaction partners. But elucidating the basis of conformational coupling is more complex than simply identifying domain/domain interactions. Coupling is necessarily a state-dependent process that depends on the conformations of two coupled domains. To understand the possible interaction mechanisms, it is useful to consider the energetics of coupling in the context of a gating cycle such as that describing voltage-sensor/gate coupling in Fig. 3 A. The coupling energy between a single voltage sensor and gate (ΔΔGCORA=kTln(D)) can be expressed in terms of the free energies of the four states involved (AC, AO, RC, and RO):ΔΔGCORA=[ΔGCOAΔGCOR]=[(GCAGOA)(GCRGOR)]=GCAGOAGCR+GOR(1)Therefore, the change in coupling produced by changes in the free energy of each state is:ΔΔΔGCORA=[ΔGCAΔGOAΔGCR+ΔGOR](2)Eq. 2 indicates that coupling is enhanced (ΔΔΔGCORA>0) by decreasing GCR or GOA (stabilizing the RC or AO state) or by increasing GCA or GOR (destabilizing the AC or RO state). Coupling could also be enhanced by changing the free energy of multiple states, provided the net contribution (Eq. 2) is positive. However, interactions that are state independent or depend on the conformation of only one domain (e.g., ΔGCA=ΔGCR or ΔG0R=ΔGCR) will have no effect on coupling (i.e.,ΔΔGCORA=0).This analysis has several important implications for elucidating mechanisms of conformational coupling. First, to identify coupling interactions by structural methods requires that structures be solved in four conformations defined by two domains and have sufficient resolution to observe state-dependent changes in interaction. In practice, because changes in interaction energy may be subtle or difficult to estimate based on structure, coupling interactions would have to be verified and quantified by structure-function analysis. When only one or two structures of a channel are available, as is currently the case for BK and homologous channels, then structure-function analysis is required to determine if observed interactions participate in coupling, and residues that do not interact in an observed structure may still mediate coupling by interacting in other conformations. Second, structure-function analysis involving site-directed mutagenesis and measurements of coupling energy, as in Figs. 2 and and3,3, should be sufficient to identify residues involved in coupling interactions. However, such experiments must be interpreted cautiously. Allosteric proteins are often sensitive to modulation at domain interfaces by heterotropic ligands that bind and introduce interactions that do not normally exist, and mutations can have similar effects. For example, introduction of a positively charged residue near the BK channel Mg2+-binding site (Q397R) mimics the effect of Mg2+ by introducing an electrostatic interaction (Yang et al., 2007). Thus, the ability of Q397R to enhance coupling does not indicate that Q397 is normally involved in this process, just that it is located at a sensitive interface where conformational changes associated with voltage-sensor activation and channel opening both occur. Such false positives cannot be avoided simply by restricting analysis to mutations that reduce coupling because, based on Eq. 2, coupling can be reduced by introducing interactions that destabilize the RC or AO states or stabilize the AC or RO states. Intracellular heme, which inhibits voltage-sensor/gate coupling in BK channels, may act in this way (Horrigan et al., 2005). Consequently, whether a mutation increases or decreases coupling is less critical than the use of mutations such as Ala that are least likely to introduce interactions. Finally, to define the mechanism of voltage-sensor/gate coupling we must not only identify mutations that alter coupling but also characterize those changes in terms of the effect on individual states (i.e., RC, RO, AC, and AO) to determine when during activation the interactions occur. Although the change in free energy of any particular state cannot be determined directly, the equilibria between states can be evaluated as in Fig. 3 to define the change in free energy of individual states relative to each other. Such analysis has indicated that Mg2+ enhances voltage-sensor/gate coupling by preferentially stabilizing the AO state (Horrigan and Ma, 2008), whereas a stabilization of the RO state is consistent with the inhibitory effect of heme (Horrigan et al., 2005). It is worth noting that, depending on their state dependence, coupling interactions may influence open probability when sensors are in a resting conformation. Thus, the basal activity of BK channels in the absence of sensor activation defined by the equilibrium constant L in the HA model cannot be interpreted simply as the intrinsic stability of the gate because it also reflects any state-dependent interactions between the gate and resting sensors.

Mechanisms of voltage-sensor/gate coupling

Kv channel structures reveal two interfaces between VSD and PGD: intrasubunit contacts between the S4–S5 linker and S6 and intersubunit contacts between S4 and S5 (Long et al., 2005a, 2007). In BK channels, additional intersubunit interactions occur between the VSD and CTD (Yang et al., 2008). All of these interfaces potentially play a role in voltage-sensor/gate coupling in BK channels, but their relative contributions and mechanisms remain unknown. The S4/S5 interface is implicated because a mutation near the top of S4 (R210E) reduces coupling energy by half (Ma et al., 2006). However, we cannot rule out that this is a false positive produced by introducing rather than disrupting an interaction because other mutations at this site (R210C and R210N) appear to constitutively activate the channel preventing measurement of voltage-sensor/gate coupling (Ma et al., 2006). Similarly, the VSD/CTD interface is implicated indirectly based on the ability of Mg2+, Q397R, and heme to modulate coupling by introducing interactions at or near this interface.Interactions between the S4–S5 linker and S6 are widely considered to underlie voltage-sensor/gate coupling, also known as electromechanical coupling, in Kv channels (Lu et al., 2002; Tristani-Firouzi et al., 2002; Long et al., 2005b; Chowdhury and Chanda, 2012) and are also likely to be important in BK channels (Sun et al., 2012). However, even in Kv channels many of the questions raised above remain to be answered, such as what are the individual residues and nature of interactions that contribute to coupling and when do they occur during gating? The importance of these questions can be illustrated by comparing three hypothetical mechanisms of voltage-sensor/gate coupling that include S4–S5/S6 interactions and are consistent with the allosteric nature of voltage gating in Slo1 (Fig. 4). Fig. 4 (A–C) depicts the four combinations of states the gate and voltage-sensor can assume in a single subunit. The possibility that S4–S5/S6 forms a rigid connection that forces sensor and gate to move as a unit can be ruled out because voltage sensors can activate while channels remain closed. However, it is conceivable that the S4–S5 linker remains bound to S6 at all times, whereas flexibility in adjoining regions allows sensor and gate to move as if coupled by a spring (Fig. 4 A). Another possibility is the S4–S5 linker binds to the open gate only when the voltage sensor is activated, stabilizing the AO state (Fig. 4 B). Alternatively, the S4–S5 linker of the resting voltage sensor might clash with the open gate to destabilize the RO state (Fig. 4 C). These and other mechanisms could account for the ability of voltage-sensor activation to promote opening but make different predictions concerning the source of coupling and the role of S4–S5/S6 interaction. Fig. 4 A suggests that many parts of the S4–S5 linker could contribute to coupling by influencing the mechanical properties of the linkage. In this case, S4–S5/S6 interaction acts merely as a passive connection that affects coupling only to the extent that it is intact or disrupted. In contrast, Fig. 4 (B and C) predicts that the S4–S5/S6 interaction energy is the main determinant of coupling energy. However, the panels in Fig. 4 differ in the nature of the proposed interaction (Fig. 4 B, binding; Fig. 4 C, steric hindrance) and the state dependence of the interaction. The analysis of coupling energy and equilibria for such gating cycles, as outlined above, should allow such mechanisms and predictions to be distinguished in BK channels.Open in a separate windowFigure 4.Three models of voltage-sensor/gate coupling. Interaction of S4–S5 linker with the S6 gate in a single subunit. The voltage sensor can be in a resting (R) or activated (A) state, and the gate is closed (C) or open (O). (A) Flexible linkage with S4–S5/S6 interacting in all states. (B) S4–S5/S6 binding stabilizes AO state. (C) Steric hindrance destabilizes the RO state.Despite substantial homology between BK and KV channels, important differences also exist that could be relevant to voltage-sensor/gate coupling. First, BK channels are weakly voltage dependent compared with Kv channels and exhibit a different pattern and contribution of voltage-sensing residues in the VSD, suggesting that conformational changes associated with voltage-sensor activation may differ (Ma et al., 2006). Second, the ability of intracellular blockers and Cys-modifying reagents to access the inner pore of closed BK channels (Wilkens and Aldrich, 2006; Zhou et al., 2011) suggests that the BK channel gate may be formed by the selectivity filter (Cox and Hoshi, 2011; Thompson and Begenisich, 2012), as in cyclic-nucleotide–gated channels (Contreras et al., 2008), rather than by crossing of S6 segments at the inner mouth of the pore, as in Kv channels (del Camino and Yellen, 2001). These differences do not require that voltage-sensor/gate coupling occurs through fundamentally different mechanisms. For example, if the BK channel gate is in the selectivity filter it is still likely to be strongly coupled to S6 movement, as in CNG channels (Flynn and Zagotta, 2001), and therefore subject to control by S4–S5/S6 interaction. Indeed, there is considerable evidence for S6 movement associated with BK channel opening (Li and Aldrich, 2006; Wu et al., 2009; Chen and Aldrich, 2011). However, the coupling mechanisms for BK and Kv channels may differ in detail, and a selectivity gate in the BK channel could potentially support a larger role for S4/S5 interaction in voltage-sensor/gate coupling. It should also be noted that regulatory β and γ subunits have been reported to modulate voltage-sensor/gate coupling (Bao and Cox, 2005; Yan and Aldrich, 2010); thus, additional state-dependent interactions may form in BK channels between the VSD and regulatory subunits.

Mechanisms of coupling Ca2+ sensors to the gate and voltage sensors

Of the three coupling interactions in the HA model, Ca2+-sensor/gate coupling is easiest to measure and best understood at a molecular level. Nonetheless, many questions and controversies remain to be resolved concerning the conformational changes that occur upon Ca2+ binding and the mechanisms coupling these changes to the gate and voltage sensors.Ca2+-dependent activation is generally consistent with the idea, originally proposed for MthK, that Ca2+ binding causes a conformational change in the gating ring that opens the channel by pulling on the RCK1-S6 linker (Jiang et al., 2002). Crystal structures in the presence and absence of Ca2+ suggest that MthK and BK channel gating rings expand in diameter by 8 and 12 Å, respectively, upon Ca2+ binding (Ye et al., 2006; Wu et al., 2010; Yuan et al., 2012). Experimental evidence in BK channels shows a monotonic relationship between channel activation and linker length, suggesting that the linker is under constant tension in the presence or absence of Ca2+, like a spring (Niu et al., 2004). Likewise, effects on Ca2+-sensor/gate coupling and Ca2 sensitivity of mutations in the N-terminal AC region of the BK channel RCK1 domain suggest that the flexibility of this region is important for transmitting Ca2+-dependent conformational changes to the gate (Yang et al., 2010). Although the latter results are consistent with the linker hypothesis, they also have raised the possibility that conformational changes in the AC region could be coupled to the gate through direct contact with the PGD. Consistent with this possibility, the N-terminal half of RCK1 undergoes a substantial reorientation relative to the rest of the BK channel gating ring upon Ca2+ binding (Yuan et al., 2012). However, determining whether a CTD/PDG interface exists and contributes to coupling may have to wait until a complete BK channel structure is available.The BK channel CTD contains two high affinity Ca2+-binding sites (one in each RCK domain) that have been identified by site-directed mutagenesis (Bao et al., 2004; Zhang et al., 2010). By using mutations to eliminate each site individually and carefully measuring the effects of Ca2+ on Po at −80 mV (similar to Fig. 2 B), Sweet and Cox (2008) determined that the contribution to Ca2+-sensor/gate coupling of the RCK1 site (3.74 kCal mol−1) was slightly greater than that of the RCK2 Ca2+ bowl site (3.04 kCal mol−1), but the RCK2 site has a higher affinity for Ca2+. In addition, the summed contribution of the individual sites exceeded their combined effect in the WT channel (ΔΔGCOCa=5.0kcal mol1; Fig. 2 B), consistent with negative cooperativity between sites in the same subunit (Sweet and Cox, 2008). In contrast, another group performing similar experiments at +50 mV concluded that positive cooperativity exists between the two binding sites (Qian et al., 2006). In the latter case, positive cooperativity could potentially be accounted for if both Ca2+ sites were coupled to the voltage sensor because the voltage sensor is not held in a resting state at +50 mV. However Sweet and Cox (2008) concluded that the RCK1 site is solely responsible for Ca2+-sensor/voltage-sensor coupling. A recent study combining Ca2+ site mutations and voltage clamp fluorometry to monitor the effects of Ca2+ on steady-state voltage-sensor activation and Po reached conclusions similar to that of Sweet and Cox (2008) regarding intrasubunit cooperativity and Ca2+-sensor/voltage-sensor coupling (Savalli et al., 2012). However, this study also suggested that Ca2+ site mutations may have effects other than elimination of Ca2+ binding. This caveat, together with the fact that none of the studies mentioned measured Ca2+-sensor/voltage-sensor coupling directly, suggests that the extent to which the individual Ca2+-binding sites are coupled to the voltage sensor or to each other, as well as the molecular mechanisms mediating such interactions, is still open to question.Although the CTD contains two high affinity Ca2+-binding sites, only the Ca2+ bowl is occupied in the Ca2+-bound gating-ring structure (Yuan et al., 2012), raising questions of to what extent this structure resembles the Ca2+-saturated conformation in the intact channel. In principal, the two Ca2+ sites could have independent effects on activation by stabilizing a single Ca2+-bound open conformation. However, mutations in the AC region of RCK1 selectively alter the Ca2+ sensitivity of the RCK1 site (Yang et al., 2010), suggesting that some conformational changes may be coupled to RCK1 occupancy alone. This, together with the fact that the gating ring of the intact channel is expected to contact the VSD, suggests that significant differences in structure could exist between the crystal structure and intact Ca2+-saturated gating ring. A related question is to what extent gating ring expansion is determined by channel opening (i.e., RCK1-S6 linker tension) versus Ca2+ binding. Because gating ring expansion is observed upon Ca2+ binding in isolated gating rings, it seems reasonable to suppose that expansion and channel opening might not occur simultaneously. If the gating ring could expand into a high-Ca2+-affinity conformation while the channel is closed, it could have important implications for gating models and for the possible state dependence of CTD/VSD coupling. However, there is as yet no evidence supporting this possibility. Indeed Sweet and Cox (2008) concluded that their results were best fit by assuming gating-ring expansion is tightly coupled to channel opening, consistent with the HA model.A final question relates to conformational events that occur in Ca2+-binding sites during channel opening. As in any allosteric model of ligand-dependent gating, the HA model predicts that Ca2+-binding sites must have a higher affinity for Ca2+ in the open than the closed conformation. Therefore, understanding how Ca2+ coordination changes upon channel opening is fundamental to the mechanism of Ca2+-sensor/gate coupling. In principal, the contribution of individual Ca2+-coordinating residues to state-dependent binding can be evaluated by determining the effect on coupling energy of mutating such sites (Purohit et al., 2012). Previous studies have generally identified likely Ca2+-coordination sites based on a reduced sensitivity of mutated channels to Ca2+ in the physiological range (≤100 µM). However, experiments at higher [Ca2+] may be required to saturate mutated sites and determine if Ca2+-sensor/gate coupling is altered.

Summary

The BK channel is an important example of a voltage- and ligand-gated channel whose function depends on conformational coupling between multiple domains. Many questions remain about the molecular basis of domain–domain coupling, but this channel represents a favorable system for studying such processes owing to its unique functional properties and methods that have allowed the energetics of coupling to be studied in detail. Combining these methods with emerging structural information about domain/domain interfaces and conformational changes in the channel should provide further insight into coupling mechanisms that are important in BK channels and in other voltage-gated or ligand-gated channels. BK channels also constitute a powerful system for understanding the interplay between ligand- and voltage-dependent gating. Defining the interactions that mediate coupling between voltage and Ca2+ sensors in this channel should provide unique insight into processes that may be relevant to other multimodal channels such as HCN or TRP.This Perspectives series includes articles by Andersen, Colquhoun and Lape, and Chowdury and Chanda.  相似文献   

8.
Ion conductance and ion selectivity of potassium channels in snail neurones   总被引:13,自引:0,他引:13  
Summary Delayed potassium channels were studied in internally perfused neurone somata from land snails. Relaxation and fluctuation analysis of this class of ion channels revealed Hodgkin-Huxley type K channels with an average single channel conductance ( K) of 2.40±0.15 pS. The conductance of open channels is independent of voltage and virtually all K channels seem to be open at maximum K conductance (g K) of the membrane. Voltage dependent time constants of activation ofg K, calculated from K current relaxation and from cut-off frequencies of power spectra, are very similar indicating dominant first-order kinetics. Ion selectivity of K channels was studied by ion substitution in the external medium and exhibited the following sequence: T1+>K+>Rb+>Cs+>NH 4 + >Li+>Na+. The sequence of the alkali cations does not conform to any of the sequences predicted by Eisenman's theory. However, the data are well accommodated by a new theory assuming a single rate-limiting barrier that governs ion movement through the channel.This paper is dedicated to the memory of Walther Wilbrandt.  相似文献   

9.
10.

Background  

Taste receptor cells are responsible for transducing chemical stimuli from the environment and relaying information to the nervous system. Bitter, sweet and umami stimuli utilize G-protein coupled receptors which activate the phospholipase C (PLC) signaling pathway in Type II taste cells. However, it is not known how these cells communicate with the nervous system. Previous studies have shown that the subset of taste cells that expresses the T2R bitter receptors lack voltage-gated Ca2+ channels, which are normally required for synaptic transmission at conventional synapses. Here we use two lines of transgenic mice expressing green fluorescent protein (GFP) from two taste-specific promoters to examine Ca2+ signaling in subsets of Type II cells: T1R3-GFP mice were used to identify sweet- and umami-sensitive taste cells, while TRPM5-GFP mice were used to identify all cells that utilize the PLC signaling pathway for transduction. Voltage-gated Ca2+ currents were assessed with Ca2+ imaging and whole cell recording, while immunocytochemistry was used to detect expression of SNAP-25, a presynaptic SNARE protein that is associated with conventional synapses in taste cells.  相似文献   

11.
The ion channel of the nicotinic acetylcholine receptor (nAChR) is believed to be lined by transmembrane M2 helices. A 4-8-12 sequence motif, comprising serine (S) or threonine (T) residues at positions 4, 8 and 12 of M2, is conserved between different members, anion and cation selective, of the nAChR superfamily. Parallel bundles of 4-8-12 motif-containing helices are considered as simplified models of ion channels. The relationship between S and T sidechain conformations and channelion interactions is explored via evaluation of interaction energies of K+ and of Cl ions with channel models. Energy calculations are used to determine optimal 2 (C-C\-O-H) values in the presence of K+ or Cl ions. 4-8-12 motif-containing bundles may form favourable interactions with either cations or anions, dependent upon the 2 values adopted. Parallel-helix and tilted-helix bundles are considered, as are heteromeric models designed to mimic the Torpedo nAChR. The main conclusion of the study is that conformational flexibility at 2 enables both S and T residues to form favourable interactions with anions or cations. Consequently, there is apparently no difference between S and T residues in their interactions with permeant ions, which suggests that the presence of T vs. S residues within the 4-8-12 motif is not a major mechanism whereby anion/cation selectivity may be generated. The implications of these studies with respect to more elaborate models of nAChR and related receptors are considered.Abbreviations nAChR, GluR, NMDA-R, 5HT3-R, GABAAR, GlyR nicotinic acetylcholine, glutamate, NMDA, 5HT3, GABAA and glycine receptors, respectively - PhTx philanthotoxin - M2 second membrane-spanning helix of receptor-channel subunits  相似文献   

12.
Lau WK  Chow AW  Au SC  Ko WH 《PloS one》2011,6(7):e22363

Background

Cysteinyl leukotriene (CysLT) is one of the proinflammatory mediators released by the bronchi during inflammation. CysLTs exert their biological effects via specific G-protein-coupled receptors. CysLT1 receptor antagonists are available for clinical use for the treatment of asthma. Recently, crosstalk between CysLT1 and P2Y6 receptors has been delineated. P2Y receptors are expressed in apical and/or basolateral membranes of virtually all polarized epithelia to control the transport of fluid and electrolytes. Previous research suggests that CysLT1 receptor antagonists inhibit the effects of nucleotides acting at P2Y receptors. However, the detailed molecular mechanism underlying the inhibition remains unresolved.

Methodology/Principal Findings

In this study, western blot analysis confirmed that both CysLT1 and P2Y6 receptors were expressed in the human bronchial epithelial cell line 16HBE14o-. All three CysLT1 antagonists inhibited the uridine diphosphate (UDP)-evoked ISC, but only montelukast inhibited the UDP-evoked [Ca2+]i increase. In the presence of forskolin or 8-bromoadenosine 3′5′ cyclic monophosphate (8-Br-cAMP), the UDP-induced ISC was potentiated but was reduced by pranlukast and zafirlukast but not montelukast. Pranlukast inhibited the UDP-evoked ISC potentiated by an Epac activator, 8-(4-Chlorophenylthio)-2′-O-methyladenosine-3′,5′-cyclic monophosphate (8-CPT-2′-O-Me-cAMP), while montelukast and zafirlukast had no such effect. Pranlukast inhibited the real-time increase in cAMP changes activated by 8-CPT-2′-O-Me-cAMP as monitored by fluorescence resonance energy transfer imaging. Zafirlukast inhibited the UDP-induced ISC potentiated by N6- Phenyladenosine- 3′, 5′- cyclic monophosphorothioate, Sp- isomer (Sp-6-Phe-cAMP; a PKA activator) and UDP-activated PKA activity.

Conclusions/Significance

In summary, our data strongly suggest for the first time that in human airway epithelia, the three specific CysLT1 receptor antagonists exert differential inhibitory effects on P2Y6 receptor-coupled Ca2+ signaling pathways and the potentiating effect on ISC mediated by cAMP and Epac, leading to the modulation of ion transport activities across the epithelia.  相似文献   

13.

Background

Cystic fibrosis (CF) respiratory epithelia are characterized by a defect Cl? secretion and an increased Na+ absorption through epithelial Na+ channels (ENaC). The present study aimed to find an effective inhibitor of human ENaC with respect to replacing amiloride therapy for CF patients. Therefore, we developed specific antisense oligonucleotides (AON) that efficiently suppress Na+ hyperabsorption by inhibiting the expression of the α‐ENaC subunit.

Methods

We heterologously expressed ENaC in oocytes of Xenopus laevis for mass screening of AON. Additionally, primary cultures of human nasal epithelia were transfected with AON and were used for Ussing chamber experiments, as well as biochemical and fluorescence optical analyses.

Results

Screening of several AON by co‐injection or sequential microinjection of AON and ENaC mRNA in X. laevis oocytes led to a sustained decrease in amiloride‐sensitive current and conductance. Using primary cultures of human nasal epithelia, we show that AON effectively suppress amiloride‐sensitive Na+ absorption mediated by ENaC in CF and non‐CF tissues. In western blot experiments, it could be shown that the amount of ENaC protein is effectively reduced after AON transfection.

Conclusions

Our data comprise an initial step towards a preclinical test with AON to reduce Na+ hyperabsorption in CF epithelia. Copyright © 2009 John Wiley & Sons, Ltd.
  相似文献   

14.

Background  

Ca2+ is known to be involved in a number of metastatic processes including motility and proliferation which can result in store-depletion of Ca2+. Up regulation of genes which contribute to store operated channel (SOC) activity may plausibly be necessary for these processes to take place efficiently. TRPC proteins constitute a family of conserved Ca2+-permeable channels that have been shown to contribute to SOC activity.  相似文献   

15.
16.
Recently we have reported that the αC-helix in the cyclic nucleotide binding domain (CNBD) is required for channel regulation and function of cyclic nucleotide gated ion channels (CNGCs) in Arabidopsis. A mutation at arginine 557 to cysteine (R557C) in the αC-helix of the CNBD caused an alteration in channel regulation. Protein sequence alignments revealed that R557 is located in a region that is important for calmodulin (CaM) binding. It has been hypothesized that CaM negatively regulates plant CNGCs similar to their counter parts in animals. However, only a handful of studies has been published so far and we still do not have much information about the regulation of CNGCs by CaM. Here, we conducted in silico binding prediction of CaM and Arabidopsis CNGC12 (AtCNGC12) to further study the role of R557. Our analysis revealed that R557 forms salt bridges with both D79 and E83 in AtCaM1. Interestingly, a mutation of R557 to C causes the loss of these salt bridges. Our data further suggests that this alteration in CaM binding causes the observed altered channel regulation and that R557 plays an important role in CaM binding.Key words: calmodulin, CaM, cyclic nucleotide gated ion channels, CNGC, environmental effect, defense responses, temperatureWe recently reported about the role of the αC-helix in the cyclic nucleotide binding domain (CNBD) for channel regulation and function of cyclic nucleotide gated ion channels (CNGCs) in Arabidopsis.1 CNGCs were first discovered in vertebrate retinal photoreceptors and olfactory sensory neurons.2,3 They are composed of a cytoplasmic N-terminus, six membrane spanning regions (S1–S6), a pore domain located between S5 and S6 and a cytoplasmic C-terminus and share similarities with the voltage-gated outward rectifying K+-selective ion channel (Shaker) proteins.2 However, CNGCs are ligand gated and opened by the direct binding of cyclic nucleotides monophosphates (cNMPs), such as cAMP and cGMP (cNMPs).4 The cytoplasmic C-terminus contains a CNBD that is connected to the pore domain by a C-linker region.5 The function and regulation of CNGCs has been extensively studied in retinal photoreceptor and olfactory sensory neurons and it has been reported that their channel activity is moderated by Ca2+/calmodulin (CaM).6 A variety of Ca2+ permeable channels are known to be regulated by Ca2+/CaM. In many cases, CaM downregulates their activity, thereby creating negative feedback regulation for Ca2+ entry.7 CNGCs are also considered to follow this mode of regulation. Some animal CNGCs possess a CaM binding domain in the cytoplasmic N-terminus. For example, it has been reported that CaM binds to a short segment in the N-terminal region of the A2 subunit of CNGCs of olfactory sensory neurons in a Ca2+ dependent manner.8,9The first plant CNGC, HvCBT1 (Hordeum vulgare calmodulin (CaM)-binding transporter), was identified as a CaM-binding protein in barley.10 Interestingly, in contrast to animal CNGCs, the CaM binding domain in HvCBT1 was shown to be located at the cytoplasmic C-terminal region.10 Subsequently, several CNGCs were identified from Arabidopsis and Nicotiana tabacum11,12 and some of these CNGCs have also been shown to possess the CaM binding domain in the C-terminal region that partially overlaps with the CNBD.1114 This difference in the location of the CaM binding domain between animal and plant CNGCs implies that different mechanisms for CNGC regulation may have evolved in animals and plants.The Arabidopsis mutant constitutive expresser of PR genes 22 (cpr22), which contains a novel chimeric CNGC, AtCNGC11/12, shows environmentally sensitive constitutive defense responses, such as heightened salicylic acid accumulation and hypersensitive response-like spontaneous programmed cell death.15,16 It has been reported that the expression of AtCNGC11/12 and its channel activity is attributable for the cpr22 phenotype.17,18 A genetic screen for mutants that suppress cpr22-conferred phenotypes identified over 20 novel mutant alleles in AtCNGC11/12.1,18 These intragenic mutants are excellent tools to study the structure-function relationship of plant CNGCs. One of these mutants, suppressor S58, possesses a single amino acid substitution, arginine 557 to cysteine (R557C), in the αC-helix of the CNBD. The suppressor S58 lost all cpr22 related phenotypes, such as spontaneous cell death formation, under ambient temperature conditions. However, these phenotypes were recovered at 16°C, suggesting that the stability of channel function is affected by temperature.1 Interestingly, this temperature sensitivity was also observed in the original mutant, cpr22.19 All salicylic acid-dependent phenotypes of cpr22 are enhanced under low temperature and low humidity conditions. Furthermore, this type of environmental sensitivity has been reported not only for cpr22, but also for various other pathogen resistance mutants as well as for defense responses in wild type plants.16 Therefore, it is possible that the basis of the temperature sensitivity observed in S58 may be related to a general environmental sensitivity in defense responses.Characterization of S58 and functional complementation using heterologous expression analyses suggested that R557 in the αC-helix of the CNBD is important for channel regulation, but not for basic channel function.1 To further investigate this, we aimed to elucidate the molecular mechanism by which R557C (S58 mutation) alters channel regulation to determine the functional role of R557. Since R557 is located in the CNBD, it is possible that R557C alters cNMP binding resulting in disruption of channel opening. In animal systems, it has been reported that cNMPs bind within the pocket formed by the αC-helix and the β-barrel that is composed of the eight β sheets in the CNBD.20,21 The αC-helix was suggested to function as the lid of this pocket that stabilizes cNMP binding by forming hydrophobic interactions with the bound cNMP.21 However, it is unlikely that R557 interacts directly with the bound cNMP because of its hydrophilic nature.Interestingly, a 19–20 amino acid sequence of the αC-helix was suggested to function as CaM binding domain in AtCNGC1 and AtCNGC2 by Köhler and Neuhaus using yeast two hybrid analysis.14 Arazi et al.13 biochemically demonstrated that a 23 amino acid sequence that overlaps with this 19–20 amino acids is the CaM binding domain in the tobacco CNGC, NtCBP4. Furthermore, they reported that the 4 additional amino acids (W R T/S W) which are located just outside of the 19–20 amino acid sequence are crucial for efficient binding to CaM. Sequence alignment revealed that R557 is located in this crucial sequence (W R T/S W).1 Thus, it can be hypothesized that the R557C mutation causes a modification in the binding affinity to CaM resulting in an alteration in channel regulation. Therefore, we generated in silico models of CaM binding with the αC-helix of the CNBD in AtCNGC12 (identical to AtCNGC11/12) and AtCNGC12:R557C (S58). We first modeled Arabidopsis CaM1 (AtCaM1) based on the crystal structure of a potato CaM, PCM6 (PDB# 1RFJ).22 There are seven different CaM genes in Arabidopsis that encode two sets of identical isoforms (CaM1 and CaM4; CaM2, CaM3 and CaM5) and two additional distinct isoforms (CaM6 and CaM7).23,24 All of them share high similarity with PCM6. Using the AtCaM1 model, we modeled possible interactions between AtCaM1 and the αC-helix of the CNBD in AtCNGC12 or AtCNGC12:R557C. As shown in Figure 1 (center part), a hydrophobic pocket of CaM that is necessary for binding with target proteins25 was seen in AtCaM1 creating a binding pocket for the αC-helix of the CNBD in AtCNGC12. In this model, R557 creates salt bridges with both D79 and E83 of AtCaM1 (indicated by a box) and these salt bridges appear to play a role for binding to CaM. This type of salt bridges have been reported to be crucial for CaM binding with several different target proteins.26 Interestingly, these salt bridges are no longer seen in AtCNGC12:R557C (Fig. 1, right part, indicated by a box); indicating that the mutation may cause an affinity change in CaM binding. Such an affinity change will likely cause an alteration in channel regulation. Since CaM is likely a negative regulator of CNGCs, this alteration in CaM affinity does not provide a simple mechanism for the R557C mutation. However, the change in CaM binding affinity may cause complex regulatory changes in CNGCs. Thus, we are currently conducting various biochemical analyses to validate this hypothesis.Open in a separate windowFigure 1Computational structural modeling of CaM binding with AtCNGC12 and AtCNGC12:R557C. Modeling of the tertiary structure of AtCaM1, and the αC-helix of AtCNGC12 and AtCNGC12:R557 was conducted using the crystallized structures of the potato CaM, PCM 6 (PDB# 1RFJ)22 and the cytoplasmic C-terminus of the invertebrate CNGC, SpIH (Flynn et al. 2007, PDB# 2PTM),27 respectively, as templates. The protein fold recognition server (Phyre)28 was used to model these proteins. The binding modeling was performed using an algorithm for molecular docking (PatchDock).29 All the images were generated using PyMOL.30 CaM is colored in cyan and the αC-helix is shown in magenta. Left part: overall binding model between AtCaM1 and AtCNGC12, Center part: close up of the boxed area of the left part in AtCNGC12, Right part: the same area in AtCNGC12:R557C. M73, M52 and M37 of AtCaM1 create a hydrophobic pocket together with F562 and I564 of AtCNGC12, which is a typical binding configuration between CaM and target proteins. R557 creates salt bridges with both D79 and E83 (center part). These salt bridges are no longer seen between AtCaM1 and AtCNGC12:R557C (right part).Although plants CNGCs have only recently been revealed to mediate multiple stress responses and also play important roles in some developmental pathways, studies that aim to elucidate their structural and regulatory properties are still very much in their infancy. Our current study will certainly contribute to a better understanding of the structure-function relationship and regulation of plant CNGCs.  相似文献   

17.

Background

Outer hair cells are the specialized sensory cells that empower the mammalian hearing organ, the cochlea, with its remarkable sensitivity and frequency selectivity. Sound-evoked receptor potentials in outer hair cells are shaped by both voltage-gated K+ channels that control the membrane potential and also ligand-gated K+ channels involved in the cholinergic efferent modulation of the membrane potential. The objectives of this study were to investigate the tonotopic contribution of BK channels to voltage- and ligand-gated currents in mature outer hair cells from the rat cochlea.

Methodology/Principal

Findings In this work we used patch clamp electrophysiology and immunofluorescence in tonotopically defined segments of the rat cochlea to determine the contribution of BK channels to voltage- and ligand-gated currents in outer hair cells. Although voltage and ligand-gated currents have been investigated previously in hair cells from the rat cochlea, little is known about their tonotopic distribution or potential contribution to efferent inhibition. We found that apical (low frequency) outer hair cells had no BK channel immunoreactivity and little or no BK current. In marked contrast, basal (high frequency) outer hair cells had abundant BK channel immunoreactivity and BK currents contributed significantly to both voltage-gated and ACh-evoked K+ currents.

Conclusions/Significance

Our findings suggest that basal (high frequency) outer hair cells may employ an alternative mechanism of efferent inhibition mediated by BK channels instead of SK2 channels. Thus, efferent synapses may use different mechanisms of action both developmentally and tonotopically to support high frequency audition. High frequency audition has required various functional specializations of the mammalian cochlea, and as shown in our work, may include the utilization of BK channels at efferent synapses. This mechanism of efferent inhibition may be related to the unique acetylcholine receptors that have evolved in mammalian hair cells compared to those of other vertebrates.  相似文献   

18.

Background

Sperm have but one purpose, to fertilize an egg. In various species including Drosophila melanogaster female sperm storage is a necessary step in the reproductive process. Amo is a homolog of the human transient receptor potential channel TRPP2 (also known as PKD2), which is mutated in autosomal dominant polycystic kidney disease. In flies Amo is required for sperm storage. Drosophila males with Amo mutations produce motile sperm that are transferred to the uterus but they do not reach the female storage organs. Therefore Amo appears to be a mediator of directed sperm motility in the female reproductive tract but the underlying mechanism is unknown.

Methodology/Principal Findings

Amo exhibits a unique expression pattern during spermatogenesis. In spermatocytes, Amo is restricted to the endoplasmic reticulum (ER) whereas in mature sperm, Amo clusters at the distal tip of the sperm tail. Here we show that flagellar localization of Amo is required for sperm storage. This raised the question of how Amo at the rear end of sperm regulates forward movement into the storage organs. In order to address this question, we used in vivo imaging of dual labelled sperm to demonstrate that Drosophila sperm navigate backwards in the female reproductive tract. In addition, we show that sperm exhibit hyperactivation upon transfer to the uterus. Amo mutant sperm remain capable of reverse motility but fail to display hyperactivation and directed movement, suggesting that these functions are required for sperm storage in flies.

Conclusions/Significance

Amo is part of a signalling complex at the leading edge of the sperm tail that modulates flagellar beating and that guides a backwards path into the storage organs. Our data support an evolutionarily conserved role for TRPP2 channels in cilia.  相似文献   

19.

Background

Dimebon is an antihistamine compound with a long history of clinical use in Russia. Recently, Dimebon has been proposed to be useful for treating neurodegenerative disorders. It has demonstrated efficacy in phase II Alzheimer's disease (AD) and Huntington's disease (HD) clinical trials. The mechanisms responsible for the beneficial actions of Dimebon in AD and HD remain unclear. It has been suggested that Dimebon may act by blocking NMDA receptors or voltage-gated Ca2+ channels and by preventing mitochondrial permeability pore transition.

Results

We evaluated the effects of Dimebon in experiments with primary striatal neuronal cultures (MSN) from wild type (WT) mice and YAC128 HD transgenic mice. We found that Dimebon acts as an inhibitor of NMDA receptors (IC50 = 10 μM) and voltage-gated calcium channels (IC50 = 50 μM) in WT and YAC128 MSN. We further found that application of 50 μM Dimebon stabilized glutamate-induced Ca2+ signals in YAC128 MSN and protected cultured YAC128 MSN from glutamate-induced apoptosis. Lower concentrations of Dimebon (5 μM and 10 μM) did not stabilize glutamate-induced Ca2+ signals and did not exert neuroprotective effects in experiments with YAC128 MSN. Evaluation of Dimebon against a set of biochemical targets indicated that Dimebon inhibits α-Adrenergic receptors (α1A, α1B, α1D, and α2A), Histamine H1 and H2 receptors and Serotonin 5-HT2c, 5-HT5A, 5-HT6 receptors with high affinity. Dimebon also had significant effect on a number of additional receptors.

Conclusion

Our results suggest that Ca2+ and mitochondria stabilizing effects may, in part, be responsible for beneficial clinical effects of Dimebon. However, the high concentrations of Dimebon required to achieve Ca2+ stabilizing and neuroprotective effects in our in vitro studies (50 μM) indicate that properties of Dimebon as cognitive enhancer are most likely due to potent inhibition of H1 histamine receptors. It is also possible that Dimebon acts on novel high affinity targets not present in cultured MSN preparation. Unbiased evaluation of Dimebon against a set of biochemical targets indicated that Dimebon efficiently inhibited a number of additional receptors. Potential interactions with these receptors need to be considered in interpretation of results obtained with Dimebon in clinical trials.  相似文献   

20.
1. Caffeine at 0.3–10 mM enhanced the binding of [3H]ryanodine to calcium-release channels of rabbit muscle sarcoplasmic reticulum. A variety of other xanthines were as efficacious as caffeine or nearly so, but none appeared markedly more potent.2. Caffeine at 1 mM markedly inhibited binding of [3H]diazepam to GABAA receptors in rat cerebral cortical membranes.3. Other xanthines also inhibited binding with certain dimethylpropargylxanthines being nearly fivefold more potent than caffeine.4. Caffeine at 1 mM stimulated binding of [35S]TBPS to GABAA receptors as did certain other xanthines.5. The dimethylpropargylxanthines had little effect. 1,3-Dipropy1-8-cyclopentylxan- thine at 100 M had no effect on [3H]diazepam binding, but markedly inhibited [35S]TBPS binding.6. Structure–activity relationships for xanthines do differ for calcium-release channels and and for different sites on GABAA receptors, but no highly selective lead compounds were identified.  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