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 46 毫秒
1.
Calmodulin regulation (calmodulation) of the family of voltage-gated CaV1-2 channels comprises a prominent prototype for ion channel regulation, remarkable for its powerful Ca2+ sensing capabilities, deep in elegant mechanistic lessons, and rich in biological and therapeutic implications. This field thereby resides squarely at the epicenter of Ca2+ signaling biology, ion channel biophysics, and therapeutic advance. This review summarizes the historical development of ideas in this field, the scope and richly patterned organization of Ca2+ feedback behaviors encompassed by this system, and the long-standing challenges and recent developments in discerning a molecular basis for calmodulation. We conclude by highlighting the considerable synergy between mechanism, biological insight, and promising therapeutics.The ancient Ca2+ sensor protein calmodulin (CaM) has emerged as a pervasive modulator of ion channels (Saimi and Kung, 2002), manifesting remarkable Ca2+ sensing capabilities in this context (Dunlap, 2007; Tadross et al., 2008) and furnishing essential Ca2+ feedback in numerous biological settings (Alseikhan et al., 2002; Xu and Wu, 2005; Mahajan et al., 2008; Adams et al., 2010; Morotti et al., 2012). Such modulation was discovered via mutations in CaM that altered the motile behavior of Paramecium, stemming from blunted activation of Ca2+-dependent Na+ current or K+ current (Kink et al., 1990). Since then, numerous ion channels have been found to be modulated by CaM, as extensively reviewed (Budde et al., 2002; Saimi and Kung, 2002; Wei et al., 2003; Halling et al., 2006; Tan et al., 2011; Nyegaard et al., 2012). This paper focuses on the CaM regulation (calmodulation) of voltage-gated Ca2+ channels, which are extensively regulated by intracellular Ca2+ binding to CaM (Ca2+/CaM; Lee et al., 1999; Peterson et al., 1999; Zühlke et al., 1999; Haeseleer et al., 2000; DeMaria et al., 2001; Budde et al., 2002; Halling et al., 2006). Such feedback regulation by Ca2+/CaM demonstrates versatile functional capabilities, represents a prominent prototype for ion-channel modulation, and holds far-reaching biological consequences. These attributes contribute much to the standing of voltage-gated Ca2+ channels as the queen of ion channels, molecules that not only sculpt membrane electrical waveforms but also serve as gatekeepers of the ubiquitous second messenger Ca2+ (Yue, 2004).

Classic history of discovery

The earliest signs of Ca2+-dependent regulation of Ca2+ channels came from data showing that increased Ca2+ could accelerate the inactivation of Ca2+ currents in Paramecium (Brehm and Eckert, 1978), invertebrate neurons (Tillotson, 1979), and insect muscle (Ashcroft and Stanfield, 1981). These results gave birth to the concept of Ca2+-dependent inactivation (CDI), the then counterintuitive notion that entities other than transmembrane voltage could modulate the rapid gating of ion channels (Eckert and Tillotson, 1981).Methodological advances permitting routine isolation and recording of Ca2+ currents within vertebrate preparations clearly revealed the existence of CDI of cardiac L-type Ca2+ currents (carried by CaV1.2 channels) in both multicellular preparations (Kass and Sanguinetti, 1984; Mentrard et al., 1984) and isolated myocytes (Lee et al., 1985). Classic data illustrating CDI of cardiac L-type currents is shown in Fig. 1 (A–C). Fig. 1 A (Ca) shows Ca2+ current from frog cardiocytes, elicited by a test voltage pulse to +80 mV (Mentrard et al., 1984). CDI becomes apparent upon Ca2+ entry during a voltage prepulse to +80 mV (Fig. 1 B), which causes sharply attenuated Ca2+ current in a subsequent test pulse (diminished area of red shading). To exclude voltage depolarization itself as the cause of inactivation, a prepulse to +120 mV (Fig. 1 C) can be demonstrated to produce far less inactivation within a subsequent test-pulse current. Because this prepulse imposed strong depolarization, but diminished Ca2+ entry via decreased driving force, the reemergence of test Ca2+ current suggests that the strong inactivation evident in Fig. 1 B could be attributed to CDI. Data of this type specifies an inactivation mechanism that depends on voltage as a U-shaped function, now considered a hallmark of CDI (Eckert and Tillotson, 1981).Open in a separate windowFigure 1.Classic U-shaped signature of CDI. (A–C) Early voltage-clamp recordings from a multicellular preparation of frog atrial trabeculae cells demonstrates CDI of Ca2+ currents. All listed voltages are relative to resting potential ER (∼−80 mV) Adapted from Mentrard et al. (1984). (A) Ca2+ currents (shaded pink) evoked in response to an +80-mV depolarizing test pulse show robust inactivation. Fast capacitive transient was estimated by blocking these Ca2+ currents with 3 mM Co2+ solution (Co). Vertical bar corresponds to 0.5 µA of current; horizontal bar, 100 ms. (B) The Ca2+ current in response to the test pulse is sharply attenuated when preceded by an +80 mV prepulse. This reduction in current amplitude reflects the inactivation of Ca2+ channels. Format is as in A. (C) Further increase in the prepulse potential to +120 mV restores Ca2+ current amplitude (compare with B). Here, the diminished Ca2+ entry during the prepulse was insufficient to trigger CDI. The format is as in A. (D–F) Single channel recordings of L-type Ca2+ channels from adult rat ventricular myocytes also exhibit U-shaped dependence of Ca2+ channel inactivation. Adapted from Imredy and Yue (1994) with permission from Elsevier. (D) Depolarizing voltage pulse to +20 mV evokes representative elementary Ca2+ currents. Ensemble average currents are shown on the bottom. (E) When depolarizing pulse was preceded by +40 mV prepulse, the elementary Ca2+ currents are sparser and the first opening is delayed. This reduction in open probability parallels the sharp attenuation of macroscopic Ca2+ currents seen in B. The ensemble average is shown on the bottom. (F) Further increase in prepulse voltage to +120 mV led to a reversal of this gating pattern, once again highlighting the U-shaped dependence of CDI. Ensemble average is shown on the bottom.Still, the actual nature of effects of Ca2+ upon gating remained unclear. Fig. 1 (D–F) reproduces the profile of CDI at the single-molecule level (Yue et al., 1990; Imredy and Yue, 1992, 1994). Resolving data such as these is technically challenging, owing to the diminutive size of unitary Ca2+ currents (∼0.3 pA) and the sub-millisecond gating timescale involved. These data demonstrate the existence of U-shaped inactivation in the gating of a single cardiac L-type Ca2+ channel fluxing Ca2+, as present in an adult rat ventricular myocyte (Imredy and Yue, 1994). These results established that Ca2+ influx of a single channel suffices to trigger CDI, and demonstrated the clear correspondence of single-channel and multicellular behavior (Fig. 1). Thus, CDI emerged as a legitimate molecular-level modulatory process.

Advent of Ca2+ channel calmodulation

The Ca2+ sensor mediating Ca2+ regulation of Ca2+ channels has long been a matter of debate, with proposals of direct binding of Ca2+ to the channel complex (Standen and Stanfield, 1982; Plant et al., 1983; Eckert and Chad, 1984), and of Ca2+-dependent phosphorylation and/or dephosphorylation of channels (Chad and Eckert, 1986; Armstrong, D., C. Erxleben, D. Kalman, Y. Lai, A. Nairn, and P. Greengard. 1988. Abstracts of papers presented at the forty-second annual meeting of the Society of General Physiologists. Abstr. 20). The road toward identification of the actual Ca2+ sensor began with the cloning and expression of recombinant Ca2+ channels, thereby enabling structure–function studies (Snutch and Reiner, 1992). This phase of discovery was initiated by a bioinformatic observation (Babitch, 1990), pointing out that the carboxy terminus of voltage-gated Ca2+ channels contains a region with weak homology to an EF hand Ca2+-binding motif (Fig. 2 A, EF1). After the recognition of EF1, the existence of a second EF hand–like motif became apparent (Fig. 2 A, EF2). This led to chimeric channel analysis (de Leon et al., 1995), wherein segments of the carboxy termini of L-type (CaV1.2) and R-type (CaV2.3) Ca2+ channels were swapped. These two types of channels exhibit strong and weak CDI, respectively, under conditions of strong intracellular Ca2+ buffering (Liang et al., 2003), making them advantageous for chimeric analysis. Results from these experiments revealed that the proximal third of the carboxy terminus (Fig. 2 A, CI region) was important for CDI, and that EF1 of CaV1.2 channels was essential for the strong CDI in these channels (de Leon et al., 1995). One possible explanation was that EF1 might be the Ca2+ binding site for CDI postulated previously (Standen and Stanfield, 1982; Plant et al., 1983; Eckert and Chad, 1984). However, mutations within the Ca2+ binding loop of EF1 failed to eliminate CDI (Zhou et al., 1997; Peterson et al., 2000), focusing the search for the Ca2+ sensor elsewhere.Open in a separate windowFigure 2.Calmodulation: the ingredients and the flavors. (A) Channel diagram illustrates overall arrangement of structural landmarks critical for CDI. The Ca2+-inactivation (CI) region, spanning ∼160 residues of the channel carboxy terminus, is highly conserved across CaV1/2 channel families and is elemental for CDI. The proximal segment (PCI) of the CI region includes the dual vestigial EF hand (EF) segments (shaded purple and blue-green). The IQ domain, a canonical CaM binding motif critical for CDI, is located just downstream of the PCI segment. The NSCaTE element in the amino terminus of CaV1.2/1.3 channels is an N-lobe Ca2+/CaM effector site. The CBD element (gray) in the carboxy terminus of CaV2 channels is a CaM binding segment thought to be critical for CDI. (B) The table outlines functional bipartition of CaM in CaV channels and the corresponding spatial Ca2+ selectivities. In CaV1.2 and CaV1.3 channels, both the C-lobe and N-lobe of CaM enable fast CDI with local Ca2+ selectivity. Latent CDI of CaV1.4 channels is revealed upon deletion of an autoinhibitory domain in the distal carboxy terminus. Throughout the CaV2 channel family, the N-lobe of CaM evokes slow CDI with global Ca2+ spatial selectivity. In CaV2.1 channels, the C-lobe of CaM supports ultra-fast facilitation (CDF) with local Ca2+ selectivity. No C-lobe triggered modulation has been reported for CaV2.2 and CaV2.3 channels.Targets of CaM binding themselves often resemble CaM, a bilobed molecule with each lobe composed of two EF hands (Jarrett and Madhavan, 1991). The dual vestigial EF hands in the carboxy terminus of channels (Fig. 2 A) may be seen as resembling a lobe of CaM, which suggests that CaM itself might be the Ca2+ sensor for CDI. Initial studies exploring this notion, however, showed that pharmacological inhibition of CaM did not eliminate CDI of L-type Ca2+ channels (Imredy and Yue, 1994; Zühlke and Reuter, 1998). Nonetheless, deletions within the CaV1.2 channel CI region identified an IQ motif (Fig. 2 A) as a critical element in CDI (Zühlke and Reuter, 1998), and mutagenesis of this IQ domain weakens CDI (Qin et al., 1999). As IQ motifs often bind Ca2+-free CaM (apoCaM; Jurado et al., 1999), the possibility that CaM acts as a CDI sensor reemerged. Definitive evidence came with studies involving a Ca2+-insensitive mutant form of CaM (CaM1234), in which point mutations within all four EF hands eliminate Ca2+ binding (Xia et al., 1998). If “preassociation” of apoCaM with target molecules were a prerequisite for subsequent regulation by Ca2+ binding to this apoCaM, then CaM1234 could act as a dominant negative to silence regulation, as seen for small-conductance Ca2+-activated K channels (Xia et al., 1998). Indeed, when CaV1.2 channels were coexpressed with CaM1234 or its analogues, CDI was ablated or strongly suppressed (Peterson et al., 1999; Zühlke et al., 1999). Additionally, apoCaM was found to bind to the CaV1.2 CI region, with critical dependence upon the IQ segment (Erickson et al., 2001; Pitt et al., 2001; Erickson et al., 2003). In retrospect, apoCaM preassociation with CaV1.2 channels would sterically protect CaM from pharmacological effects (Dasgupta et al., 1989), rationalizing the prior insensitivity of CDI to such small-molecule perturbation. In all, these data firmly substantiated CaM as the sensor for CaV1.2 channel Ca2+ regulation.Initially, it was believed that only CaV1.2 L-type channels (Fig. 2 B) were subject to CaM-mediated Ca2+ regulation (Zamponi, 2003). However, this system of calmodulation was gradually recognized to pertain to most of the CaV1 and CaV2 (but not CaV3) branches of the CaV channel superfamily (Liang et al., 2003; Fig. 2 B). P-type (CaV2.1) channels were found to be Ca2+ regulated (Lee et al., 1999), and a Ca2+/CaM binding site downstream of the IQ element (CBD; Fig. 2 A) was argued to be important. In this initial study, no role for CaM preassociation was recognized, and no role for the IQ domain was found. Moreover, Ca2+ regulation of CaV2.1 channels manifests as a facilitation of current (Ca2+-dependent facilitation; CDF), followed by a slowly developing CDI. Thus, it appeared possible that the Ca2+ regulation of CaV2.1 channels might diverge mechanistically from that of CaV1.2 channels. However, apoCaM was later found to preassociate with CaV2.1 channels (Erickson et al., 2001), Ca2+-insensitive mutant CaM molecules were found to eliminate Ca2+ regulation in CaV2.1 channels (DeMaria et al., 2001; Lee et al., 2003), and the IQ domain was determined to be structurally essential for this regulation (DeMaria et al., 2001; Lee et al., 2003; Kim et al., 2008; Mori et al., 2008). That said, the role of the CBD segment remains contentious, with some studies still arguing for this segment’s importance in CDI (Lee et al., 2003). In contrast, deletion of the entire carboxy terminus after the IQ domain (including the CBD element) completely spared CDF and CDI of CaV2.1 channels (DeMaria et al., 2001; Chaudhuri et al., 2005). Overall, CaV2.1 and CaV1.2 channels appear to be Ca2+ regulated by a largely conserved scheme.Soon thereafter, calmodulation was established for the remaining members of the CaV2 class of channels (Fig. 2 B, CaV2.2 and CaV2.3), which indicates that this modulatory system pertains throughout the CaV1-2 channel superfamily (Liang et al., 2003). Strong CaM-mediated CDI was found in CaV1.3 channels (Xu and Lipscombe, 2001; Shen et al., 2006; Yang et al., 2006); a latent capacity for CaM-mediated CDI was observed in CaV1.4 channels (Singh et al., 2006; Wahl-Schott et al., 2006); and potential indications of CaM-mediated regulation of CaV1.1 channels have been described (Stroffekova, 2008, 2011).A few more nuanced points nonetheless merit attention. First, some types of calmodulation are insensitive to strong intracellular Ca2+ buffering (e.g., CaV1.2 CDI), whereas others are not (CaV2.3 CDI; Fig. 2 B). This contrasting sensitivity enables the use of chimeric channels under increased Ca2+ buffering to identify key structural determinants (de Leon et al., 1995), despite the existence of calmodulation across the channel superfamily. Second, the mechanisms underlying CDF in CaV1.2 channels remain mysterious, as the CDF of native channels of the heart is weak to start, and attenuated by blockade of Ca2+ release from neighboring ryanodine receptor channels (RYRs; Wu et al., 2001). Additionally, CDF of CaV1.2 channels is strongly manifest only in recombinant channels containing point mutations within the IQ domain, and only when they are expressed in frog oocytes (Zühlke et al., 2000). Curiously, CDF is not observed in recombinant CaV1.2 channels expressed in mammalian cell lines (Peterson et al., 1999). In contrast, CDI of CaV1.2 channels is strong and universally observed across experimental platforms. Third, Ca2+/CaM-dependent dephosphorylation by calcineurin was initially suggested as a mechanism for CDI of Ca2+ channels (Chad and Eckert, 1986); however, this proposition has remained controversial as previously reviewed (Budde et al., 2002). Recently, calcineurin was found to preassemble with the CaV1.2 channel complex through the scaffolding protein AKAP79/150 bound to the distal carboxy terminus of the channel (Oliveria et al., 2007). Furthermore, two additional maneuvers were reported to impede CDI of these channels (Oliveria et al., 2012): the overexpression of a mutant AKAP79/150 incapable of binding calcineurin, and inhibition of calcineurin activity. However, others have found that direct inhibition of calcineurin has no effect on the CDI of L-type channels within multiple neuronal preparations (Branchaw et al., 1997; Victor et al., 1997; Zeilhofer et al., 1999). Moreover, deletion of the entire distal carboxy terminus (including the AKAP79/150 binding segment) of CaV1.2 channels preserves strong CDI (de Leon et al., 1995; Erickson et al., 2001; Crump et al., 2013). Similarly, short variants of CaV1.3 channels lacking the AKAP harboring distal carboxy termini (Marshall et al., 2011) also fully support CDI (Xu and Lipscombe, 2001; Shen et al., 2006; Yang et al., 2006; Bock et al., 2011). Altogether, it may be that calcineurin and AKAP79/150 are indirect and context-dependent modulators of CDI, rather than direct effector molecules (Budde et al., 2002).

Functional bipartition of CaM and selectivity for local/global Ca2+ sources

The calmodulatory mechanism for Ca2+ channels supports remarkable forms of Ca2+ decoding. This feature echoes earlier discoveries of “functional bipartition” of CaM in Paramecium (Kink et al., 1990; Saimi and Kung, 2002). Here, “under-excitable” behavioral strains had mutations only in the N-lobe of CaM, whereas “over-excitable” strains had mutations only in the C-lobe. Thus, one lobe of CaM can mediate signaling to one set of functions, whereas the other lobe signals to an alternative set of operations. It was unclear, however, whether this bipartition extended to mammals. The requirement that Ca2+ regulation of mammalian Ca2+ channels requires apoCaM preassociation permitted direct exploration of this question. Coexpressing channels with “half mutant” CaM molecules (CaM12 and CaM34, where only C- or N-terminal lobes, respectively, bind Ca2+) revealed that the individual lobes of CaM can evoke distinct components of channel regulation (Fig. 2 B). In most CaV1 channels, the C-lobe induces a kinetically rapid phase of CDI, whereas the N-lobe yields a slower component (Peterson et al., 1999; Yang et al., 2006; Dick et al., 2008). For CaV1.2 channels, early studies that utilized high intracellular Ca2+ buffering found only minimal N-lobe–mediated CDI (Peterson et al., 1999). However, when interrogated under low intracellular Ca2+ buffering, slow but recognizable N-lobe–mediated CDI of CaV1.2 channels emerges (Dick et al., 2008; Simms et al., 2013). Most strikingly, in CaV2.1 channels, the lobes of CaM produce opposing polarities of regulation: the C-lobe of CaM triggers a kinetically rapid CDF, whereas the N-lobe evokes a slower CDI (DeMaria et al., 2001; Lee et al., 2003). CaV2.2 and CaV2.3 channels manifest CDI triggered mainly by the N-lobe of CaM (Liang et al., 2003). Whether this bipartition orchestrates divergent classes of behavior in higher-order animals remains an intriguing and open question (Wei et al., 2003).Beyond simply splitting the Ca2+ signal, however, the calmodulation of mammalian Ca2+ channels revealed that the lobes of CaM can selectively decode Ca2+ in different ways. Hints as to this capability came from the differential sensitivity of calmodulation in various channels to Ca2+ buffering. Processes evoked by the C-lobe of CaM are invariably insensitive to introduction of strong Ca2+ buffering, whereas N-lobe–mediated processes in CaV2 channels can be eliminated by the same maneuver. Strong buffering would only spare Ca2+ signals near the cytoplasmic mouth of channels where strong point-source Ca2+ influx would overwhelm buffer capacity (Neher, 1986; Stern, 1992). This argues that local Ca2+ influx through individual channels triggers C-lobe signaling; in other words, the C-lobe of CaM exhibits a “local Ca2+ selectivity.” In contrast, a spatially global elevation of Ca2+ (present in the absence of strong Ca2+ buffering) is required for N-lobe signaling of CaV2 channels (DeMaria et al., 2001; Lee et al., 2003; Liang et al., 2003); thus, the N-lobe in this context exhibits a “global selectivity.” The existence of global selectivity is notable, given that local Ca2+ influx yields far larger Ca2+ increases near a channel than does globally sourced Ca2+ (Tay et al., 2012). One exception to this pattern is the local Ca2+ selectivity of the N-lobe component of CDI in CaV1.2/1.3 channels (see two paragraphs below). Corroboration of spatial Ca2+ selectivity is provided by the ability of single CaV1.2 channels to undergo CDI (Fig. 1, D–F). By definition, only a local Ca2+ source is present in the single-channel configuration; thus, the presence of CDI in this context indicates local Ca2+ selectivity. Additionally, single-channel records of CaV2.1 channels exhibit CDF (driven by C-lobe), but not CDI (triggered by N-lobe; Chaudhuri et al., 2007), which is consistent with the proposed differential selectivities for this channel. Fig. 2 B summarizes the arrangement of spatial Ca2+ selectivities according to the lobes of CaM.The mechanisms underlying these contrasting spatial Ca2+ selectivities can be interpreted as emergent behaviors of a system in which a lobe of apoCaM must transiently detach from a preassociation site before binding Ca2+, and where a Ca2+-bound lobe of CaM must associate with a channel effector site to mediate regulation (Tadross et al., 2008). When the slow Ca2+-unbinding kinetics of the C-lobe of CaM are imposed on this scenario, local Ca2+ sensitivity invariably results. In contrast, when the rapid Ca2+-unbinding kinetics of the N-lobe are interfaced with this architecture, global Ca2+ selectivity arises if the channel preferentially binds apoCaM versus Ca2+/CaM, as in the case of CaV2 channels (Fig. 2 B). When the channel preferentially interacts with Ca2+/CaM, local Ca2+ selectivity emerges, as in the context of CaV1 channels (Fig. 2 B). The relative roles of the affinities and kinetics of Ca2+ binding to the two lobes of CaM in mediating local and global Ca2+ selectivities have been considered at length elsewhere (Tadross et al., 2008).As an explicit example of how spatial Ca2+ selectivity of N-lobe regulation is specified, we consider the effects of an N-terminal spatial Ca2+-transforming element (NSCaTE) present only on the amino terminus of CaV1.2/1.3 channels (Fig. 2 A). NSCaTE is a Ca2+/CaM binding site whose presence enhances the aggregate channel affinity for Ca2+/CaM over apoCaM (Ivanina et al., 2000; Dick et al., 2008), endowing the N-lobe component of CDI of these channels with a largely local Ca2+ selectivity of N-lobe CDI in these channels. Elimination of the NSCaTE site in these channels, which tilts channels toward a preference for apoCaM, then switches their N-lobe CDI to a global profile. Conversely, donation of NSCaTE to CaV2 channels, which causes channels to favor Ca2+/CaM binding, then endows their N-lobe CDI with local Ca2+ selectivity (Dick et al., 2008). In this manner, the presence of NSCaTE can tune the spatial Ca2+ selectivity of N-lobe–mediated channel regulation. This overall arrangement of CaM interactions with a target molecule could endow numerous other regulatory systems with spatial Ca2+ selectivity.

Molecular basis of calmodulation

Elucidating the arrangement of apoCaM and Ca2+/CaM on Ca2+ channels is fundamental for the field, given the biological influence of calmodulation, and the importance of this system as an ion-channel regulatory prototype (Dunlap, 2007). This critical task, however, remains an ongoing challenge, given the >2,000 amino acids comprising the main pore-forming α1 subunit alone, the ability of multiple peptide segments of the channel to bind CaM in vitro, and the formidable nature of obtaining atomic structures for these channels.The stoichiometry of CaM interaction with channels has been debated. Crystal structures of Ca2+/CaM complexed with portions of the carboxy tail CI region of CaV1.2 channels (Fallon et al., 2009; Kim et al., 2010) suggest that multiple CaM molecules may interact to produce the full spectrum of Ca2+ regulatory functions. Moreover, CaM can interact with multiple peptide segments of the channel (Ivanina et al., 2000; Tang et al., 2003; Zhou et al., 2005; Dick et al., 2008; Ben Johny et al., 2013). In contrast, covalent fusion of single CaM molecules to CaV1.2 channels suggests that only one CaM per channel suffices to elicit CDI (Mori et al., 2004). Moreover, live-cell FRET studies of the interactions between CaM and holochannels (CaV1.2) also point to a 1:1 CaM/channel ratio (Ben Johny et al., 2012). In the holochannel, the various CaM binding segments may be arranged in close proximity so as to allow the two lobes of CaM to preferentially interact with the distinct channel segments during Ca2+ regulation (Dick et al., 2008; Ben Johny et al., 2013). Recent biochemical studies of the NSCaTE element show that this segment preferentially interacts with a single lobe of CaM (N-lobe) at a time (Liu and Vogel, 2012). Furthermore, the NSCaTE element can interact also with Ca2+/CaM prebound to an IQ domain peptide (Taiakina et al., 2013). Thus, we favor the interpretation that only one CaM interacts within holochannels, although peptide fragments of Ca2+ channels may bind to multiple CaM molecules.The IQ domain (Fig. 3 A, blue circle) is critical for calmodulation. Mutations within this domain markedly modulate the Ca2+ regulation of CaV1.2 channels (Zühlke and Reuter, 1998; Qin et al., 1999; Zühlke et al., 1999, 2000; Erickson et al., 2003), CaV1.3 channels (Yang et al., 2006; Bazzazi et al., 2013; Ben Johny et al., 2013), CaV2.1 channels (DeMaria et al., 2001; Lee et al., 2003; Kim et al., 2008; Mori et al., 2008), and CaV2.2 and CaV2.3 channels (Liang et al., 2003). Furthermore, apoCaM preassociation with CaV1.2/1.3 channels requires the IQ domain (Erickson et al., 2001; Pitt et al., 2001; Erickson et al., 2003; Bazzazi et al., 2013; Ben Johny et al., 2013). Finally, Ca2+/CaM binds well to IQ-domain peptides of many CaV channels (Peterson et al., 1999; Zühlke et al., 1999; DeMaria et al., 2001; Pitt et al., 2001; Liang et al., 2003; Bazzazi et al., 2013; Ben Johny et al., 2013). All these results gave rise to the IQ-centric hypothesis shown in Fig. 3 A. Here, the IQ domain is important for both apoCaM preassociation (left) and Ca2+/CaM binding (effector configuration shown on the right). This IQ-centric paradigm has motivated efforts to resolve crystal structures of Ca2+/CaM complexed with IQ-domain peptides of various CaV1-2 channels (Fallon et al., 2005; Van Petegem et al., 2005; Kim et al., 2008; Mori et al., 2008), as shown in Fig. 3 (B and C). Throughout, the IQ peptide segment appears as an α-helical entity with N and C termini as labeled.Open in a separate windowFigure 3.Toward an atomic-level understanding of CDI. (A) IQ-centric view of calmodulation of CaV channels. In this mechanistic scheme, ApoCaM is preassociated with the IQ domain (blue). Upon Ca2+ binding, CaM rebinds the same IQ domain with a higher affinity, and subtle conformational rearrangements are presumed to trigger CaM-mediated channel regulation. (B, left) Crystal structure of CaV1.2 IQ domain peptide in complex with Ca2+/CaM (Protein Data Bank accession no. 2BE6). The IQ domain is colored blue. CaM is show in cyan (N-lobe, pale cyan; C-lobe, cyan). Ca2+ ions are depicted as yellow spheres. The key isoleucine residue (red) serves as a hydrophobic anchor for CaM. Ca2+/CaM adopts a parallel arrangement with the IQ domain in which the N-lobe binds closer to the amino terminus of the IQ domain and the C-lobe binds further downstream. (B, right) Crystal structure of Ca2+/CaM bound to mutant CaV1.2 IQ domain with alanines substituted for the key isoleucine-glutamine residues (accession no. 2F3Z). This double mutation abolishes CDI. Structurally, however, Ca2+/CaM hugs the mutant IQ domain in a similar conformation as to its interaction with the wild-type (left). (C) Crystal structure of Ca2+/CaM bound to CaV2.1 IQ domain (left, accession no. 3BXK; right, accession no. 3DVM). The IQ domain is shaded green, CaM in cyan. Ca2+/CaM was reported to bind to the CaV2.1 IQ domain in both parallel (left) and antiparallel (right) arrangements. The antiparallel arrangement in which the C-lobe of CaM binds upstream of IQ domain has led to speculation that CDF may result from this inverted polarity of Ca2+/CaM association. (D) Simplified configurations of the CaM/channel complex relevant for calmodulation (E, A, IC, IN, and ICN). In configuration E, channels lack preassociated CaM and therefore cannot undergo CDI. Configuration A corresponds to channels bound to apoCaM, and thus capable of undergoing robust CDI. Ca2+ binding to the C-lobe and N-lobe of CaM leads to the inactivated configurations IC and IN, respectively. Ca2+ binding to both lobes of CaM then yields configuration ICN, with both C and N lobes of CaM engaged toward CDI. This final transition may exhibit positive cooperativity as specified by the constant λ >> 1. Under endogenous levels of CaM, channels may reside in any of the five configurations. Strong coexpression of wild-type or mutant CaM restricts accessibility to various states. Reproduced with permission from Ben Johny et al., 2013.However, this viewpoint remains problematic in three regards. (1) The atomic structures of Ca2+/CaM bound to wild-type and mutant IQ peptides of CaV1.2 (Fig. 3 B) show that a central isoleucine (side chains explicitly shown) in the IQ element is deeply buried within the C-lobe of Ca2+/CaM (Van Petegem et al., 2005), and that alanine substitution at this site hardly changes structure (Fallon et al., 2005). Additionally, Ca2+/CaM dissociation constants for corresponding wild-type and mutant IQ peptides are nearly the same (Zühlke et al., 2000; Bazzazi et al., 2013). Why then would alanine substitution at this well-ensconced site influence the rest of the channel to blunt regulation (Fallon et al., 2005)? (2) For CaV1.2/1.3 channels, the N-lobe of Ca2+/CaM effector site appears to be an NSCaTE element (Fig. 3 A, oval) of the channel amino terminus (Dick et al., 2008; Tadross et al., 2008; Liu and Vogel, 2012), separate from the IQ element. (3) Crystal structures of Ca2+/CaM in complex with the IQ peptide of CaV2.1 channels show that CaM can adopt both a parallel (Mori et al., 2008) and an antiparallel configuration (Kim et al., 2008; Fig. 3 C, left and right, respectively). The apparent inversion in configuration of CaM binding to IQ domain between CaV1.2 and CaV2.1 has been proposed as a mechanism for the opposing polarity of Ca2+ regulation observed in the two channels (Kim et al., 2008; Minor and Findeisen, 2010). However, detailed structural analysis along with functional systematic alanine scanning mutagenesis and chimeric channel analysis argues that the C-lobe effector site resides at a site beyond the IQ module (Mori et al., 2008).A major concern with older IQ domain analyses is that the regulatory system was not considered conceptually as a whole, as shown in Fig. 3 D (drawn with specific reference to CaV1.3 channels; Ben Johny et al., 2013). Configuration E portrays channels lacking apoCaM. Such channels can open normally, but do not manifest CDI because Ca2+/CaM from bulk solution cannot efficiently access a channel in configuration E to induce CDI (Mori et al., 2004; Yang, P.S., M.X. Mori, E.A. Antony, M.R. Tadross, and D.T. Yue. 2007. Biophysical Journal abstracts issue. 1669-Plat; Liu et al., 2010; Findeisen et al., 2011). ApoCaM binding with configuration E gives rise to configuration A, where opening can also occur normally, but CDI can now take place. For CDI, Ca2+ binding to both lobes of CaM yields configuration ICN, which underlies a fully inactivated channel with strongly reduced opening. For intermediate configurations, Ca2+ binding only to the C-lobe brings about configuration IC, equivalent to a C-lobe inactivated channel; Ca2+ binding only to the N-lobe yields the N-lobe–inactivated arrangement (IN). Of particular importance, ensuing entry into ICN likely involves positively cooperative interactions specified by cooperativity factor λ >> 1.This scheme emphasizes several challenges for older analyses of the IQ domain, where CDI was mostly assessed with only endogenous CaM present. Results thus obtained would be ambiguous, because IQ-domain mutations could alter calmodulation via changes at multiple steps depicted in Fig. 3 D, whereas other interpretations largely attributed the effects to altered Ca2+/CaM binding with an IQ effector site. In contrast, IQ mutations could well weaken apoCaM preassociation and reduce CDI by favoring configuration E. Moreover, functional deficits caused by mutations that do attenuate interaction with one lobe of Ca2+/CaM may be masked by positively cooperative steps (λ in Fig. 3 D).To minimize these issues, CDI of CaV1.3 channels was recently characterized during strong coexpression with various mutant CaM molecules (Bazzazi et al., 2013; Ben Johny et al., 2013) as shown in Fig. 4 (A–C, top). For orientation, CDI of channels expressed with only endogenous CaM present is shown in the upper portion of Fig. 4 A. Strong CDI produces a rapid decay of whole-cell Ca2+ current (red trace), compared with the negligible decline of Ba2+ current (black trace). Because Ba2+ binds CaM poorly (Chao et al., 1984), the decline of Ca2+ versus Ba2+ current after 300 ms of depolarization quantifies CDI (CDI parameter plotted below in bar graphs). One can isolate the diamond-shaped subsystem lacking configuration E (Fig. 4 B, top) by leveraging mass action with strong coexpression of wild-type CaM (CaMWT). Further simplification of CDI was obtained by strongly coexpressing channels with CaM12 (Fig. 4 C, top), which empties configuration E by mass action, and forbids configurations IN and ICN. Thus, the isolated C-lobe component of CDI can be studied, with the signature rapid time course of current decay shown near the top of Fig. 4 C. Critically, this layout avoids interplay with cooperative λ steps in Fig. 3 D. Likewise, strongly coexpressing CaM34 focuses on the slower N-lobe form of CDI, with attendant simplifications.Open in a separate windowFigure 4.Residue-level roadmap for calmodulation of CaV1.3 channels. Systematic alanine scanning mutagenesis of the entire CI domain of CaV1.3 channels. (A) CDI of wild-type and mutant CaV1.3 channels recombinantly expressed in HEK293 cells was measured with only endogenous CaM present. CDI observed here reflects properties of the entire system (stick figure) diagrammed in Fig. 3 D. Exemplar whole-cell current shows robust CDI for wild-type (WT) CaV1.3, reflecting their high affinity for apoCaM. Here and throughout, the vertical bar pertains to 0.2 nA of Ca2+ current (black); the Ba2+ current (gray) has been scaled ∼3-fold downward to aid comparison of decay kinetics. Horizontal bar, 100 ms. CDI is measured as the fractional reduction of Ca2+ current in comparison to Ba2+ at 300 ms (Ben Johny et al., 2013). Bottom, bar graph summary of CDI for alanine substitutions for indicated residues. CDI for WT channels, blue-green vertical line. Alanine substitutions in both EF hand regions and IQ domain resulted in diminished CDI. (B) Overexpression of CaMWT isolates the behavior of the diamond-shaped subsystem. Such overexpression of CaMWT should rescue CDI for mutations that weakened apoCaM preassociation. For WT CaV1.3, exemplar current shows that CDI is unaltered, as expected for a channel with high affinity for apoCaM. Bottom, bar graph summary of CDI for various mutant channels with CaMWT overexpressed. Format is as in A. CDI is rescued for several mutations in the EF hand region (EF2) and the IQ domain, reflecting the preassociation of N-lobe and C-lobe of apoCaM. (C) Overexpression of CaM12 isolates C-lobe CDI. Exemplar currents for WT channels show the fast C-lobe form of CDI. Bottom, bar graph summary shows the footprint of C-lobe CDI. Mutants in both the first EF hand and the IQ domain disrupted the C-lobe CDI (see the LGF loci in the EF hand regions and TFL and IQD loci in the IQ domain). The format is as in A. (D) Overexpression of CaM34 isolates N-lobe CDI. N-lobe CDI was largely preserved by mutations in the CI region. Mutations that weakened N-lobe apoCaM preassociation exhibited enhanced N-lobe CDI (see EF2 segment). Reassuringly, alanine scanning mutagenesis of NSCaTE element resulted in specific disruption of N-lobe CDI of CaV1.3 channels (Tadross et al., 2008). Adapted with permission from Ben Johny et al. (2013) and Bazzazi et al. (2013).Accordingly, a new framework of CaM/channel configurations that underlie calmodulation could be deduced, as diagrammed in Fig. 5 (Ben Johny et al., 2013). For convenience, Fig. 4 compiles results from a systematic alanine scan covering the entire CI domain of the carboxyl tail of CaV1.3 channels (Bazzazi et al., 2013; Ben Johny et al., 2013), where substitution positions are denoted to the left of the bar graph in Fig. 4 A. The N-terminal end of the CI segment corresponds to the top of the graph, purple and blue-green regions indicate EF1 and EF2, and the lavender zone denotes the IQ domain. Electrophysiological characterization for each of the substitutions was performed for all the various subsystems (Fig. 4, A–D, top), and bar graphs plot the strength of CDI for the corresponding subsystems in each of the panels. The blue-green vertical lines reference the profile for wild-type channels. The results thus obtained were combined with CaM binding assays (not depicted) to identify meaningful structure–function correlations.Open in a separate windowFigure 5.Next-generation blueprint for CaM/CaBP regulation of CaV channels. (A) De novo molecular model of CaV1.3 CI region docked to apoCaM. The N-lobe of apoCaM is thought to preassociate with the PCI region (green), whereas the C-lobe binds to the IQ domain (blue). The NSCaTE segment (tan) is unoccupied. This model corresponds to configuration A in Fig. 3 D, where the channels are charged with an apoCaM and ready to undergo CDI. (B) Proposed model for the Ca2+ inactivated state of the channel. The N-lobe of Ca2+/CaM is shown binding to the NSCaTE segment based on a recent NMR structure (Protein Data Bank accession no. 2LQC). This configuration is believed to result in N-lobe CDI. The de novo molecular model shows C-lobe of Ca2+/CaM, the EF hand segment, and the IQ domain forming a tripartite complex resulting in C-lobe CDI. Overall, this model corresponds to the inactive configuration ICN in Fig. 3 D. (C) Proposed allosteric mechanism of CaBP action. CaBP and apoCaM may dually bind the CaV1 channel. However, CaBP may allosterically inhibit CDI of CaV1 channels by interacting with the NT, III-IV loop, or PCI segment, thereby preventing Ca2+/CaM from reaching its effector configuration. (A–C) Adapted with permission from Yang et al. (2014).Under the regimen in Fig. 4 A, where all configurations are potentially accessible, diminished CDI by alanine substitutions occurred not only in the IQ domain, but upstream in the EF hand regions (between LGF and TLF residues). The latter effects strongly suggested that something outside the IQ domain was essential. Some of these deficits in CDI could be attributed to weakening of apoCaM preassociation with channels, because overexpressing wild-type CaM to depopulate configuration E (in Fig. 4 B) substantially recovered many of these CDI deficits. Binding assays of apoCaM with various CI segments confirmed this interpretation (Bazzazi et al., 2013; Ben Johny et al., 2013).Turning to potential deficits relating to diminished Ca2+/CaM action through channel effector sites, Fig. 4 (C and D) separately interrogates the C-lobe and N-lobe components of CDI. Importantly, isolating the C- and N-lobe components minimizes masking of mutational effects by positive cooperativity, allowing CDI deficits to be more sensitively observed in this regimen. That said, Fig. 4 C displays C-lobe CDI deficits in the EF hand region (strongest for LGF substitution), as well as in the IQ domain (with the strongest effect upon substitution at the central isoleucine). These outcomes support a hypothesis where the Ca2+/CaM effector configuration for the C-lobe component of CDI involves both of these functional hotspots, as developed two paragraphs below. For the N-lobe component of CDI (Fig. 4 D), no appreciable deficits were observed, as would be expected if interaction with the channel NSCaTE element (in the channel amino terminus) serves as the effector element.Fig. 5 A displays a proposed configuration A for apoCaM interaction with the channel. This includes a homology model of the apoCaM C-lobe complexed with the IQ domain (blue), based on analogous NaV structures (Chagot and Chazin, 2011; Feldkamp et al., 2011). The portrayal of the apoCaM N-lobe incorporates ab initio structural prediction of the CI domain, with two vestigial EF hands (EF), and a protruding helix (preIQ subelement). The EF hand module (EF) resembles the structure of a homologous NaV segment (Chagot et al., 2009; Wang et al., 2012), and the preIQ helix resembles a helical segment observed in crystal structures of analogous CaV1.2 peptides (Fallon et al., 2009; Kim et al., 2010). The atomic structure of the apoCaM N-lobe (1CFD) was interfaced with shape-complementarity docking algorithms. Regarding the proposed interaction interface of the N-lobe with the channel EF domain, we note that alanine substitution therein produced a curious enhancement of N-lobe CDI seen with certain alanine substitutions, especially in the EF region (Fig. 4 D). This feature is interesting because weakening of channel interaction with the N-lobe of apoCaM would be predicted to strengthen CDI produced by the N-lobe of Ca2+/CaM (Tadross et al., 2008).Fig. 5 B displays a proposed arrangement for the Ca2+/CaM-bound configuration ICN. The N-lobe complex with NSCaTE is an NMR structure (Liu and Vogel, 2012). An alternative ab initio model of the PCI is computationally docked with the C-lobe of Ca2+/CaM (Protein Data Bank accession no. 3BXL) and IQ module, which together form a ternary complex. This ternary arrangement is consistent with the importance of both IQ and EF domains for C-lobe CDI (Fig. 4 C). Intriguingly the C-lobe configuration resembles a rather canonical CaM–peptide complex, where the channel EF module contributes a surrogate lobe of CaM. Importantly, assays of Ca2+/CaM binding to the IQ segment alone would correlate poorly with functional effects relating to the ternary complex, as experimental studies observe (Bazzazi et al., 2013; Ben Johny et al., 2013). Overall, the proposed framework in Fig. 5 (A and B) may aid future structural biology and structure–function work. In addition, this scheme of CaM exchange within its target molecule may generalize beyond the Ca2+ channel family.

Biological consequences and prospects for new disease therapies

The biological consequences of Ca2+ regulation by CaM promise to be wide ranging and immense. In the heart, elimination of CaV1.2 CDI by means of dominant-negative CaM elicits marked prolongation of ventricular action potential duration (APD), implicating CDI as a dominant control factor in specifying APD (Alseikhan et al., 2002; Mahajan et al., 2008). As APD is one of the main determinants of electrical stability and arrhythmias in the heart, pharmacological manipulation of such regulation looms as a future antiarrhythmic strategy (Mahajan et al., 2008; Anderson and Mohler, 2009). Most recently, genome-wide linkage analysis in humans has uncovered heritable and de novo CaM mutations as the probable cause of several cases of catecholaminergic polymorphic ventricular tachycardia (CPVT), with altered CaM-ryanodine receptor function implicated as a major contributing factor (Nyegaard et al., 2012). Whole-exome sequencing has revealed an association between three de novo missense CaM mutations and severe long-QT (LQT) syndrome with recurrent cardiac arrest (Crotti et al., 2013), quite plausibly acting via disruption of CaV1.2/1.3 channel CDI (Limpitikul et al., 2014).In brain, state-of-the-art analysis of genome-wide single-nucleotide polymorphisms (SNPs) has identified CaV channels as a major risk factor for several psychiatric disorders (Cross-Disorder Group of the Psychiatric Genomics Consortium, 2013). More specifically to calmodulation, CaV1.3 channels constitute a prominent Ca2+ entry portal into pacemaking and oscillatory neurons (Bean, 2007), owing to the more negative voltages required to open these ion channels. These channels are subject to extensive alternative splicing (Hui et al., 1991; Xu and Lipscombe, 2001; Shen et al., 2006; Bock et al., 2011; Tan et al., 2011) and RNA editing (Huang et al., 2012) in the carboxy tail, in ways that strongly modulate the strength of CDI (Shen et al., 2006; Liu et al., 2010; Huang et al., 2012). This fine tuning of CDI appears to be important for circadian rhythms (Huang et al., 2012). From the specific perspective of disease, CaV1.3 channels contribute a substantial portion of Ca2+ entry into substantia nigral neurons (Bean, 2007; Chan et al., 2007; Puopolo et al., 2007; Guzman et al., 2009), which exhibit high-frequency pacemaking (Chan et al., 2007) that drives dopamine release important for movement control. Notably, loss of these neurons is intimately related to Parkinson’s disease, and Ca2+ disturbances and overload are critical to this neurodegeneration (Bezprozvanny, 2009; Surmeier and Sulzer, 2013). Accordingly, an attractive therapeutic possibility for Parkinson’s disease involves discovery of small molecules that selectively down-regulate the opening of CaV1.3 versus other closely related Ca2+ channels (Kang et al., 2012). Thus, understanding the mechanisms underlying the modulation of CaV1.3 CDI is crucial, particularly to furnish specific molecular interfaces as targets of rational screens for small-molecule modulators.The fundamental mechanism of action of a natural modulator of CaV1.3 CDI may be especially relevant. In particular, recent discoveries identify Ca2+-binding proteins (CaBPs), a family of CaM-like brain molecules (Haeseleer et al., 2000) that may also bind CaV channels and other targets (Yang et al., 2002; Kasri et al., 2004). Like CaM, CaBPs are bilobed, with each lobe containing two EF hand Ca2+ binding motifs (Haeseleer et al., 2000), and a recent crystal structure shows overall similarity to CaM (Findeisen and Minor, 2010). However, whereas all four EF hands bind Ca2+ in CaM, one lobe is nonfunctional in CaBPs. Coexpressing CaBPs with CaV1 channels eliminates their CDI, thereby potentially influencing diverse biological processes (Zhou et al., 2004; Yang et al., 2006). Some have argued that CaBPs may simply compete for apoCaM on channels (Lee et al., 2002; Findeisen et al., 2013; Oz et al., 2013). More recent data argue for a different mechanism (Fig. 5 C), in which CaBP and apoCaM can both preassociate with the channel (Yang et al., 2014). In particular, by binding to channel regions that overlap Ca2+/CaM effector loci, CaBPs may preemptively retard the ability of Ca2+/CaM to reach its effector configuration (Fig. 5 B), allowing low concentrations of CaBP molecules in the CNS to still exert functional effects in the presence of much higher CaM levels (Yang et al., 2014). If this scheme is correct, the basic mechanisms of CaBP action may have implications for drug discovery, in that small molecules that target channel interaction surfaces for CaBP and/or Ca2+/CaM may exert potent modulatory actions.In conclusion, this overview of the calmodulation of voltage-gated Ca2+ channels highlights an impressive synergy among elegant molecular regulatory mechanisms, vital biological functions, and pathogenesis and potential therapy. As such, this field promises considerable mystery, enrichment, and enlightenment in the years ahead.  相似文献   

2.
Acidic Ca2+ stores are important sources of Ca2+ during cell signaling but little is known about how Ca2+ enters these stores. In this issue, Melchionda et al. (2016. J. Cell Biol. http://dx.doi.org/10.1083/jcb.201510019) identify a Ca2+/H+ exchanger (CAX) that is required for Ca2+ uptake and cell migration in vertebrates.Intracellular Ca2+ signaling is of fundamental importance in processes such as cell migration but we do not fully understand the contribution made by different intracellular Ca2+ stores to this particular function. Elevation of cytosolic Ca2+ by 10- to 100-fold the normal resting levels can occur by entry of external Ca2+ across the plasma membrane and release of Ca2+ from intracellular organelles such as the ER. Ca2+ ions are transported across membranes by ligand-gated ion channels, energy-dependent pumps, and transporters (Berridge et al., 2003; Lloyd-Evans et al., 2010). Intracellular Ca2+ levels are regulated in this manner from simple organisms, such as yeast, through to complex multicellular organisms, suggesting a degree of conservation across the taxonomic kingdoms (Patel and Cai, 2015). Recent evidence has indicated that “acidic Ca2+ stores” such as lysosomes in mammalian cells are a key intracellular Ca2+ signaling store, like the ER (Lloyd-Evans and Platt, 2011; Patel and Muallem, 2011). The Ca2+ concentration of the lysosome (500 µM) is similar to the ER (Christensen et al., 2002; Lloyd-Evans et al., 2008) but lysosomes are smaller in volume and their impact on cellular Ca2+ signaling seems localized to events that regulate endocytosis, vesicular fusion, and recycling (Ruas et al., 2010; López-Sanjurjo et al., 2013). However, there is a significant amount of evidence emerging that lysosomes are capable of triggering much larger changes in cytosolic Ca2+ during signaling via the induction of Ca2+ release from the ER. This effect appears to be mediated by the most potent intracellular Ca2+–releasing second messenger nicotinic acid adenine dinucleotide phosphate (NAADP), which triggers Ca2+ release from lysosomes via two-pore channels (Brailoiu et al., 2009; Calcraft et al., 2009). In addition to two-pore channels, acidic stores also express other Ca2+-permeable channels (summarized in Fig. 1).Open in a separate windowFigure 1.Lysosomal Ca2+ transporters and channels. Our current understanding of lysosomal Ca2+ transport and the proteins that regulate the transport of Ca2+ into and out of the lysosome is heavily stacked in favor of Ca2+ release channels. To date, voltage-gated (CaV2.1/CACNA1A), ligand-gated (TRPML1 and TRPM2), and nucleotide-gated (TPC1, TPC2, and P2X4) channels have all been identified or implicated in lysosomal Ca2+ release (Patel and Cai, 2015). Much less is known about the mechanisms of Ca2+ entry into lysosomes. In lower order organisms, CAX mediates lysosomal Ca2+ entry against the proton gradient. In this issue, Melchionda et al. (2016) provide the first evidence for a mammalian lysosomal Ca2+ uptake mechanism in nonplacental mammals. These findings provide further support for the key role of the lysosome as an intracellular Ca2+ store.Despite recent advances in our knowledge of lysosomal Ca2+ release channels, we have so far failed to identify the transport proteins that fill the lysosome with Ca2+. Ca2+ entering the cell by endocytosis is removed by the early endosome after the initiation of endosomal acidification by the vATPase; therefore, it is likely that lysosomes have their own transporters or pumps to take up Ca2+ (Gerasimenko et al., 1998; Christensen et al., 2002). Although there have been studies suggesting the presence of ATPases and putative ion exchangers on mammalian cells (Styrt et al., 1988), the identity of the proteins that mediate lysosomal Ca2+ uptake remains elusive. In this issue, Melchionda et al. describe the first lysosomal CAX in nonplacental mammals and link lysosomal Ca2+ import via CAX to the maintenance of normal cellular migration during development.To identify novel regulators of Ca2+ transport in vertebrates, Melchionda et al. (2016) searched gene databases for homologues of the CAX proteins, which are known to use the proton gradient across the vacuole to drive Ca2+ uptake in plant and yeast cells (Dunn et al., 1994). They identified putative CAX genes in many species, from sea urchins and frogs to reptiles and birds. The CAX homologues discovered in the genomes of the platypus and Tasmanian devil are the first lysosomal Ca2+ exchangers to be identified in any mammalian species. This new work is a significant finding as it suggests that these mechanisms do clearly exist in some mammalian cells and are required for lysosomal Ca2+ store filling. To examine the regulation of lysosomal Ca2+ uptake by vertebrate CAX transporters, the authors cloned full-length CAX from the frog and found that expression of frog CAX could rescue Ca2+ transport in yeast lacking their own CAX. Furthermore, the authors show that the frog CAX channels correctly localize to lysosomes when expressed in human cell lines and that these CAX are capable of manipulating lysosomal and cytosolic Ca2+ levels (in a manner perhaps comparable to plasma membrane Ca2+ ATPases). The findings reported in Melchionda et al. (2016) also have significance for researchers who are using simpler model organisms to characterize mechanisms regulating acidic store Ca2+. A study by Churchill et al. (2002) that used acidic stores purified from sea urchin egg homogenate to monitor acidic store Ca2+ entry concluded that vanadate-sensitive Ca2+ pumps were absent and suggested instead the presence of a CAX. This now appears to be the case through the reported cloning of sea urchin CAX. The findings of Melchionda et al. (2016) are a step forward in unraveling the molecular mechanisms of Ca2+ handling in model animals.Ca2+ signaling plays an important role in development, particularly for cellular migration, where localized elevations in intracellular Ca2+ drive rearrangement of the cytoskeleton, cellular contraction, and adhesion (Wei et al., 2009; Sumoza-Toledo et al., 2011; Praitis et al., 2013). A concentration gradient of Ca2+ exists across the migrating cell, with higher levels at the rear that contribute to cellular detachment and contraction (Praitis et al., 2013). Recent evidence has highlighted the presence of Ca2+ flickers at the leading edge of the migrating cell that have been shown to underlie changes in direction (Wei et al., 2009). Despite the clear importance of Ca2+ in mediating cellular migration events and the emergent role of lysosomes in maintaining intracellular Ca2+ signaling, very little is known about the roles of lysosomal Ca2+ stores in cellular migration. ER Ca2+ channels including the inositol 1,4,5-trisphosphate receptors and ryanodine receptors as well as the secretory pathway Ca2+ ATPase and lysosomal TRPM2 have all been implicated in regulating changes in intracellular Ca2+ to mediate cellular migration, but to date no lysosomal transporters have been implicated in this process (Wei et al., 2009; Sumoza-Toledo et al., 2011; Praitis et al., 2013). Melchionda et al. (2016) investigated the migration of neural crest cells during frog development to find out whether or not CAX transporters control cell motility. CAX proteins are expressed in the neural crest of developing frogs and morpholino-mediated knockdown of CAX expression increased cytosolic Ca2+ levels and impeded neural crest cell migration. Confocal imaging of neural crest tissue in vitro revealed the dynamic recruitment of CAX-containing vesicles to the protrusions that contain focal adhesion complexes at the leading edge of the migrating neural cells. Loss of CAX protein expression reduced the ability of neural crest cells to form stable focal adhesions and undergo the initial cell spreading required for migration. The work presented by Melchionda et al. (2016) is a significant discovery providing evidence that lysosomal Ca2+ uptake is involved in cell migration and that lower organisms are useful model systems to investigate the role of acidic store Ca2+ in this critical cellular function during embryo development.Melchionda et al. (2016) have made a significant step forward in our understanding of the mechanisms that regulate lysosomal/vacuolar Ca2+ entry. However, we remain in the dark about the identity of the transporters that pump Ca2+ into the lysosomes of placental mammals. What led to the loss of CAX genes in these organisms is as much a mystery as the identity of the transporters that have replaced CAX. Evidence from a study using purified mammalian lysosomes to observe Ca2+ uptake indicates that the process is ATP-dependent (Styrt et al., 1988). Placental mammals may have completely different ATP-dependent mechanisms governing lysosomal Ca2+ uptake compared with lower order organisms and nonplacental mammals. Interestingly, defects in lysosomal Ca2+ uptake are associated with two human diseases, Niemann-Pick type C and Chediak-Higashi syndrome (CHS; Styrt et al., 1988; Lloyd-Evans et al., 2008). The lysosomal accumulation of sphingosine, a Ca2+ ATPase inhibitor (Lloyd-Evans and Platt, 2011), leads to reduced lysosomal Ca2+ levels in Niemann-Pick type C disease cells and defects in NAADP-mediated lysosomal Ca2+ release (Lloyd-Evans et al., 2008). In CHS, there have been reports of enhanced lysosomal Ca2+ ATPase transporter activity in neutrophils (Styrt et al., 1988). Interestingly, CHS leukocytes show alterations in chemotaxis with a reduced response to chemotactic factors (Clark and Kimball, 1971), which is supportive of the findings of Melchionda et al. (2016). Much remains to be elucidated about the enigma of mammalian lysosomal Ca2+ uptake, but the work of Melchionda et al. (2016) begins to pick this mystery apart.  相似文献   

3.
4.
Calcium signaling and reactive oxygen species signaling are directly connected, and both contribute to cell-to-cell signal propagation in plants.Calcium (Ca2+) is an important second messenger with diverse functions not only in mammals but also in plants. It is released in response to a variety of stimuli like biotic and abiotic stresses and facilitates a tight regulation of response reactions as well as of developmental processes (Sanders et al., 2002; Steinhorst and Kudla, 2012). Ca2+ accumulation events are characterized by distinct temporal and spatial features, and they can vary in terms of amplitude, frequency, and duration (Webb et al., 1996; Scrase-Field and Knight, 2003; Dodd et al., 2010; Kudla et al., 2010). Spatially defined Ca2+ signals can be generated due to the especially slow diffusion rate of the Ca2+ ion in the cytoplasm in combination with tightly regulated release and uptake from and into different intracellular stores and the apoplast. Together, these characteristics encode information about particular stimuli, for example, drought stress that is presented to the cell as so-called Ca2+ signatures (Webb et al., 1996). This information has to be decoded and transmitted by a signaling machinery in order to initiate adequate response reactions, for example, stomatal closure (Allen et al., 2000, 2001; Sanders et al., 2002). Ca2+ signatures can be sensed by proteins that bind Ca2+ via helix-loop-helix EF-hand motifs. Arabidopsis (Arabidopsis thaliana) possesses at least 250 putative EF-hand proteins, 100 of which have been classified as Ca2+ sensor proteins (Day et al., 2002; Hashimoto and Kudla, 2011). Given that each member of this intricate set of Ca2+ sensor proteins can exhibit characteristic expression and subcellular localization profiles as well as distinct Ca2+ affinities, plants are equipped with a complex signal-decoding machinery to process a wide range of different Ca2+ signals (Batistič and Kudla, 2004; Batistič and Kudla, 2010). Ca2+ functions in concert with other important second messengers like reactive oxygen species (ROS). ROS can be generated in a controlled manner by several types of enzymes, such as NADPH oxidases, in order to contribute to pathogen defense and cell signaling. Recent findings point to direct connections between ROS and Ca2+ signaling pathways that enable cell-to-cell communication and thereby long-distance transmission of signals in plants. In this Update, we focus on new findings in the field of plant Ca2+ signaling during the past 3 years since the status of the field was discussed in comprehensive reviews (Dodd et al., 2010; Kudla et al., 2010; Mazars et al., 2011; Reddy et al., 2011) and put special emphasis on the contribution of a plant-specific Ca2+ signaling network to deciphering defined Ca2+ signals and its integration with ROS signaling.  相似文献   

5.
Regions of close apposition between two organelles, often referred to as membrane contact sites (MCSs), mostly form between the endoplasmic reticulum and a second organelle, although contacts between mitochondria and other organelles have also begun to be characterized. Although these contact sites have been noted since cells first began to be visualized with electron microscopy, the functions of most of these domains long remained unclear. The last few years have witnessed a dramatic increase in our understanding of MCSs, revealing the critical roles they play in intracellular signaling, metabolism, the trafficking of metabolites, and organelle inheritance, division, and transport.

Introduction

The compartmentalization of cells allows the segregation and regulation of the myriad reactions that occur within them. The tremendous benefits of intracellular compartmentalization also come at a price; to function optimally, cells must transmit signals and exchange material between compartments. Numerous mechanisms have evolved to facilitate these exchanges. One that has not been well appreciated until the last few years is the transmission of signals and molecules between organelles that occurs at regions where the organelles are closely apposed, often called membrane contact sites (MCSs). These sites were first characterized because of their critical roles in the intracellular exchange of lipids and calcium, which can be directly channeled between organelles via MCSs. More recently, it has also become apparent that MCSs are important sites for intracellular signaling, organelle trafficking, and inheritance, and that MCSs are specialized regions where regulatory complexes are assembled (English and Voeltz, 2013; Helle et al., 2013).A hallmark of MCSs is that membranes from two organelles (or compartments of the same organelle) are tethered to one another, but not all instances in which membranes interact with or are tethered to one another are considered MCSs. True MCSs have four properties: (1) membranes from two intracellular compartments are tethered in close apposition, typically within 30 nm, (2) the membranes do not fuse (though they may transiently hemi-fuse), (3) specific proteins and/or lipids are enriched at the MCS, and (4) MCS formation affects the function or composition of at least one of the two organelles in the MCS.This review will discuss what we know about proteins that tether organelles, the exchange of small molecules at MCSs, and other emerging functions of MCSs.

MCS tethers

An MCS tether is a protein or complex of proteins (Fig. 1) that simultaneously binds the two apposing membranes at an organelle contact site and plays a role in maintaining the site (English and Voeltz, 2013; Helle et al., 2013). In many cases it is not yet clear if these proteins and complexes are genuine tethers, which are necessary to maintain MCSs, or function at MCSs but are not necessary to sustain contacts. Distinguishing between these possibilities is an important challenge for the field, especially when more than one protein or complex of proteins independently hold together the membranes at an MCS.Open in a separate windowFigure 1.Proteins proposed to mediate tethering at MCSs. Mammalian proteins are shown on a yellow background, yeast proteins on a blue background, and proteins found in both mammals and yeast are on a green background. Tethering complexes not described in the text are indicated with red numbers: (1) StARD3-VAPs (Alpy et al., 2013), (2) NPC1-ORP5 (Du et al., 2011), (3) Psd2-Pdr17 (Riekhof et al., 2014), (4) Vac8-Nvj1 (Pan et al., 2000), (5) Nvj2 (Toulmay and Prinz, 2012), (6) PTPIP51-VAPs (De Vos et al., 2012), (7) Orai1-STIM1 (Nunes et al., 2012), (8) DGAT2-FATP1 (Xu et al., 2012), and (9) IncD-CERT-VAPs (Derré et al., 2011; Elwell et al., 2011).As a growing number of potential tethers are identified, three trends are emerging. First, most MCSs are maintained by several tethers. One of the best-characterized examples of this is the junction of the ER and plasma membrane (PM) in Saccharomyces cerevisiae. Recent work showed that it was necessary to eliminate six ER resident proteins to dramatically reduce the normally extensive interactions between the ER and PM (Manford et al., 2012; Stefan et al., 2013). This suggests that these six proteins mediate tethering independently of each other. Four of the six proteins (three calcium and lipid-binding domain proteins 1–3, also called Tcb1–3, and Ist2) are integral ER membrane proteins that have cytosolic domains that bind the plasma membranes (Fischer et al., 2009; Toulmay and Prinz, 2012). The other two proteins, Scs2 and Scs22 (Scs, suppressor of Ca2+ sensitivity), are homologues of mammalian VAPs (vesicle-associated membrane protein–associated proteins). VAPs are integral membrane tail-anchored proteins in the ER that bind proteins containing FFAT (phenylalanines in an acid tract) motifs (Loewen et al., 2003). A number of proteins that contain these motifs also have domains that bind lipids and proteins in the PM, allowing them to simultaneously bind and tether the ER and PM. For example, some oxysterol-binding protein (OSBP)–related proteins (ORPs) have FFAT motifs and pleckstrin homology (PH) domains that bind phosphoinositides (PIPs) in the plasma membrane (Levine and Munro, 1998; Weber-Boyvat et al., 2013). Thus, ORPs and other FFAT motif-containing proteins can mediate ER–PM tethering via VAPs. It should be noted that VAPs and proteins bound by VAPs also mediate tethering between the ER and organelles in addition to the PM. These are shown in Fig. 1.A second emerging trend is that tethering seems to be a dynamic, regulated process, and we are beginning to understand the mechanisms of dynamic apposition of membranes at MCSs by tethers. One example is ER–PM tethering mediated by proteins called extended synaptotagmins (E-Syts), which are homologues of the yeast Tcb tethers. The tethering of the ER and PM by E-Syts is regulated by Ca2+ and the PM-enriched lipid PI(4,5)P2 (Chang et al., 2013; Giordano et al., 2013). Binding of these molecules by E-Syts may control both the extent of ER–PM contact and the distance between these organelles at MCSs. A second example of regulated MCS formation is provided by a recent study on OSBP. This protein and other FFAT motif-containing proteins have been thought to mediate ER–Golgi tethering by simultaneously binding VAPs in the ER and PIPs in the Golgi complex (Kawano et al., 2006; Peretti et al., 2008). In an elegant set of experiments, Mesmin et al. (2013) showed that OSBP regulates its own ability to mediate ER–Golgi tethering by modulating PI4P levels in the Golgi complex. When PI4P levels in the Golgi complex are high, OSBP tethers the ER and Golgi complex and also transports PI4P from the Golgi to the ER. When the PI4P reaches the ER, it is hydrolyzed by the phosphatase Sac1, preventing it from being transferred back to the Golgi. The reduction in Golgi complex PI4P levels by OSBP causes OSBP to dissociate from the Golgi, decreasing ER–Golgi tethering. Thus, OSBP negatively regulates its own tethering of the ER and Golgi membranes. Lipid transport by OSBP and similar proteins will be discussed in more detail in the section on lipid transport at MCSs.The third important feature of many MCS tethering complexes is that most have functions in addition to tethering. This is well illustrated by complexes proposed to mediate ER–mitochondria tethering in mammalian cells, where four such complexes have been described (Fig. 1). For example, Mfn2 (mitofusin-2) acts as a tether (de Brito and Scorrano, 2008), but the primary function of this dynamin-like protein is to mediate mitochondrial fusion. Although Mfn2 is largely in the outer mitochondrial membrane (OMM), a small fraction also resides the ER, and it has been proposed that the interaction of Mfn2 in the ER with Mfn2 in the OMM tethers the ER and mitochondria (de Brito and Scorrano, 2008). The other ER–mitochondria tethering complexes proposed in mammals (Fig. 1) also have additional functions—either Ca2+ signaling or apoptotic signaling between these organelles.

Tethers within organelles

MCSs may form not only between organelles but also between compartments of the same organelle. In two cases, proteins necessary for these intra-organelle contacts are known. The Golgi complex is divided into a number of cisternae that remain closely apposed in some cell types, forming stacked compartments. Two tethering proteins maintain connections between Golgi cisternae. Golgi reassembly stacking protein 65 (GRASP65) forms contacts between cis- and medial-Golgi cisternae and GRASP55 mediates medial- to trans-cisternal interactions (Fig. 1; Barr et al., 1997; Shorter et al., 1999). The Golgi stack disassembles when both GRASPs are depleted, indicating that they are the primary or sole tethers (Xiang and Wang, 2010). Tethering by these proteins is regulated by kinases to allow Golgi cisternal disassembly during the cell cycle. Whether the inter-Golgi contacts formed by GRASPs mediate signaling or lipid exchange between cisternae is not yet known (Tang and Wang, 2013).MCSs also form inside organelles with internal membranes: mitochondria, chloroplasts, and multivesicular bodies. These MCSs may form between membranes within these organelles or between internal membranes and the outer membrane of the organelle. Recently, three groups discovered a tethering complex involved in forming contacts between mitochondrial cisternae and between cisternae and the mitochondrial outer membrane (Harner et al., 2011; Hoppins et al., 2011; von der Malsburg et al., 2011). This complex, called the mitochondrial contact site and cristae organizing system (MICOS), is conserved from yeast to humans and contains at least six proteins (Fig. 1). It is necessary to maintain inner membrane organization and also interacts with protein complexes in the outer membrane, including the translocase of the outer membrane (TOM) complex and the sorting and assembly machinery (SAM) complex (van der Laan et al., 2012; Zerbes et al., 2012).

Lipid exchange at MCSs

Lipid exchange between organelles at MCSs may serve a number of important functions. One is that it allows cells to rapidly modulate the lipid composition of an organelle independently of vesicular trafficking. In addition, some organelles, such as mitochondria and chloroplasts, must obtain most of the lipids they require for membrane biogenesis by nonvesicular lipid trafficking that almost certainly occurs at MCSs (Osman et al., 2011; Wang and Benning, 2012; Horvath and Daum, 2013). Finally, and perhaps most importantly, lipid transfer at MCSs may play an important role in lipid metabolism by channeling lipids to or away from enzymes in different compartments.Some lipid exchange at MCSs is facilitated by soluble lipid transport proteins (LTPs), which can shuttle lipid monomers between membranes (Fig. 2 A). In other cases, known LTPs do not seem to be required and lipids may be exchanged at MCSs by other mechanisms (Fig. 2, B and C), which will be discussed next.Open in a separate windowFigure 2.Possible mechanisms of lipid exchange at MCSs. (A) Transfer by LTPs using CERT as an example. The targeting PH domain (pink) and FFAT motif (blue) are shown. CERT could shuttle between membranes (left) or transfer while binding both membranes (right). (B) Some transfer could occur through hydrophobic channels or tunnels (in green) bridging the two membranes at a MCS. (C) Lipid exchange between hemifused membranes. Hemifusion could be promoted and regulated by proteins (red).Most LTPs fall into at least five superfamilies that differ structurally but that all have a hydrophobic pocket or groove that can bind a lipid monomer, and often have a lid domain that shields the bound lipid from the aqueous phase (D’Angelo et al., 2008; Lev, 2010). This allows LTPs to shuttle lipid monomers between membranes. LTPs probably transfer lipids between organelles in cells most efficiently at MCSs, where they have only a short distance to diffuse between membranes. LTPs that may transfer lipids at contact sites are: OSBP, ceramide transport protein (CERT), the yeast OSBP homologues Osh6 and Osh7, protein tyrosine kinase 2 N-terminal domain–interacting receptor 2 (Nir2), and Ups1 (Hanada, 2010; Connerth et al., 2012; Chang et al., 2013; Maeda et al., 2013; Mesmin et al., 2013).LTPs could function by shuttling between membranes at MCSs or while simultaneously bound to both membranes (Fig. 2 A). Many LTPs have domains that target them to the two membranes at an MCS. For example, OSBP and CERT have FFAT motifs, which bind ER resident VAPs, and PH domains that bind PIPs in the Golgi complex or PM.Another important emerging aspect of lipid exchange by some LTPs is that it may be driven by their ability to exchange one lipid for another. For example, OSBP can transfer both cholesterol and PI4P. At ER–Golgi MCSs, OSBP may facilitate the net movement of cholesterol from the ER to the Golgi and PI4P in the opposite direction (Mesmin et al., 2013). The difference in the PI4P concentrations in the ER and Golgi (lower in the ER than in the Golgi) may drive the net transfer of cholesterol to the Golgi. The ability to exchange one lipid for another has been found for other LTPs (Schaaf et al., 2008; de Saint-Jean et al., 2011; Kono et al., 2013) and may be critical for driving directional lipid exchange at MCSs.Some lipid exchange at MCSs does not seem to be facilitated by LTPs. The best evidence for this comes from studies on lipid transfer between the ER and mitochondria. It has long been known that lipids are exchanged between these two organelles; mitochondria must acquire most of the lipid it requires for membrane biogenesis from the rest of the cell. Lipid exchange at ER–mitochondria MCSs occurs by a mechanism that does not require energy, at least in vitro, and does not require any cytosolic factors (Osman et al., 2011; Vance, 2014).How this lipid transfer occurs is not known, and two possible types of mechanism are shown in Fig. 2, B and C. One is that some MCS proteins form a hydrophobic channel that allows lipids to move between membranes. Such a channel would be similar to an LTP, but whereas lipids enter and exit LTPs by the same opening, they enter and exit channels by different openings. This difference could allow lipid exchange by a channel to be regulated and, if the channel could bind two different lipids simultaneously, it might couple the transfer of the lipids. A domain that may form channels at MCSs has been identified. Called the synaptotagmin-like mitochondrial lipid-binding protein (SMP) domain, it has been predicted to be part of a superfamily of proteins that includes cholesterol ester transfer protein (CETP; Kopec et al., 2010). CETP has a tubular lipid-binding domain that transfers lipids between high-density and low-density lipoproteins, probably while simultaneously bound to both (Qiu et al., 2007; Zhang et al., 2012). SMP domains could transfer lipids between membranes by a similar mechanism. Consistent with this possibility, all SMP-containing proteins in budding yeast localize to MCSs and many mammalian SMP-containing proteins do as well (Toulmay and Prinz, 2012). Interestingly, SMP domains are present in three of the five proteins in a yeast ER–mitochondria tethering complex called ERMES (Kornmann et al., 2009). Whether ERMES facilitates lipid exchange between the ER and mitochondria is not yet clear. Mitochondria derived from cells missing ERMES have altered lipid composition (Osman et al., 2009; Tamura et al., 2012; Tan et al., 2013), indicating that lipid exchange between the ER and mitochondria could be altered in these strains. On the other hand, little or no defect in the rates of phospholipid exchange between ER and mitochondria were found in ERMES mutants (Kornmann et al., 2009; Nguyen et al., 2012; Voss et al., 2012). Thus, whether proteins that contain SMP domains actually facilitate lipid exchange remains to be determined.As second possible mechanism of lipid transfer at MCSs that does not require LTPs is membrane hemifusion (Fig. 2 C), which could allow rapid exchange of large amounts of lipids between compartments. Recent indirect evidence suggests that hemifusion may occur between the ER and chloroplasts (Mehrshahi et al., 2013). This is consistent with an earlier study using optical tweezers that found the ER and chloroplasts remained attached to one another even when a stretching force of 400 pN was applied (Andersson et al., 2007). Whether hemifusion occurs at MCSs in animal cells remains to be determined.

Calcium signaling at MCSs

MCSs between the ER and PM and the ER and mitochondria play central roles in intracellular Ca2+ storage, homeostasis, and signaling in mammalian cells. MCSs between the ER and lysosomes may also be important, though they are less well understood (Helle et al., 2013; Lam and Galione, 2013).One of the best-characterized MCSs is the one formed between the PM and ER in muscle cells. In both cardiac and skeletal muscle cells, deep invaginations of the PM, called T (transverse)-tubules, allow it to form extensive contacts with the ER, called the sarcoplasmic reticulum (SR) in muscle cells. These contacts are essential for coupling excitation and contraction. Before excitation, Ca2+ levels in the cytoplasm of muscle cells are low, whereas the Ca2+ concentrations in the SR and outside muscle cells are high. During muscle excitation, Ca2+ rapidly flows into the cytosol through channels in the PM and the SR (Fig. 3 A). The channels in the PM, called dihydropyridine receptors (DHPRs), and those in the SR, known ryanodine receptors RyRs, directly interact with each other where the SR and PM are closely apposed, allowing the opening of both types of channels to be coordinated (Fabiato, 1983; Bannister, 2007; Beam and Bannister, 2010; Rebbeck et al., 2011).Open in a separate windowFigure 3.Ca2+ trafficking at ER–PM MCSs. (A) In muscle cells, the interaction of the RyR in the SR and with DHPR in the PM allows the coordinated release of Ca2+ during muscle excitation and contraction. See text for details. (B) When STIM1 senses low Ca2+ concentration in the ER, it undergoes a conformational change that allows it to oligomerize and bind to the PM, to the protein Orai1, and to accumulate at ER–PM MCSs. Ca2+ influx at these sites facilitates Ca2+ import into the ER by sarco/endoplasmic reticulum Ca2+-ATPase (SERCA). (C) Calcium channeling from the ER lumen to the mitochondrial matrix. Calcium exits the ER through the inositol trisphosphate receptor (IP3R) channel, enters mitochondria via VDAC, and then uses the mitochondrial Ca2+ uniporter (MCU) to move into the mitochondrial matrix.The extensive contacts between the SR and PM in muscle cells are largely maintained by tethering proteins called junctophilins, which have a single transmembrane domain in the SR and a large cytosolic domain that interacts with the PM. Expression of junctophilins in cells lacking them induces ER–PM contacts (Takeshima et al., 2000) and cells lacking junctophilins have abnormal SR–PM MCSs and defects in Ca2+ signaling (Ito et al., 2001; Komazaki et al., 2002; Hirata et al., 2006). Thus, junctophilins are both necessary and sufficient for generating functional SR–PM contacts. However, cells lacking junctophilins still maintain some SR–PM contacts, indicating that other proteins also tether the SR and the PM. Some of this residual tethering probably comes from the interaction of DHPRs and RyRs.ER–PM contacts also play a role in regulating intracellular Ca2+ levels in non-excitable cells. When the Ca2+ concentration in the ER lumen is low it triggers Ca2+ entry into the cytosol and ER from outside cells (Fig. 3 B), a process known as store-operated Ca2+ entry (SOCE). The PM channel responsible for Ca2+ entry is Orai1, and the sensor of Ca2+ concentration in the ER lumen is the integral membrane protein stromal interaction molecule-1 (STIM1). When STIM1 senses that the Ca2+ concentration in the ER is low, it oligomerizes and undergoes a conformational change that exposes a basic cluster of amino acids in its C terminus that binds PIPs in the PM (Stathopulos et al., 2006, 2008; Liou et al., 2007; Muik et al., 2011). STIM1 also binds to Orai1 in the PM and activates it (Kawasaki et al., 2009; Muik et al., 2009; Park et al., 2009; Wang et al., 2009). Activation of STIM1 causes it to shift from being relatively evenly distributed on the ER to forming a number of puncta, which are regions were the ER and PM are closely apposed. It seems likely that STIM1 accumulates at and expands preexisting ER–PM MCSs and may also drive the formation of new MCSs (Wu et al., 2006; Lur et al., 2009; Orci et al., 2009).The interaction of STIM1 and Orai1 at ER–PM contacts during SOCE is an elegant mechanism for channeling both signals and small molecules at an MCS. The signal that ER luminal Ca2+ concentration is low is transmitted directly from STIM1 in the ER to Orai1 in the PM. The close contact of PM and ER also allows Ca2+ to move from outside the cell into the lumen of the ER without significantly increasing cytosolic Ca2+ levels (Jousset et al., 2007). During SOCE, ER Ca2+ levels are restored by the sarco/endoplasmic reticulum Ca2+-ATPase (SERCA) pump (Sampieri et al., 2009; Manjarrés et al., 2011). This pump is enriched in ER–PM contacts with STIM1 and may interact directly with it, suggesting how Ca2+ can be effectively channeled from outside cells directly into the ER lumen at ER–PM MCSs (Fig. 3 B).Interestingly, it has become clear that proteins that are not part of the SOCE pathway also facilitate ER–PM connections during Ca2+ signaling. The E-Syts have multiple domains that probably bind Ca2+. They have been shown to regulate both the number of the ER–PM contacts and the distance between the ER and PM at MCSs during Ca2+ signaling (Chang et al., 2013; Giordano et al., 2013).MCSs between the ER and mitochondria similarly facilitate Ca2+ movement from the ER lumen to mitochondria (Rizzuto et al., 1998; Csordás et al., 2006). Ca2+ channels in the ER and OMM interact with each other at MCSs (Fig. 3 C). The channel in the ER is called the inositol trisphosphate receptor (IP3R), while the voltage-dependent anion channel (VDAC) in the outer mitochondrial membrane is a nonspecific pore that allows Ca2+ entry into mitochondria. These proteins, together with the cytosolic chaperone Grp75, form a complex that links the ER and mitochondria and facilitates Ca2+ exchange (Szabadkai et al., 2006).More evidence that Ca2+ transfer from the ER to mitochondria occurs at MCSs came from studies on the channel that allows Ca2+ to move across the inner mitochondrial membrane, called the mitochondrial Ca2+ uniporter (MCU). Surprisingly, this channel has an affinity for Ca2+ that is lower than the typical Ca2+ concentration in the cytosol (Kirichok et al., 2004). However, Ca2+ release by the ER at ER–mitochondrial MCSs suggests a solution to this puzzle; the local Ca2+ concentration at these MCSs is probably high enough for MCU to function (Csordás et al., 2010). Close contacts between the ER and mitochondria are therefore essential for channeling Ca2+ from the ER lumen to the mitochondrial matrix.It is thought that MCSs between the ER (or SR) and lysosomes regulate Ca2+ release by lysosomes, but the mechanism is not yet understood (Kinnear et al., 2004, 2008; Galione et al., 2011; Morgan et al., 2011).

Enzymes working in trans and signaling at MCSs

MCSs allow rapid and efficient signaling between intracellular compartments. We are still just beginning to understand the mechanisms and functions of this signaling. One way that signals are transmitted between the two compartments at an MCS is for an enzyme in one compartment to modify substrates in the second; that is, for the enzyme to work in trans. Although there are currently only a few examples of this, which are discussed here, it seems likely that many more will be uncovered.The protein tyrosine phosphatase PTP1B regulates a number of receptor tyrosine kinases. PTP1B resides on the surface of the ER with its active site in the cytosol, and yet the receptor tyrosine kinases it modifies are in the PM. Although this was initially puzzling, it was found that PTP1B probably encounters its substrates at MCSs, either at ER–PM junctions or at contacts between the ER and endocytic recycling compartments (Haj et al., 2002; Boute et al., 2003; Anderie et al., 2007; Eden et al., 2010; Nievergall et al., 2010). Interestingly, in some cases the interaction of PTP1B with substrates in the PM occurs on portions of the PM that are part of cell–cell contacts (Haj et al., 2012), suggesting that ER–PM contacts could play a role in signaling, not only between the ER and PM but between cells as well. Dephosphorylation of receptor tyrosine kinases by PTP1B at contact sites probably allows their kinase activity to be regulated in response to changes in the ER or changes in cellular architecture that alter MCSs. For example, the dephosphorylation of epidermal growth factor receptor (EGFR) by PTP1B occurs at regions of close contact between the ER and multivesicular bodies, causing EGFR to become sequestered with multivesicular bodies (Eden et al., 2010). This may provide a mechanism for cells to regulate EGFR levels on the PM in response to signals in the ER.Lipid metabolism enzymes can also work in trans at MCSs. In two cases, both in yeast, enzymes that reside in the ER have been found to modify lipids in the PM at MCSs. In one instance, the phosphatase Sac1, which is on the surface of the ER, can dephosphorylate PIPs in the PM (Stefan et al., 2011). In the second, the ER enzyme Opi3 methylates phosphatidylethanolamine in the PM, a reaction that is required for the conversion of phosphatidylethanolamine to phosphatidylcholine (Tavassoli et al., 2013). Remarkably, the PIP-binding protein Osh3 (Tong et al., 2013) regulates both reactions, suggesting that lipid metabolism at ER–PM junctions is regulated by PIPs. It seems likely that ER–PM junctions play important roles in integrating lipid metabolism in both organelles.

MCSs and organelle trafficking and inheritance

In addition to being sites at which signals and small molecules are exchanged between cellular compartments, there is growing evidence that MCS formation also regulates organelle trafficking and inheritance.In budding yeast, organelle transport is polarized from the mother cell to the growing bud and is required for proper organelle inheritance. The transport of peroxisomes and mitochondria to the bud is regulated by their association with the ER or PM.Knoblach et al. (2013) found that tethering of the ER to peroxisomes requires Pex3, an integral membrane protein that resides in both compartments, and Inp1, a cytosolic protein that binds to Pex3. This tether keeps peroxisomes in mother cells. When peroxisomes divide they are transferred to the bud by the myosin V motor Myo2 and become attached to the ER in the bud. In cells lacking the ER–peroxisome tether, peroxisomes accumulate in daughter cells. Thus, tethering plays a critical role in ensuring that some peroxisomes are retained in mother cells and that both cells inherit peroxisomes.Mitochondrial inheritance in yeast is regulated by close contacts with both the ER and PM. Mitochondria–PM contacts mediated by a complex containing Num1 and Mdm36 ensure that mitochondria are properly distributed between mother and daughter cells and seem to be particularly important for retaining mitochondria in the mother cells (Klecker et al., 2013; Lackner et al., 2013). Interestingly, Num1–Mdm36-mediated contacts also associate with the ER (Lackner et al., 2013), suggesting that three membranes may somehow associate at these MCSs. An ER–mitochondria tether containing the protein Mmr1, which anchors mitochondria to bud tips, also plays a role in mitochondrial inheritance (Swayne et al., 2011). Thus, the Num1-tethering complex and Mmr1-tethering complex seem to play antagonistic roles in mitochondrial distribution; the Num1 complex promotes mitochondrial retention in the mother, whereas the Mmr1 complex favors retention in the bud.MCSs also play a role in endosomal trafficking in mammalian cells. One of the complexes that tethers the ER to endosomes contains VAPs and ORP1L, which is an OSBP homologue that can bind cholesterol (Fig. 1). ORPlL can also binds the p150Glued subunit of the dynein–dynactin motor that participates in endosome transport along microtubules (Johansson et al., 2007). When cellular cholesterol levels are high, ORP1L associates with p150Glued but not VAPs and endosomes are transported on microtubules. However, when cholesterol levels decrease, ORP1L undergoes a conformation change that dissociates it from p150Glued and allows it to bind to VAPs on the surface of the ER, thus forming a tether between endosomes and the ER (Rocha et al., 2009). Under these conditions, endosome transport on microtubules is blocked. ORPlL is therefore a cholesterol sensor that regulates a switch between the association of endosomes with either motors or the ER.

MCSs and organelle division

A groundbreaking study revealed a new and unexpected role for MCSs between the ER and mitochondria: the ER regulates mitochondrial fission (Friedman et al., 2011). Although a mechanistic understanding of how ER participates in mitochondrial fission is not yet available, the sequence of events is beginning to come into focus (Fig. 4). The ER encircles mitochondria at sites where scission will occur. The ERMES complex is present at these sites (Murley et al., 2013). Because mammalian cells lack ERMES, another tethering complex must perform the same function in higher eukaryotes. Mitochondrial division requires membrane scission by the dynamin-like protein Dnm1/Drp1, which multimerizes on the outer mitochondria membrane. Close contacts between the ER and mitochondria occur before Dnm1/Drp1 assembly, suggesting that these contacts promote or regulate the association of Dnm1/Drp1 with mitochondria and hence mitochondrial division. It is possible that when the ER encircles mitochondria it causes mitochondria to constrict to a diameter that allows Dnm1/Drp1 to assemble. The force necessary to drive constriction may come from actin polymerization. A recent study found that the ER protein, inverted formin-2, probably drives actin polymerization at these sites and is necessary for mitochondria fusion (Korobova et al., 2013).Open in a separate windowFigure 4.Model of ER-mediated regulation of mitochondrial fission at sites of contact. (A) The ER and mitochondria are tethered by ERMES in yeast (other tethers are used in higher eukaryotes). (B) The ER encircles mitochondria at sites where division will occur. (C) Actin polymerization facilitated by formin 2 may cause mitochondrial constriction. (D) The dynamin-like protein Drp1 is recruited to the mitochondrial surface, where it multimerizes and causes mitochondrial scission. (E) After fission, the ER remains associated with the mitochondrion that retains the ERMES complex.Understanding the assembly and regulation of the mitochondrial division machinery at ER–mitochondria MCSs and how this is linked to mitochondrial and perhaps ER function remain fascinating questions for the future. Another interesting question is whether other MCSs play roles in the fission of other organelles.

Proposed functions of ER–mitochondrial MCSs

A growing number of studies have suggested that ER–mitochondria MCSs play critical roles in autophagy, apoptosis, inflammation, reactive oxygen species signaling, and metabolic signaling. ER–mitochondria MCSs have also been implicated in Alzheimer’s disease, Parkinson’s disease, and some viral infections. These topics have been recently reviewed (Eisner et al., 2013; Raturi and Simmen, 2013; Marchi et al., 2014; Vance, 2014) and will not be discussed in detail here.One issue with most of the studies on the functions of ER–mitochondria junctions is that they rely, at least in part, on density gradient purification of the ER that associates with mitochondria. These operationally defined membranes, often called mitochondrial-associated membranes (MAMs), remain poorly defined. In fact, a significant number of proteins that are enriched in MAMs do not seem to be enriched at ER–mitochondria junctions when their localization is determined by other methods (Helle et al., 2013; Vance, 2014). Therefore, it remains unclear why some proteins and lipids are enriched in MAMs.Here, two interesting findings will be discussed that suggest the importance of ER–mitochondrial junctions in signaling in addition to their well-known role in Ca2+ signalling.The induction of apoptosis requires signal transmission between the ER and mitochondria. Part of this signaling process occurs through an interaction between the ER protein Bap31 and the mitochondrial fission protein Fission 1 homologue (Fis1; Iwasawa et al., 2011). This interaction occurs at ER–mitochondria MCSs and results in the cleavage of Bap31 by caspase-8 to form p20Bap31, which is pro-apoptotic. Both Bap31 and Fis1 are parts of larger complexes that are still being characterized. Interestingly, it has recently been found that a protein called cell death–involved p53 target-1 (CDIP1) binds to Bap31 during ER stress and promotes apoptotic signaling from the ER to mitochondria (Namba et al., 2013), suggesting how ER stress signals are transmitted from the ER to mitochondria through MCSs.Another important connection between ER–mitochondrial MCSs and signaling has to do with the target of rapamycin (TOR) kinase complexes, which are critical regulators of growth and metabolism. The mammalian TOR complex 2 (mTORC2) was found to interact with the IP3R–Grp75–VDAC complex that tethers the ER and mitochondria (Betz et al., 2013). Remarkably, this study presents evidence that mTORC regulates both the formation of ER–mitochondrial MCSs and mitochondrial function, suggesting an interesting new mechanism for how metabolic signaling can impact mitochondrial function via MCSs.

Conclusions and perspectives

The potential of MCSs to facilitate Ca2+ signaling and channel lipids between organelles was recognized some time ago (Levine and Loewen, 2006), but it has only been in the last few years that we have finally begun to have some mechanistic insight into how these processes occur and how MCSs are formed. Many fundamental questions remain to be addressed. How lipid exchange at MCSs that does not require soluble LTPs occurs or whether transient hemifusion of membranes at MCS ever occurs remain open questions. Another is the mechanisms by which Ca2+ regulates MCS formation between the ER and other organelles. One major challenge for the field will be devising better methods to visualize MCSs and identify proteins and lipids enriched at these sites. It is particularly important to better understand what the MAM fraction is and what it means for proteins and lipids to be enriched in this fraction.One of the most exciting developments in the study of MCSs in the last few years has been the discovery of the role of MCSs in organelle trafficking, inheritance, and dynamics. These studies have revealed that MCSs not only play critical roles in signaling and metabolism, but also modulate the intracellular distribution of organelles and organelle architecture. Understanding how MCSs perform these functions will probably shed light on the connection between the still murky relationship between organelle structure and function as well as the role of the ER as a regulator of other organelles. Given the current pace of discovery, it seems likely that in the next few years our knowledge of the functions of MCSs will grow dramatically.  相似文献   

6.
Ca2+-dependent protein kinases (CPKs) form a large family of 34 genes in Arabidopsis (Arabidopsis thaliana). Based on their dependence on Ca2+, CPKs can be sorted into three types: strictly Ca2+-dependent CPKs, Ca2+-stimulated CPKs (with a significant basal activity in the absence of Ca2+), and essentially calcium-insensitive CPKs. Here, we report on the third type of CPK, CPK13, which is expressed in guard cells but whose role is still unknown. We confirm the expression of CPK13 in Arabidopsis guard cells, and we show that its overexpression inhibits light-induced stomatal opening. We combine several approaches to identify a guard cell-expressed target. We provide evidence that CPK13 (1) specifically phosphorylates peptide arrays featuring Arabidopsis K+ Channel KAT2 and KAT1 polypeptides, (2) inhibits KAT2 and/or KAT1 when expressed in Xenopus laevis oocytes, and (3) closely interacts in plant cells with KAT2 channels (Förster resonance energy transfer-fluorescence lifetime imaging microscopy). We propose that CPK13 reduces stomatal aperture through its inhibition of the guard cell-expressed KAT2 and KAT1 channels.Stomata are microscopic organs at the leaf surface, each made of two so-called guard cells forming a pore. Opening or closing these pores is the way through which plants control their gas exchanges with the atmosphere (i.e. carbon dioxide uptake to feed the photosynthetic process and transpirational loss of water vapor). Stomatal movements result from osmotically driven fluxes of water, which follow massive exchanges of solutes, including K+ ions, between the guard cells and the surrounding tissues (Hetherington, 2001; Nilson and Assmann, 2007).Both Ca2+-dependent and Ca2+-independent signaling pathways are known to control stomatal movements (MacRobbie, 1993, 1998; Blatt, 2000; Webb et al., 2001; Mustilli et al., 2002; Israelsson et al., 2006; Marten et al., 2007; Laanemets et al., 2013). In particular, Ca2+ signals have been reported to promote stomatal closure through the inhibition of inward K+ channels and the activation of anion channels (Blatt, 1991, 1992, 2000; Thiel et al., 1992; Grabov and Blatt, 1999; Schroeder et al., 2001; Hetherington and Brownlee, 2004; Mori et al., 2006; Marten et al., 2007; Geiger et al., 2010; Brandt et al., 2012; Scherzer et al., 2012). However, little is known about the molecular identity of the links between Ca2+ events and Shaker K+ channel activity. Several kinases and phosphatases are believed to be involved in both the Ca2+-dependent and Ca2+-independent signaling pathways. Plants express two large kinase families whose activity is related to Ca2+ signaling. Firstly, CBL-interacting protein kinases (CIPKs; 25 genes in Arabidopsis [Arabidopsis thaliana]) are indirectly controlled by their interaction with a set of calcium sensors, the calcineurin B-like proteins (CBLs; 10 genes in Arabidopsis). This complex forms a fascinating network of potential Ca2+ signaling decoders (Luan, 2009; Weinl and Kudla, 2009), which have been addressed in numerous reports (Xu et al., 2006; Hu et al., 2009; Batistic et al., 2010; Held et al., 2011; Chen et al., 2013). In particular, some CBL-CIPK pairs have been shown to regulate Shaker channels such as Arabidopsis K+ Transporter1 (AKT1; Xu et al., 2006; Lan et al., 2011) or AKT2 (Held et al., 2011). Second, Ca2+-dependent protein kinases (CPKs) form an even larger family (34 genes in Arabidopsis) of proteins combining a kinase domain with the ability to bind Ca2+, thanks to the so-called EF hands (Harmon et al., 2000; Harper et al., 2004). CPKs, which, interestingly, are not found in animal cells, exhibit different calcium dependencies (Boudsocq et al., 2012). With respect to this, three types of CPKs can be considered: strictly Ca2+-dependent CPKs, Ca2+-stimulated CPKs (with a significant basal activity in the absence of Ca2+), and essentially Ca2+-insensitive CPKs (however, structurally close to kinases of groups 1 and 2).Pioneering work by Luan et al. (1993) demonstrated in Vicia faba guard cells that inward K+ channels were regulated by some Ca2+-dependent kinases. Then, such a Ca2+-dependent kinase was purified from guard cell protoplasts of V. faba and shown to actually phosphorylate the in vitro-translated KAT1 protein, a Shaker channel subunit natively expressed in Arabidopsis guard cells (Li et al., 1998). KAT1 regulation by CPK was shown by the inhibition of KAT1 currents after the coexpression of KAT1 and CDPK from soybean (Glycine max) in oocytes (Berkowitz et al., 2000). Since then, several cpk mutant lines of Arabidopsis have been shown to be impaired in stomatal movements, for example cpk10 (Ca2+ insensitive), cpk4/cpk11 (Ca2+ dependent), and cpk3/cpk6/cpk23 (Ca2+ dependent; Mori et al., 2006; Geiger et al., 2010; Munemasa et al., 2011; Hubbard et al., 2012).Of the nine genes encoding voltage-dependent K+ channels (Shaker) in Arabidopsis (Véry and Sentenac, 2002, 2003; Lebaudy et al., 2007; Hedrich, 2012), six are expressed in guard cells and play a role in stomatal movements: the Gated Outwardly-Rectifying K+ (GORK) gene, encoding an outward K+ channel subunit, and the AKT1, AKT2, Arabidopsis K+ Rectifying Channel1 (AtKC1), KAT1, and KAT2 genes, encoding inward K+ channel subunits (Pilot et al., 2001; Szyroki et al., 2001; Hosy et al., 2003; Pandey et al., 2007; Lebaudy et al., 2008a). Shaker channels result from the assembly of four subunits, and it has been shown that inward subunits tend to heterotetramerize, thus potentially widening the functional and regulatory scope of inward K+ conductance in guard cells (Xicluna et al., 2007; Jeanguenin et al., 2008; Lebaudy et al., 2008a, 2010). Inhibition of inward K+ channels has been shown to reduce stomatal opening (Liu et al., 2000; Kwak et al., 2001). This has grounded a strategy for disrupting inward K+ channel conductance in guard cells by expressing a nonfunctional KAT2 subunit (dominant negative mutation) in a kat2 knockout Arabidopsis line. The resulting Arabidopsis lines, named kincless, have no functional inward K+ channels and exhibit delayed stomatal opening (Lebaudy et al., 2008b) with, in the long term, a biomass reduction compared with the Arabidopsis wild-type line.Among the CPKs presumably expressed in Arabidopsis guard cells (Leonhardt et al., 2004), we looked for CPK13, which belongs to the atypical Ca2+-insensitive type of CPKs (Kanchiswamy et al., 2010; Boudsocq et al., 2012; Liese and Romeis, 2013) and whose role remains unknown in stomatal movements. Here, we confirm first that CPK13 kinase activity is independent of Ca2+ and show that CPK13 expression is predominant in Arabidopsis guard cells using CPK13-GUS lines. We then report that overexpression of CPK13 in Arabidopsis induces a dramatic default in stomatal aperture. Based on the previously reported kincless phenotype (Lebaudy et al., 2008b), we propose that CPK13 could reduce the activity of inward K+ channels in guard cells, particularly that of KAT2. We confirm this hypothesis by voltage-clamp experiments and show an inhibition of KAT2 and KAT1 activity by CPK13 (but not that of AKT2). In addition, we present peptide array phosphorylation assays showing that CPK13 targets, with some specificity, several KAT2 and KAT1 polypeptides. Finally, we demonstrate that KAT2 and CPK13 interact in planta using Förster resonance energy transfer (FRET)-fluorescence lifetime imaging microscopy (FLIM).  相似文献   

7.
Ca2+ and nitric oxide (NO) are essential components involved in plant senescence signaling cascades. In other signaling pathways, NO generation can be dependent on cytosolic Ca2+. The Arabidopsis (Arabidopsis thaliana) mutant dnd1 lacks a plasma membrane-localized cation channel (CNGC2). We recently demonstrated that this channel affects plant response to pathogens through a signaling cascade involving Ca2+ modulation of NO generation; the pathogen response phenotype of dnd1 can be complemented by application of a NO donor. At present, the interrelationship between Ca2+ and NO generation in plant cells during leaf senescence remains unclear. Here, we use dnd1 plants to present genetic evidence consistent with the hypothesis that Ca2+ uptake and NO production play pivotal roles in plant leaf senescence. Leaf Ca2+ accumulation is reduced in dnd1 leaves compared to the wild type. Early senescence-associated phenotypes (such as loss of chlorophyll, expression level of senescence-associated genes, H2O2 generation, lipid peroxidation, tissue necrosis, and increased salicylic acid levels) were more prominent in dnd1 leaves compared to the wild type. Application of a Ca2+ channel blocker hastened senescence of detached wild-type leaves maintained in the dark, increasing the rate of chlorophyll loss, expression of a senescence-associated gene, and lipid peroxidation. Pharmacological manipulation of Ca2+ signaling provides evidence consistent with genetic studies of the relationship between Ca2+ signaling and senescence with the dnd1 mutant. Basal levels of NO in dnd1 leaf tissue were lower than that in leaves of wild-type plants. Application of a NO donor effectively rescues many dnd1 senescence-related phenotypes. Our work demonstrates that the CNGC2 channel is involved in Ca2+ uptake during plant development beyond its role in pathogen defense response signaling. Work presented here suggests that this function of CNGC2 may impact downstream basal NO production in addition to its role (also linked to NO signaling) in pathogen defense responses and that this NO generation acts as a negative regulator during plant leaf senescence signaling.Senescence can be considered as the final stage of a plant’s development. During this process, nutrients will be reallocated from older to younger parts of the plant, such as developing leaves and seeds. Leaf senescence has been characterized as a type of programmed cell death (PCD; Gan and Amasino, 1997; Quirino et al., 2000; Lim et al., 2003). During senescence, organelles such as chloroplasts will break down first. Biochemical changes will also occur in the peroxisome during this process. When the chloroplast disassembles, it is easily observed as a loss of chlorophyll. Mitochondria, the source of energy for cells, will be the last cell organelles to undergo changes during the senescence process (Quirino et al., 2000). At the same time, other catabolic events (e.g. protein and lipid breakdown, etc.) are occurring (Quirino et al., 2000). Hormones may also contribute to this process (Gepstein, 2004). From this information we can infer that leaf senescence is regulated by many signals.Darkness treatment can induce senescence in detached leaves (Poovaiah and Leopold, 1973; Chou and Kao, 1992; Weaver and Amasino, 2001; Chrost et al., 2004; Guo and Crawford, 2005; Ülker et al., 2007). Ca2+ can delay the senescence of detached leaves (Poovaiah and Leopold, 1973) and leaf senescence induced by methyl jasmonate (Chou and Kao, 1992); the molecular events that mediate this effect of Ca2+ are not well characterized at present.Nitric oxide (NO) is a critical signaling molecule involved in many plant physiological processes. Recently, published evidence supports NO acting as a negative regulator during leaf senescence (Guo and Crawford, 2005; Mishina et al., 2007). Abolishing NO generation in either loss-of-function mutants (Guo and Crawford, 2005) or transgenic Arabidopsis (Arabidopsis thaliana) plants expressing NO degrading dioxygenase (NOD; Mishina et al., 2007) leads to an early senescence phenotype in these plants compared to the wild type. Corpas et al. (2004) showed that endogenous NO is mainly accumulated in vascular tissues of pea (Pisum sativum) leaves. This accumulation is significantly reduced in senescing leaves (Corpas et al., 2004). Corpas et al. (2004) also provided evidence that NO synthase (NOS)-like activity (i.e. generation of NO from l-Arg) is greatly reduced in senescing leaves. Plant NOS activity is regulated by Ca2+/calmodulin (CaM; Delledonne et al., 1998; Corpas et al., 2004, 2009; del Río et al., 2004; Valderrama et al., 2007; Ma et al., 2008). These studies suggest a link between Ca2+ and NO that could be operating during senescence.In animal cells, all three NOS isoforms require Ca2+/CaM as a cofactor (Nathan and Xie, 1994; Stuehr, 1999; Alderton et al., 2001). Notably, animal NOS contains a CaM binding domain (Stuehr, 1999). It is unclear whether Ca2+/CaM can directly modulate plant NOS or if Ca2+/CaM impacts plant leaf development/senescence through (either direct or indirect) effects on NO generation. However, recent studies from our lab suggest that Ca2+/CaM acts as an activator of NOS activity in plant innate immune response signaling (Ali et al., 2007; Ma et al., 2008).Although Arabidopsis NO ASSOCIATED PROTEIN1 (AtNOA1; formerly named AtNOS1) was thought to encode a NOS enzyme, no NOS-encoding gene has yet been identified in plants (Guo et al., 2003; Crawford et al., 2006; Zemojtel et al., 2006). However, the AtNOA1 loss-of-function mutant does display reduced levels of NO generation, and several groups have used the NO donor sodium nitroprusside (SNP) to reverse some low-NO related phenotypes in Atnoa1 plants (Guo et al., 2003; Bright et al., 2006; Zhao et al., 2007). Importantly, plant endogenous NO deficiency (Guo and Crawford, 2005; Mishina et al., 2007) or abscisic acid/methyl jasmonate (Hung and Kao, 2003, 2004) induced early senescence can be successfully rescued by application of exogenous NO. Addition of NO donor can delay GA-elicited PCD in barley (Hordeum vulgare) aleurone layers as well (Beligni et al., 2002).It has been suggested that salicylic acid (SA), a critical pathogen defense metabolite, can be increased in natural (Morris et al., 2000; Mishina et al., 2007) and transgenic NOD-induced senescent Arabidopsis leaves (Mishina et al., 2007). Pathogenesis related gene1 (PR1) expression is up-regulated in transgenic Arabidopsis expressing NOD (Mishina et al., 2007) and in leaves of an early senescence mutant (Ülker et al., 2007).Plant cyclic nucleotide gated channels (CNGCs) have been proposed as candidates to conduct extracellular Ca2+ into the cytosol (Sunkar et al., 2000; Talke et al., 2003; Lemtiri-Chlieh and Berkowitz, 2004; Ali et al., 2007; Demidchik and Maathuis, 2007; Frietsch et al., 2007; Kaplan et al., 2007; Ma and Berkowitz, 2007; Urquhart et al., 2007; Ma et al., 2009a, 2009b). Arabidopsis “defense, no death” (dnd1) mutant plants have a null mutation in the gene encoding the plasma membrane-localized Ca2+-conducting CNGC2 channel. This mutant also displays no hypersensitive response to infection by some pathogens (Clough et al., 2000; Ali et al., 2007). In addition to involvement in pathogen-mediated Ca2+ signaling, CNGC2 has been suggested to participate in the process of leaf development/senescence (Köhler et al., 2001). dnd1 mutant plants have high levels of SA and expression of PR1 (Yu et al., 1998), and spontaneous necrotic lesions appear conditionally in dnd1 leaves (Clough et al., 2000; Jirage et al., 2001). Endogenous H2O2 levels in dnd1 mutants are increased from wild-type levels (Mateo et al., 2006). Reactive oxygen species molecules, such as H2O2, are critical to the PCD/senescence processes of plants (Navabpour et al., 2003; Overmyer et al., 2003; Hung and Kao, 2004; Guo and Crawford, 2005; Zimmermann et al., 2006). Here, we use the dnd1 mutant to evaluate the relationship between leaf Ca2+ uptake during plant growth and leaf senescence. Our results identify NO, as affected by leaf Ca2+ level, to be an important negative regulator of leaf senescence initiation. Ca2+-mediated NO production during leaf development could control senescence-associated gene (SAG) expression and the production of molecules (such as SA and H2O2) that act as signals during the initiation of leaf senescence programs.  相似文献   

8.
The role of calcium-mediated signaling has been extensively studied in plant responses to abiotic stress signals. Calcineurin B-like proteins (CBLs) and CBL-interacting protein kinases (CIPKs) constitute a complex signaling network acting in diverse plant stress responses. Osmotic stress imposed by soil salinity and drought is a major abiotic stress that impedes plant growth and development and involves calcium-signaling processes. In this study, we report the functional analysis of CIPK21, an Arabidopsis (Arabidopsis thaliana) CBL-interacting protein kinase, ubiquitously expressed in plant tissues and up-regulated under multiple abiotic stress conditions. The growth of a loss-of-function mutant of CIPK21, cipk21, was hypersensitive to high salt and osmotic stress conditions. The calcium sensors CBL2 and CBL3 were found to physically interact with CIPK21 and target this kinase to the tonoplast. Moreover, preferential localization of CIPK21 to the tonoplast was detected under salt stress condition when coexpressed with CBL2 or CBL3. These findings suggest that CIPK21 mediates responses to salt stress condition in Arabidopsis, at least in part, by regulating ion and water homeostasis across the vacuolar membranes.Drought and salinity cause osmotic stress in plants and severely affect crop productivity throughout the world. Plants respond to osmotic stress by changing a number of cellular processes (Xiong et al., 1999; Xiong and Zhu, 2002; Bartels and Sunkar, 2005; Boudsocq and Lauriére, 2005). Some of these changes include activation of stress-responsive genes, regulation of membrane transport at both plasma membrane (PM) and vacuolar membrane (tonoplast) to maintain water and ionic homeostasis, and metabolic changes to produce compatible osmolytes such as Pro (Stewart and Lee, 1974; Krasensky and Jonak, 2012). It has been well established that a specific calcium (Ca2+) signature is generated in response to a particular environmental stimulus (Trewavas and Malhó, 1998; Scrase-Field and Knight, 2003; Luan, 2009; Kudla et al., 2010). The Ca2+ changes are primarily perceived by several Ca2+ sensors such as calmodulin (Reddy, 2001; Luan et al., 2002), Ca2+-dependent protein kinases (Harper and Harmon, 2005), calcineurin B-like proteins (CBLs; Luan et al., 2002; Batistič and Kudla, 2004; Pandey, 2008; Luan, 2009; Sanyal et al., 2015), and other Ca2+-binding proteins (Reddy, 2001; Shao et al., 2008) to initiate various cellular responses.Plant CBL-type Ca2+ sensors interact with and activate CBL-interacting protein kinases (CIPKs) that phosphorylate downstream components to transduce Ca2+ signals (Liu et al., 2000; Luan et al., 2002; Batistič and Kudla, 2004; Luan, 2009). In several plant species, multiple members have been identified in the CBL and CIPK family (Luan et al., 2002; Kolukisaoglu et al., 2004; Pandey, 2008; Batistič and Kudla, 2009; Weinl and Kudla, 2009; Pandey et al., 2014). Involvement of specific CBL-CIPK pair to decode a particular type of signal entails the alternative and selective complex formation leading to stimulus-response coupling (D’Angelo et al., 2006; Batistič et al., 2010).Several CBL and CIPK family members have been implicated in plant responses to drought, salinity, and osmotic stress based on genetic analysis of Arabidopsis (Arabidopsis thaliana) mutants (Zhu, 2002; Cheong et al., 2003, 2007; Kim et al., 2003; Pandey et al., 2004, 2008; D’Angelo et al., 2006; Qin et al., 2008; Tripathi et al., 2009; Held et al., 2011; Tang et al., 2012; Drerup et al., 2013; Eckert et al., 2014). A few CIPKs have also been functionally characterized by gain-of-function approach in crop plants such as rice (Oryza sativa), pea (Pisum sativum), and maize (Zea mays) and were found to be involved in osmotic stress responses (Mahajan et al., 2006; Xiang et al., 2007; Yang et al., 2008; Tripathi et al., 2009; Zhao et al., 2009; Cuéllar et al., 2010).In this report, we examined the role of the Arabidopsis CIPK21 gene in osmotic stress response by reverse genetic analysis. The loss-of-function mutant plants became hypersensitive to salt and mannitol stress conditions, suggesting that CIPK21 is involved in the regulation of osmotic stress response in Arabidopsis. These findings are further supported by an enhanced tonoplast targeting of the cytoplasmic CIPK21 through interaction with the vacuolar Ca2+ sensors CBL2 and CBL3 under salt stress condition.  相似文献   

9.
10.
Extracellular ATP induces a rise in the level of cytosolic free calcium ([Ca2+]cyt) in plant cells. To expand our knowledge about the function of extracellular nucleotides in plants, the effects of several nucleotide analogs and pharmacological agents on [Ca2+]cyt changes were studied using transgenic Arabidopsis (Arabidopsis thaliana) expressing aequorin or the fluorescence resonance energy transfer-based Ca2+ sensor Yellow Cameleon 3.6. Exogenously applied CTP caused elevations in [Ca2+]cyt that displayed distinct time- and dose-dependent kinetics compared with the purine nucleotides ATP and GTP. The inhibitory effects of antagonists of mammalian P2 receptors and calcium influx inhibitors on nucleotide-induced [Ca2+]cyt elevations were distinct between CTP and purine nucleotides. These results suggest that distinct recognition systems may exist for the respective types of nucleotides. Interestingly, a mutant lacking the heterotrimeric G protein Gβ-subunit exhibited a remarkably higher [Ca2+]cyt elevation in response to all tested nucleotides in comparison with the wild type. These data suggest a role for Gβ in negatively regulating extracellular nucleotide signaling and point to an important role for heterotrimeric G proteins in modulating the cellular effects of extracellular nucleotides. The addition of extracellular nucleotides induced multiple temporal [Ca2+]cyt oscillations, which could be localized to specific root cells. The oscillations were attenuated by a vesicle-trafficking inhibitor, indicating that the oscillations likely require ATP release via exocytotic secretion. The results reveal new molecular details concerning extracellular nucleotide signaling in plants and the importance of fine control of extracellular nucleotide levels to mediate specific plant cell responses.The calcium ion, Ca2+, is a ubiquitous second messenger that is used to regulate a wide range of cellular processes (Clapham, 2007). A number of plant environmental and developmental responses are encoded to distinct Ca2+ signal patterns with specific frequencies and amplitudes of cytosolic free Ca2+ concentration ([Ca2+]cyt). These signal patterns can take the form of pulsating [Ca2+]cyt spiking/oscillations (Berridge et al., 2003). In plants, such [Ca2+]cyt oscillations occur in various cell types (e.g. stomatal guard cells, pollen tubes, and legume root hairs) and play a critical role in responding to environmental signals (Evans et al., 2001; Oldroyd and Downie, 2008; McAinsh and Pittman, 2009).ATP is a ubiquitous compound in all living cells; it not only provides the energy to drive many biochemical reactions but also functions in signal transduction as a substrate for kinases, adenylate cyclases, etc. However, ATP was also shown to be an essential signaling agent outside of cells in animals, where it is referred to as extracellular ATP. Extracellular ATP is involved in numerous cellular processes, including neurotransmission, immune responses, cell growth, and cell death (Khakh and Burnstock, 2009). In mammalian cells, plasma membrane-localized receptors, purinoceptors of the P2X and P2Y classes, bind ATP as well as other nucleotides at the cell surface to activate intracellular signaling cascades via second messengers. Binding of extracellular ATP to P2X receptors gates calcium influx, whereas activation of P2Y receptors stimulates the recruitment of heterotrimeric G proteins to trigger cytoplasmic signaling and gene expression. As a common phenomenon, the activated receptors induce the elevation of [Ca2+]cyt, which in turn activates the production of downstream messengers such as nitric oxide and reactive oxygen species (ROS; Shen et al., 2005; Fields and Burnstock, 2006).A possible physiological role for extracellular ATP in plants was first reported in studies in which exogenously applied ATP was found to stimulate closure of the Venus flytrap (Dionaea muscipula; Jaffe, 1973), to induce the formation of nucleases in excised Avena leaves (Udvardy and Farkas, 1973), and to induce potassium ion uptake into cells of maize (Zea mays) leaf slices (Lüttge et al., 1974). Over the past several years, extracellular ATP was found to be an important signaling compound in plants that induces various plant responses, including root-hair growth (Lew and Dearnaley, 2000; Kim et al., 2006), stress responses (Thomas et al., 2000; Jeter et al., 2004; Song et al., 2006), gravitropism (Tang et al., 2003), cell viability (Chivasa et al., 2005), pathogen responses (Chivasa et al., 2009), and thigmotropism (Weerasinghe et al., 2009). The release of extracellular ATP from root cells was directly imaged by Kim et al. (2006) using a luciferase construct engineered to bind to plant cell wall cellulose. Recently, using this reporter, Weerasinghe et al. (2009) measured the release of ATP from root cells in response to touch. This documentation of the presence of extracellular ATP in plants at levels sufficient to induce cellular responses suggests that extracellular ATP likely plays an important role throughout plant growth and development. However, no P2 receptor homologs have been identified in plants, despite the fact that plants share a number of cellular responses to ATP with animal cells. For example, the addition of exogenous ATP or ADP triggers an increase in [Ca2+]cyt levels in whole seedlings, dissected root tissues, and root epidermal protoplasts of Arabidopsis (Arabidopsis thaliana; Demidchik et al., 2003, 2009; Jeter et al., 2004). The production of ROS in response to ATP addition was detected in various plant tissues (Kim et al., 2006; Song et al., 2006; Wu et al., 2008; Demidchik et al., 2009). More recently, the plasma membrane NADPH oxidase RBOHC (for respiratory burst oxidase homolog C) was shown to be required for extracellular ATP-induced ROS production in Arabidopsis primary roots (Demidchik et al., 2009). Extracellular ATP also stimulates the production of nitric oxide in tomato (Solanum lycopersicum) culture cells and in Salvia miltiorrhiza hairy roots (Foresi et al., 2007; Wu and Wu, 2008). These reports suggest that extracellular ATP signals across the plasma membrane by triggering elevation in [Ca2+]cyt, which activates the production of downstream messengers. Ultimately, these cell responses induce the expression of various genes, such as MAPKs, LOX, and ACS6 (Jeter et al., 2004; Song et al., 2006), and cause physiological responses, as described above.In animal cells, extracellular ATP-evoked elevations in [Ca2+]cyt are often observed in the form of oscillations that result from the transient opening of Ca2+ channels located either in the plasma membrane or in cytosolic Ca2+ stores. Intracellular calcium release is often mediated through phospholipase C (PLC)-mediated signaling coupled to heterotrimeric G proteins (Mahoney et al., 1992; Visegrady et al., 2000; Hanley et al., 2004). In plants, plasma membrane Ca2+-permeable channels are known to contribute to extracellular ATP-induced [Ca2+]cyt elevation (Demidchik et al., 2009). However, neither the mechanisms underlying extracellular ATP-evoked Ca2+ signaling nor the possible involvement of heterotrimeric G proteins has been characterized in plants.In order to explore their roles as possible ligands of putative nucleotide receptors, the plant [Ca2+]cyt response to six different nucleotides (Fig. 1A) was measured using Arabidopsis seedlings expressing one of two [Ca2+]cyt sensors, either aequorin or the fluorescence resonance energy transfer (FRET)-based Ca2+ sensor Yellow Cameleon 3.6 (YC3.6). The pyrimidine nucleotide CTP as well as the purine nucleotides ATP and GTP induced a strong elevation of [Ca2+]cyt in seedlings. Interestingly, the effects of all the nucleotides on Ca2+ signaling were negatively regulated by a heterotrimeric G protein β-subunit, AGB1. The addition of ATP to aequorin-expressing seedlings induced distinct [Ca2+]cyt oscillations in the presence of the apyrase inhibitor NGXT191. However, in the absence of this inhibitor, such [Ca2+]cyt oscillations could be localized to specific root cell layers using YC3.6 fluorescence. Given the importance of [Ca2+]cyt oscillations in intracellular signaling, the data suggest an important, unexplored role of extracellular ATP in the plant signaling pathways.Open in a separate windowFigure 1.NTPs increase bioluminescence in aequorin-expressing transgenic Arabidopsis seedlings. A, Chemical structures of purine and pyrimidine derivatives. B, Individual 5-d-old aequorin seedlings were transferred to individual wells of a 96-well microplate and incubated overnight in reconstitution buffer containing coelenterazine. Each NTP was then applied at a final concentration of 100 μm. The line graph shows time-dependent changes in photon counts from representative wells of each treatment (bin size = 50 frames, 1 s, 20 bin smoothing). The inset shows a pseudocolored photon-counting image integrated over 400 s after nucleotide treatment calibrated to the inset scale.  相似文献   

11.
Although cytosolic free Ca2+ mobilization induced by microbe/pathogen-associated molecular patterns is postulated to play a pivotal role in innate immunity in plants, the molecular links between Ca2+ and downstream defense responses still remain largely unknown. Calcineurin B-like proteins (CBLs) act as Ca2+ sensors to activate specific protein kinases, CBL-interacting protein kinases (CIPKs). We here identified two CIPKs, OsCIPK14 and OsCIPK15, rapidly induced by microbe-associated molecular patterns, including chitooligosaccharides and xylanase (Trichoderma viride/ethylene-inducing xylanase [TvX/EIX]), in rice (Oryza sativa). Although they are located on different chromosomes, they have over 95% nucleotide sequence identity, including the surrounding genomic region, suggesting that they are duplicated genes. OsCIPK14/15 interacted with several OsCBLs through the FISL/NAF motif in yeast cells and showed the strongest interaction with OsCBL4. The recombinant OsCIPK14/15 proteins showed Mn2+-dependent protein kinase activity, which was enhanced both by deletion of their FISL/NAF motifs and by combination with OsCBL4. OsCIPK14/15-RNAi transgenic cell lines showed reduced sensitivity to TvX/EIX for the induction of a wide range of defense responses, including hypersensitive cell death, mitochondrial dysfunction, phytoalexin biosynthesis, and pathogenesis-related gene expression. On the other hand, TvX/EIX-induced cell death was enhanced in OsCIPK15-overexpressing lines. Our results suggest that OsCIPK14/15 play a crucial role in the microbe-associated molecular pattern-induced defense signaling pathway in rice cultured cells.Calcium ions regulate diverse cellular processes in plants as a ubiquitous internal second messenger, conveying signals received at the cell surface to the inside of the cell through spatial and temporal concentration changes that are decoded by an array of Ca2+ sensors (Reddy, 2001; Sanders et al., 2002; Yang and Poovaiah, 2003). Several families of Ca2+ sensors have been identified in higher plants. The best known are calmodulins (CaMs) and CaM-related proteins, which typically contain four EF-hand domains for Ca2+ binding (Zielinski, 1998). Unlike mammals, which possess single molecular species of CaM, plants have at least three distinct molecular species of CaM playing diverse physiological functions and whose expression is differently regulated (Yamakawa et al., 2001; Luan et al., 2002; Karita et al., 2004; Takabatake et al., 2007). The second major class is exemplified by the Ca2+-dependent protein kinases, which contain CaM-like Ca2+-binding domains and a kinase domain in a single protein (Harmon et al., 2000). In addition, a new family of Ca2+ sensors was identified as calcineurin B-like (CBL) proteins, which consists of proteins similar to both the regulatory β-subunit of calcineurin and the neuronal Ca2+ sensor in animals (Liu and Zhu, 1998; Kudla et al., 1999).Unlike CaMs, which interact with a large variety of target proteins, CBLs specifically target a family of protein kinases referred to as CBL-interacting protein kinases (CIPKs) or SnRK3s (for sucrose nonfermenting 1-related protein kinases type 3), which are most similar to the SNF family protein kinases in yeast (Luan et al., 2002). A database search of the Arabidopsis (Arabidopsis thaliana) genome sequence revealed 10 CBL and 25 CIPK homologues (Luan et al., 2002). Expression patterns of these Ca2+ sensors and protein kinases suggest their diverse functions in different signaling processes, including light, hormone, sugar, and stress responses (Batistic and Kudla, 2004). AtCBL4/Salt Overly Sensitive3 (SOS3) and AtCIPK24/SOS2 have been shown to play a key role in Ca2+-mediated salt stress adaptation (Zhu, 2002). The CBL-CIPK system has been shown to be involved in signaling pathways of abscisic acid (Kim et al., 2003a), sugar (Gong et al., 2002a), gibberellins (Hwang et al., 2005), salicylic acid (Mahajan et al., 2006), and K+ channel regulation (Li et al., 2006; Lee et al., 2007; for review, see Luan, 2009; Batistic and Kudla, 2009). However, physiological functions of most of the family members still remain largely unknown.Plants respond to pathogen attack by activating a variety of defense responses, including the generation of reactive oxygen species (ROS), synthesis of phytoalexins, expression of pathogenesis-related (PR) genes, cell cycle arrest, and mitochondrial dysfunction followed by a form of hypersensitive cell death known as the hypersensitive response (Nürnberger and Scheel, 2001; Greenberg and Yao, 2004; Kadota et al., 2004b). Transient membrane potential changes and Ca2+ influx are involved at the initial stage of defense responses (Kuchitsu et al., 1993; Pugin et al., 1997; Blume et al., 2000; Kadota et al., 2004a). Many kinds of defense responses are prevented when Ca2+ influx is compromised by Ca2+ chelators (Nürnberger and Scheel, 2001; Lecourieux et al., 2002). Since complex spatiotemporal patterns of cytosolic free Ca2+ concentration have been suggested to play pivotal roles in defense signaling (Nürnberger and Scheel, 2001; Sanders et al., 2002), multiple Ca2+ sensor proteins and their effectors should function in the defense signaling pathways. Although possible involvement of some CaM isoforms (Heo et al., 1999; Yamakawa et al., 2001), Ca2+-dependent protein kinases (Romeis et al., 2000, 2001; Ludwig et al., 2005; Kobayashi et al., 2007; Yoshioka et al., 2009), as well as Ca2+ regulation of EF-hand-containing enzymes such as ROS-generating NADPH oxidase (Ogasawara et al., 2008) have been suggested, other Ca2+-regulated signaling components still remain to be identified. No CBLs or CIPKs have so far been implicated as signaling components in defense signaling.N-Acetylchitooligosaccharides, chitin fragments, are microbe-associated molecular patterns (MAMPs) that are recognized by plasma membrane receptors (Kaku et al., 2006; Miya et al., 2007) and induce a variety of defense responses, such as membrane depolarization (Kuchitsu et al., 1993; Kikuyama et al., 1997), ion fluxes (Kuchitsu et al., 1997), ROS production (Kuchitsu et al., 1995), phytoalexin biosynthesis (Yamada et al., 1993), and induction of PR genes (Nishizawa et al., 1999), without hypersensitive cell death in rice (Oryza sativa) cells. In contrast, a fungal proteinaceous elicitor, xylanase from Trichoderma viride (TvX)/ethylene-inducing xylanase (EIX), which is recognized by two putative plasma membrane receptors, LeEix1 and LeEix2 (Ron and Avni, 2004), triggers hypersensitive cell death along with different kinetics of ROS production and activation of a mitogen-activated protein kinase, OsMPK6, previously named as OsMPK2 or OsMAPK6, in rice cells (Kurusu et al., 2005). These two fungal MAMPs thus provide excellent model systems to study innate immunity in rice cells.This study identified two CIPKs involved in various MAMP-induced layers of defense responses, including PR gene expression, phytoalexin biosynthesis, mitochondrial dysfunction, and cell death, in rice. Molecular characterization of these CIPKs, including interaction with the putative Ca2+ sensors as well as their physiological functions, is discussed.  相似文献   

12.
Fluorescence resonance energy transfer-sensitized emission of the yellow cameleon 3.60 was used to study the dynamics of cytoplasmic calcium ([Ca2+]cyt) in different zones of living Arabidopsis (Arabidopsis thaliana) roots. Transient elevations of [Ca2+]cyt were observed in response to glutamic acid (Glu), ATP, and aluminum (Al3+). Each chemical induced a [Ca2+]cyt signature that differed among the three treatments in regard to the onset, duration, and shape of the response. Glu and ATP triggered patterns of [Ca2+]cyt increases that were similar among the different root zones, whereas Al3+ evoked [Ca2+]cyt transients that had monophasic and biphasic shapes, most notably in the root transition zone. The Al3+-induced [Ca2+]cyt increases generally started in the maturation zone and propagated toward the cap, while the earliest [Ca2+]cyt response after Glu or ATP treatment occurred in an area that encompassed the meristem and elongation zone. The biphasic [Ca2+]cyt signature resulting from Al3+ treatment originated mostly from cortical cells located at 300 to 500 μ m from the root tip, which could be triggered in part through ligand-gated Glu receptors. Lanthanum and gadolinium, cations commonly used as Ca2+ channel blockers, elicited [Ca2+]cyt responses similar to those induced by Al3+. The trivalent ion-induced [Ca2+]cyt signatures in roots of an Al3+-resistant and an Al3+-sensitive mutant were similar to those of wild-type plants, indicating that the early [Ca2+]cyt changes we report here may not be tightly linked to Al3+ toxicity but rather to a general response to trivalent cations.The role of calcium ions (Ca2+) as a ubiquitous cellular messenger in animal and plant cells is well established (Berridge et al., 2000; Sanders et al., 2002; Ng and McAinsh, 2003). Cellular signal transduction pathways are elicited as a result of fluctuations of free Ca2+ in the cytoplasm ([Ca2+]cyt) in response to external and intracellular signals. These changes in [Ca2+]cyt influence numerous cellular processes, including vesicle trafficking, cell metabolism, cell proliferation and elongation, stomatal opening and closure, seed and pollen grain germination, fertilization, ion transport, and cytoskeletal organization (Hepler, 2005). [Ca2+]cyt fluctuations occur because cells have a Ca2+ signaling “toolkit” (Berridge et al., 2000) composed of on/off switches and a multitude of Ca2+-binding proteins. The on switches depend on membrane-localized Ca2+ channels that control the entry of Ca2+ into the cytosol (Piñeros and Tester, 1995, 1997; Thion et al., 1998; Kiegle et al., 2000a; White et al., 2000; Demidchik et al., 2002; Miedema et al., 2008). On the other hand, the off switches consist of a family of Ca2+-ATPases and Ca2+/H+ exchangers in the plasma membrane or endomembrane that remove Ca2+ from the cytosol, bringing the [Ca2+]cyt down to the initial resting level (Lee et al., 2007; Li et al., 2008).The numerous cellular processes regulated by Ca2+ have led investigators to ask how specificity in Ca2+ signaling is maintained. It has been proposed that specificity in Ca2+ signaling is achieved because a particular stimulus elicits a distinct Ca2+ signature, which is defined by the timing, magnitude, and frequency of [Ca2+]cyt changes. For instance, tip-growing plant cells such as root hairs and pollen tubes exhibit oscillatory elevations in [Ca2+]cyt that partly mirror the oscillatory nature of growth in these cell types (Cárdenas et al., 2008; Monshausen et al., 2008). Another example is nuclear Ca2+ spiking in root hairs of legumes exposed to NOD factors (Oldroyd and Downie, 2006; Peiter et al., 2007). Recently, it was shown that mechanical forces applied to an Arabidopsis (Arabidopsis thaliana) root can trigger a stimulus-specific [Ca2+]cyt response (Monshausen et al., 2009). Translating the Ca2+ signature into a defined cellular response is governed by a number of Ca2+-binding proteins such as calreticulin that act as [Ca2+]cyt buffers, which shape both the amplitude and duration of the Ca2+ signal or Ca2+ sensors such as calmodulin that impact other downstream cellular effectors (Berridge et al., 2000; White and Broadley, 2003; Hepler, 2005).A deeper understanding of Ca2+ signaling mechanisms in plants has been driven in large part by our ability to monitor dynamic changes in [Ca2+]cyt in the cell. Such measurements have been conducted using Ca2+-sensitive fluorescent indicator dyes (e.g. Indo and Fura), the luminescent protein aequorin (Knight et al., 1991, 1996; Legué et al., 1997; Wymer et al., 1997; Cárdenas et al., 2008), and more recently the yellow cameleon (YC) Ca2+ sensor, a chimeric protein that relies on fluorescence resonance energy transfer (FRET) as an indicator of [Ca2+]cyt changes in the cell (Allen et al., 1999; Miwa et al., 2006; Qi et al., 2006; Tang et al., 2007; Haruta et al., 2008). The YC reporter is composed of cyan fluorescent protein (CFP), the C terminus of calmodulin (CaM), a Gly-Gly linker, the CaM-binding domain of myosin light chain kinase (M13), and a yellow fluorescent protein (YFP; Miyawaki et al., 1997, 1999). The increased interaction between M13 and CaM upon binding of Ca2+ to CaM triggers a conformational change in the protein that brings the CFP and YFP in close proximity, resulting in enhanced FRET efficiency between the two fluorophores (Miyawaki, 2003). Thus, changes in FRET efficiency between CFP and YFP in the cameleon reporter are correlated with changes in [Ca2+]cyt.Since it was first introduced, improved versions of the cameleon reporter have been selected to more accurately report [Ca2+]cyt levels in the cell. For instance, the YC3.60 version was selected because of its resistance to cytoplasmic acidification and its higher dynamic range compared with the earlier cameleons. The higher dynamic range of YC3.60 is due to the use of a circularly permutated YFP called Venus (cpVenus) that is capable of absorbing a greater amount of energy from CFP (Nagai et al., 2004). Recently, the utility of YC3.60 for monitoring [Ca2+]cyt was demonstrated in Arabidopsis roots and pollen tubes using ratiometric imaging approaches (Monshausen et al., 2007, 2008, 2009; Haruta et al., 2008; Iwano et al., 2009). Here, we further evaluated YC3.60 as a [Ca2+]cyt sensor in plants using confocal microscopy and FRET-sensitized emission imaging. Unlike the direct ratiometric measurement of cpVenus and CFP reported in previous studies using YC3.60-expressing plants (Monshausen et al., 2008, 2009), the sensitized FRET approach we describe here involves the use of donor-only (CFP) and acceptor-only (YFP) controls, allowing us to correct for bleed-through and background signals from the FRET specimen (van Rheenen et al., 2004; Feige et al., 2005).For this study, we focused on monitoring [Ca2+]cyt changes in Arabidopsis seedling roots after aluminum (Al3+) exposure. Although Ca2+ signaling has long been implicated in mediating Al3+ responses in plants (Rengel and Zhang, 2003), the [Ca2+]cyt changes evoked by Al3+ reported in the literature have been inconsistent, and as such, the significance of these [Ca2+]cyt responses to mechanisms of Al3+ toxicity are not very clear. For instance, some studies reported that Al3+ caused a decrease in [Ca2+]cyt in plants (Jones et al., 1998b; Kawano et al., 2004), and others demonstrated elevated [Ca2+]cyt upon Al3+ treatment (Nichol and Oliveira, 1995; Lindberg and Strid, 1997; Jones et al., 1998a; Zhang and Rengel, 1999; Ma et al., 2002; Bhuja et al., 2004).Here, we report that Arabidopsis roots expressing the YC3.60 reporter exhibited transient elevations in [Ca2+]cyt within seconds of Al3+ exposure. The general pattern of [Ca2+]cyt changes observed after Al3+ treatment were distinct from those elicited by ATP or Glu, reinforcing the concept of specificity in [Ca2+]cyt signaling. We also observed root zone-dependent variations in the [Ca2+]cyt signatures evoked by Al3+ in regard to the shape, duration, and timing of the [Ca2+]cyt response. Other trivalent ions such as lanthanum (La3+) and gadolinium (Ga3+), which have been widely used as Ca2+ channel blockers (Monshausen et al., 2009), also induced a rapid rise in [Ca2+]cyt in root cells that were similar to those elicited by Al3+. Al3+, La3+, and Gd3+ elicited similar [Ca2+]cyt signatures in the Al3+-tolerant mutant alr104 (Larsen et al., 1998) and the Al3+-sensitive mutant als3-1 (Larsen et al., 2005), indicating that the early [Ca2+]cyt increases we report here may not be tightly linked to mechanisms of Al3+ toxicity but rather to a general trivalent cation response. Our study further shows that FRET-sensitized emission imaging of Arabidopsis roots expressing YC3.60 provides a robust method for documenting [Ca2+]cyt signatures in different root developmental zones that should be useful for future studies on Ca2+ signaling mechanisms in plants.  相似文献   

13.
Calcium plays an essential role in pollen tube tip growth. However, little is known concerning the molecular basis of the signaling pathways involved. Here, we identified Arabidopsis (Arabidopsis thaliana) CALCINEURIN B-LIKE PROTEIN-INTERACTING PROTEIN KINASE19 (CIPK19) as an important element to pollen tube growth through a functional survey for CIPK family members. The CIPK19 gene was specifically expressed in pollen grains and pollen tubes, and its overexpression induced severe loss of polarity in pollen tube growth. In the CIPK19 loss-of-function mutant, tube growth and polarity were significantly impaired, as demonstrated by both in vitro and in vivo pollen tube growth assays. Genetic analysis indicated that disruption of CIPK19 resulted in a male-specific transmission defect. Furthermore, loss of polarity induced by CIPK19 overexpression was associated with elevated cytosolic Ca2+ throughout the bulging tip, whereas LaCl3, a Ca2+ influx blocker, rescued CIPK19 overexpression-induced growth inhibition. Our results suggest that CIPK19 may be involved in maintaining Ca2+ homeostasis through its potential function in the modulation of Ca2+ influx.In flowering plants, fertilization is mediated by pollen tubes that extend directionally toward the ovule for sperm delivery (Krichevsky et al., 2007; Johnson, 2012). The formation of these elongated tubular structures is dependent on extreme polar growth (termed tip growth), in which cell expansion occurs exclusively in the very apical area (Yang, 2008; Rounds and Bezanilla, 2013). As this type of tip growth is amenable to genetic manipulation and cell biological analysis, the pollen tube is an excellent model system for the functional analysis of essential genes involved in polarity control and fertilization (Yang, 2008; Qin and Yang, 2011; Bloch and Yalovsky, 2013).It is well established that Ca2+ plays a critical role in pollen germination and tube growth (Konrad et al., 2011; Hepler et al., 2012). A steep tip-focused Ca2+ gradient has been detected at the tip of elongating pollen tubes (Rathore et al., 1991; Pierson et al., 1994; Hepler, 1997). In previous studies, artificial dissipation of the Ca2+ gradient seriously inhibited tip growth of pollen tubes, whereas elevation of internal Ca2+ level induced bending of the growth axis toward the zone of higher Ca2+. These studies suggest that Ca2+ not only controls pollen tube elongation but also modulates growth orientation (Miller et al., 1992; Malho et al., 1994; Malho and Trewavas, 1996; Hepler, 1997). These Ca2+ signatures are perceived and relayed to downstream responses by a complex toolkit of Ca2+-binding proteins that function as Ca2+ sensors (Yang and Poovaiah, 2003; Harper et al., 2004; Dodd et al., 2010).To date, four major Ca2+ sensor families have been identified in Arabidopsis (Arabidopsis thaliana), including calcium-dependent protein kinase, calmodulin (CaM), calmodulin-like (CML), and CALCINEURIN B-LIKE (CBL) proteins (Luan et al., 2002, 2009; Yang and Poovaiah, 2003; Harper et al., 2004). Calcium-dependent protein kinase family members comprise a kinase domain and a CaM-like domain in a single protein; thus, they act not only as a Ca2+ sensor but also as an effector, designated as sensor responders (Cheng et al., 2002). In contrast, CaM, CML, and CBL proteins do not have any enzymatic domains but transmit Ca2+ signals to downstream targets via Ca2+-dependent protein-protein interactions. Therefore, they have been designated as sensor relays (McCormack et al., 2005). While CaM and CML proteins interact with a diverse array of target proteins, it is generally accepted that CBLs interact specifically with a group of Ser/Thr protein kinases termed CALCINEURIN B-LIKE PROTEIN-INTERACTING PROTEIN KINASEs (CIPKs; Luan et al., 2002; Kolukisaoglu et al., 2004).In Arabidopsis, several CBLs coupled with their target CIPKs have been demonstrated to function in the regulation of ion homeostasis and stress responses (Luan et al., 2009). Under salt stress, SALT OVERLY SENSITIVE3 (SOS3)/CBL4-SOS2/CIPK24 regulate SOS1 at the plasma membrane for Na+ exclusion, whereas CBL10-CIPK24 complexes appear to regulate Na+ sequestration at the tonoplast (Liu et al., 2000; Qiu et al., 2002; Kim et al., 2007; Quan et al., 2007). For low-K+ stress, CBL1 and CBL9, with 87% amino acid sequence identity, interact with CIPK23, which regulates a voltage-gated ion channel (ARABIDOPSIS K+ TRANSPORTER1) to mediate the uptake of K+ in root hairs (Li et al., 2006; Xu et al., 2006; Cheong et al., 2007). In addition, CBL1 integrates plant responses to cold, drought, salinity, and hyperosmotic stresses (Albrecht et al., 2003; Cheong et al., 2003), and CBL9 is involved in abscisic acid signaling and biosynthesis during seed germination (Pandey et al., 2004). Over the past decade, the functions of CBL-CIPK complexes in abiotic stress tolerance have been studied extensively, but only limited studies focus on CBL family members in pollen tube growth. For example, CBL3 overexpression caused a defective phenotype in pollen tube growth (Zhou et al., 2009). Overexpression of CBL1 or its closest homolog CBL9 inhibited pollen germination and perturbed tube growth at high external K+, whereas disruption of CBL1 and CBL9 leads to a significantly reduced growth rate of pollen tubes under low-K+ conditions (Mähs et al., 2013). The potential roles of CIPKs in pollen tubes so far appear to be completely unknown.In this study, we demonstrated that Arabidopsis CIPK19, a CIPK specifically expressed in pollen grains and pollen tubes, functions in pollen tube tip growth, providing a new insight into the function of the CBL-CIPK network in the control of growth polarity during pollen tube extension in fertilization.  相似文献   

14.
Metabolic syndrome (MetS), a compilation of associated risk factors, increases the risk of type 2 diabetes and coronary artery disease (CAD, atherosclerosis), which can progress to the point of artery occlusion. Stents are the primary interventional treatment for occlusive CAD, and patients with MetS and hyperinsulinemia have increased restenosis. Because of its thrifty genotype, the Ossabaw pig is a model of MetS. We tested the hypothesis that, when fed high-fat diet, Ossabaw swine develop more features of MetS, greater native CAD, and greater stent-induced CAD than do Yucatan swine. Animals of each breed were divided randomly into 2 groups and fed 2 different calorie-matched diets for 40 wk: control diet (C) and high-fat, high-cholesterol atherogenic diet (H). A bare metal stent was placed in the circumflex artery, and pigs were allowed to recover for 3 wk. Characteristics of MetS, macrovascular and microvascular CAD, in-stent stenosis, and Ca2+ signaling in coronary smooth muscle cells were evaluated. MetS characteristics including, obesity, glucose intolerance, hyperinsulinemia, and elevated arterial pressure were elevated in Ossabaw swine compared to Yucatan swine. Ossabaw swine with MetS had more extensive and diffuse native CAD and in-stent stenosis and impaired coronary blood flow regulation compared with Yucatan. In-stent atherosclerotic lesions in Ossabaw coronary arteries were less fibrous and more cellular. Coronary smooth muscle cells from Ossabaw had impaired Ca2+ efflux and intracellular sequestration versus cells from Yucatan swine. Therefore, Ossabaw swine are a superior model of MetS, subsequent CAD, and cellular Ca2+ signaling defects, whereas Yucatan swine are leaner and relatively resistant to MetS and CAD.Abbreviations: CAD, coronary artery disease; CSM, coronary smooth muscle; IVGTT, intravenous glucose tolerance test; MetS, metabolic syndrome; SERCA, sarco–endoplasmic reticulum Ca2+ ATPase; ET1, endothelin 1; SOCE, store-operated Ca2+ entryAtherosclerotic coronary artery disease (CAD) is increased at least 2-fold in patients with metabolic syndrome (MetS)27 and is accompanied by marked microvascular dysfunction that further impairs coronary blood flow.10 MetS generally is diagnosed by the presence of 3 or more of the following conditions: obesity, insulin resistance, glucose intolerance, dyslipidemia, and hypertension.17,28 There is strong support for the role of the hyperinsulinemia component of MetS in increased restenosis after percutaneous coronary interventions.74,75,84,85 Further, our group has shown that severe coronary microvascular dysfunction occurs in MetS.5 Because MetS (so-called ‘prediabetes’) affects as much as 27% of the United States population, is increasing dramatically in prevalence,94 and can progress to type 2 diabetes, there is great need for basic research using animal models that accurately mimic MetS and the accompanying CAD. Clearly, there is need for study of MetS-induced CAD and in-stent stenosis and the underlying cellular and molecular mechanisms.Mice, rats, and swine are known to recapitulate MetS;3,12,36,60,71,72 however, none of these models fully reproduce the combined symptoms of MetS and CAD. Further, transgenic mouse models are simply not adequate for coronary vascular interventions using stents identical to those used in humans,18,23,38,55,57,79,83,86 a step that is essential for translation to the clinic. Yucatan and domestic swine are commonly used large animal models for study of cardiovascular disease due to their ability to mimic the neointimal formation and thrombosis observed in humans.86 For example, several laboratories have produced severe CAD in swine,8,24,51,61,62,68,91 but through toxin-induced pancreatic β-cell ablation and feeding of an atherogenic diet, rather than as a natural development subsequent to MetS or diabetes. Currently, there is a paucity of large animal models that reproduce MetS and CAD.3Research on the obesity-prone Ossabaw miniature swine59 clearly indicates that these animals develop MetS and cardiovascular disease when fed a high-calorie atherogenic diet,4,5,9,16,19,42,50,52,83,92 Female Ossabaw swine on this type of diet nearly doubled their percentage body fat in only 9 wk, showed insulin resistance, impaired glucose tolerance, dyslipidemia (profound increase in the ratio of low-density to high-density lipoprotein cholesterol, hypertriglyceridemia), hypertension, and early coronary atherosclerosis.16 These data contrast with those from male Yucatan miniature pigs, which did not develop MetS even after 20 wk on a comparable excess calorie atherogenic diet.8,68,95 Yucatan swine do not develop MetS through diet manipulation, unlike Ossabaw swine, which consistently recapitulate all MetS characteristics. However, important differences in study design have not allowed direct comparison between Yucatan and Ossabaw swine.Cytosolic Ca2+ signaling is involved in ‘phenotypic modulation’ of coronary smooth muscle (CSM), as characterized by proliferation and migration in several in vitro cell culture models33,35,89,90 and in vivo rodent models of the peripheral circulation (for example, reference 51). The Yucatan swine model of diabetic dyslipidemia shows altered Ca2+ extrusion,96 Ca2+ sequestration by the sarcoplasmic reticulum,32,34,98 and Ca2+ influx through voltage-gated Ca2+ channels.98 Currently, Ca2+ signaling has not been compared directly between MetS Ossabaw and Yucatan swine CSM. Therefore, the purpose of the present study was to test the hypothesis that compared with Yucatan swine on calorie-matched standard chow (for example, Yucatan maintenance diet8,95) and atherogenic diets, Ossabaw swine have a greater propensity to MetS and CAD with impaired coronary microvascular dysfunction and Ca2+ handling in CSM.  相似文献   

15.
NAD metabolism regulates diverse biological processes, including ageing, circadian rhythm and axon survival. Axons depend on the activity of the central enzyme in NAD biosynthesis, nicotinamide mononucleotide adenylyltransferase 2 (NMNAT2), for their maintenance and degenerate rapidly when this activity is lost. However, whether axon survival is regulated by the supply of NAD or by another action of this enzyme remains unclear. Here we show that the nucleotide precursor of NAD, nicotinamide mononucleotide (NMN), accumulates after nerve injury and promotes axon degeneration. Inhibitors of NMN-synthesising enzyme NAMPT confer robust morphological and functional protection of injured axons and synapses despite lowering NAD. Exogenous NMN abolishes this protection, suggesting that NMN accumulation within axons after NMNAT2 degradation could promote degeneration. Ectopic expression of NMN deamidase, a bacterial NMN-scavenging enzyme, prolongs survival of injured axons, providing genetic evidence to support such a mechanism. NMN rises prior to degeneration and both the NAMPT inhibitor FK866 and the axon protective protein WldS prevent this rise. These data indicate that the mechanism by which NMNAT and the related WldS protein promote axon survival is by limiting NMN accumulation. They indicate a novel physiological function for NMN in mammals and reveal an unexpected link between new strategies for cancer chemotherapy and the treatment of axonopathies.Axon degeneration in disease shares features with the progressive breakdown of the distal segment of severed axons as described by Augustus Waller in 1850 and named Wallerian degeneration.1 The serendipitous discovery of Wallerian degeneration slow (WldS) mice, where transected axons survive 10 times longer than in wild types (WTs),2 suggested that axon degeneration is a regulated process, akin to apoptosis of the cell bodies but distinct in molecular terms.3,4 This process appears conserved in rats, flies, zebrafish and humans.5, 6, 7, 8 WldS blocks axon degeneration in some disease models, indicating a mechanistic similarity.3 Therefore understanding the pathway it influences is an excellent route towards novel therapeutic strategies.WldS is a modified nicotinamide mononucleotide adenylyltransferase 1 (NMNAT1) enzyme, whose N-terminal extension partially relocates NMNAT1 from nuclei to axons, conferring gain of function.9,10 In mammals, three NMNAT isoforms, nuclear NMNAT1, cytoplasmic NMNAT2 and mitochondrial NMNAT3, catalyse nicotinamide adenine dinucleotide (NAD) synthesis from nicotinamide mononucleotide (NMN) and adenosine triphosphate (ATP; Figure 1a).11,12 Several reports indicate WldS protects injured axons by maintaining axonal NMNAT activity.13, 14, 15 In WT injured axons, without WldS, NMNAT activity falls when the labile, endogenous axonal isoform, NMNAT2, is no longer transported from cell bodies.16 NMNAT2 is required for axon maintenance16 and for axon growth in vivo and in vitro,17,18 and modulation of its stability by palmitoylation19 or ubiquitin-dependent processes both in mice or when ectopically expressed in Drosophila19, 20, 21 has a corresponding effect on axon survival.Open in a separate windowFigure 1FK866 acts within axons to delay degeneration after injury. (a) The salvage pathway of NAD biosynthesis from nicotinamide (Nam) and nicotinic acid (Na). Only NAD biosynthesis from Nam is sensitive to FK866, which potently inhibits NAMPT while having no effect on nicotinic acid phosphoribosyltransferase (NaPRT).29 The reaction catalysed by bacterial NMN deamidase is also shown. (b) SCG explants were treated with 100 nM FK866 for the indicated times, and then the whole explants (top panel) or the cell bodies (bottom left panel) and neurite fractions (bottom right panel) were separately collected. NAD was determined with an HPLC-based method (see Materials and Methods; n=3, mean and S.D. shown). (c) SCG neurites untreated (top panels) or treated with 100 nM FK866 the day before transection (bottom panels) and imaged after transection at the indicated time points. (d) SCG explants were treated with 100 nM FK866 1 day before or at the indicated times after cutting their neurites. Degeneration index was calculated from three fields in 2–4 independent experiments. The effect of treatment is highly significant when the drug is preincubated or added at 0–4 h after cut (mean ±S.E.M., n=6–12, one-way ANOVA followed by Bonferroni''s post-hoc test, *P<0.05, **P<0.01, ***P<0.001, ****P<0.0001, compared with untreated)WldS partially colocalizes with mitochondria14,22 and was shown to increase mitochondria motility and Ca2+-buffering capacity.23 Inhibiting mitochondrial permeability transition pore protects degenerating axons.24 However, WldS is protective in axons devoid of mitochondria,8 and targeting a cytosolic variant of NMNAT2 to mitochondria abolished its protective effect,19 suggesting a late mitochondrial involvement in Wallerian degeneration.Despite the importance of NMNAT activity in axon survival and degeneration, the molecular players remain elusive. Although NMNAT activity is required for protection,13 the hypothesis that increased NAD levels are responsible25,26 does not fit some data.27,28While further investigating the role of NAD, we found that blocking nicotinamide phosphoribosyltransferase (NAMPT, the enzyme preceding NMNAT, Figure 1a), was surprisingly axon-protective despite lowering NAD. NAMPT catalyses the synthesis of NMNAT-substrate NMN, the rate-limiting step in the NAD salvage pathway from nicotinamide (Nam) (Figure 1a). Here, we show that NMN accumulates after axon injury, and we provide genetic and pharmacological evidence supporting a role for this NMN increase in axon degeneration when NMNAT2 is depleted. We reveal an unexpected new direction for research into the degenerative mechanism, a novel class of protective proteins and new players in an axon-degeneration pathway sensitive to drugs under development for cancer.  相似文献   

16.
17.
A segmental tibial defect model in a large animal can provide a basis for testing materials and techniques for use in nonunions and severe trauma. This study reports the rationale behind establishing such a model and its design and conclusions. After ethics approval of the study, aged ewes (older than 5 y; n = 12) were enrolled. A 5-cm mid diaphyseal osteoperiosteal defect was made in the left tibia and was stabilized by using an 8-mm stainless-steel cross-locked intramedullary nail. Sheep were euthanized at 12 wk after surgery and evaluated by using radiography, microCT, and soft-tissue histology techniques. Radiology confirmed a lack of hard tissue callus bridging across the defect. Volumetric analysis based on microCT showed bone growth across the 16.5-cm3 defect of 1.82 ± 0.94 cm3. Histologic sections of the bridging tissues revealed callus originating from both the periosteal and endosteal surfaces, with fibrous tissue completing the bridging in all instances. Immunohistochemistry was used to evaluate the quality of the healing response. Clinical, radiographic, and histologic union was not achieved by 12 wk. This model may be effective for the investigation of surgical techniques and healing adjuncts for nonunion cases, where severe traumatic injury has led to significant bone loss.Abbreviations: BMP2, bone morphogenic protein 2; CATK, cathepsin K; VEGF, vascular endothelial growth factorThe human tibia is the most frequently broken long bone, often with significant bone loss.4 Segmental tibial defects can occur as a result of large tumor removal, trauma such as motor vehicle accidents, and more recently, blast injuries as seen with the escalating number of global conflicts. Treatment of these large bone and surrounding soft tissue defects is an ongoing, costly, and challenging clinical problem; no surgical technique has currently achieved preeminence.4 The general consensus on factors that affect healing include concomitant disease, age, and degree of trauma.5 When the first 2 factors, which are patient-related, are removed from the equation, healing is influenced by the size, anatomic location, and soft-tissue coverage of the defect. The ability to study these situations in a well-controlled, robust, and reproducible preclinical model would be advantageous to help establish effective surgical techniques and evaluate implants and materials.A literature review revealed that many ovine models for bone defects have been used, but all have limitations6,12,14,15,20,21,24,25,27,31,37,39,40 (Figure 1). Variations in protocols, such as age of the animals, size of the defect, and the bone and stabilization techniques used, limit meaningful comparison between studies.33,34 Although some studies have investigated material performance in the healing of defects, they did not rigorously quantify control defects,17,20 and others used no controls at all.39 There is often no explanation regarding the use of a particular defect size, leading to the question of whether the defect size was critical.24 The choice of bone used has been also varied; the femur,15 tibia,37 and metatarsus40 have all been studied. A noncritical-size defect implies that healing would eventually occur without the presence of any graft materials. One study,12 for example, used a 3-cm defect at an average of 1.8 times the diameter of the tibias in question and found that empty controls achieved as much as 26% of the stiffness of an intact tibia after 12 wk. Stabilization methods include plating,21,40 external fixtures,20 intramedullary nails,6,16 and a combination of intramedullary nails and plating.37Open in a separate windowFigure 1.A limited summary of the many studies where a segmental tibial has been used with their references.The criteria used in the present study for a critical-size segmental tibial defect model were based on the following factors. The ovine tibia closely resembles that of the human tibia in terms of size, shape, and physical properties and is commonly used when studying human orthopedic diseases.26,34 Intramedullary nailing has become the most commonly used method of tibial fracture fixation in human orthopedic surgery.8,22 An 8-mm intramedullary nail is commonly used in the treatment of human fractures, further confirming the size similarity between the ovine and human tibiae.19The aim of this study was to establish and characterize a preclinical ovine 5-cm osteoperiosteal critical-size tibial segmental defect model in mature sheep. The endpoints included those commonly used clinically, such as radiography and microCT. Histology to investigate the degree of healing and immunohistochemistry to characterize the healing process were included to complete the evaluation process.  相似文献   

18.
Excessive N-methyl-D-aspartate receptor (NMDAR) activation and the resulting activation of neuronal nitric oxide synthase (nNOS) cause neuronal injury. Homer1b/c facilitates NMDAR-PSD95-nNOS complex interactions, and Homer1a is a negative competitor of Homer1b/c. We report that Homer1a was both upregulated by and protected against NMDA-induced neuronal injury in vitro and in vivo. The neuroprotective activity of Homer1a was associated with NMDA-induced Ca2+ influx, oxidative stress and the resultant downstream signaling activation. Additionally, we found that Homer1a functionally regulated NMDAR channel properties in neurons, but did not regulate recombinant NR1/NR2B receptors in HEK293 cells. Furthermore, we found that Homer1a detached the physical links among NR2B, PSD95 and nNOS and reduced the membrane distribution of NMDAR. NMDA-induced neuronal injury was more severe in Homer1a homozygous knockout mice (KO, Homer1a−/−) when compared with NMDA-induced neuronal injury in wild-type mice (WT, Homer1a+/+). Additionally, Homer1a overexpression in the cortex of Homer1a−/− mice alleviated NMDA-induced neuronal injury. These findings suggest that Homer1a may be a key neuroprotective endogenous molecule that protects against NMDA-induced neuronal injury by disassembling NR2B-PSD95-nNOS complexes and reducing the membrane distribution of NMDARs.Glutamate (Glu) acts on glutamate receptors, such as the N-methyl-D-aspartate receptor (NMDAR), and leads to neuronal hyper-excitability and death in a dose-dependent manner.1 NMDAR activation induces Ca2+ influx and specifically activates neuronal nitric oxide synthase (nNOS) and downstream signaling pathways.2, 3, 4 Ca2+ influx is involved in glutamate-induced apoptosis caused by the activation of apoptosis-related signaling pathways, mitochondrial dysfunction and ROS induction.3, 4 Additionally, nNOS has been reported to contribute to NMDA-induced excitotoxicity.5, 6 Considering that direct NMDAR inhibition has not yet demonstrated favorable efficacy in most clinic trails and further considering the remarkable role of nNOS in NMDA-induced neuronal death,7 measures that can effectively protect neurons from NMDA-induced neuronal injury are urgently needed and represent a worthwhile research goal.Homer proteins belong to the postsynaptic density (PSD) family and consist of two major groups: the short-form Homer proteins (Homer1a and Ania3) and the long-form Homer proteins (Homer1b/c, Homer2 and Homer3).8 Homer1b/c has a conserved N-terminal Ena/VASP homology 1 domain and binds to group I metabotropic glutamate receptors (mGluRs), inositol triphosphate receptors and Shank family proteins.9, 10, 11, 12 Homer1b/c regulates surface receptor expression,13, 14 clustering,15 transient receptor potential family channels and mGluRs coupled to ion channels.10, 16, 17, 18, 19 Additionally, because of its C-terminal coiled-coil (CC) domains, Homer1b/c can self-multimerize, form multiprotein complexes and facilitate signal transduction to downstream pathways. Homer1a, which lacks the CC domain, is believed to compete with constitutive Homer1b/c and disrupt the association of multiple Homer1b/c complexes.Notably, Homer1b/c can interact with the Glu-induced Ca2+ influx pathway by binding to Shank, a NMDAR complex adaptor protein (NMDAR-PSD95-GKAP-Shank-Homer1b/c).12, 20 Furthermore, Homer1a also interacts with Shank, NMDA, nNOS and other Homer1b/c target proteins. Homer1a has a negative regulatory role by physically replacing certain target proteins, and is involved in the regulation of a variety of cellular and molecular functions in neurological diseases.21, 22, 23, 24, 25 Nevertheless, the mechanisms of action and associations between Homer1a and NMDA-induced neuronal injury have not yet been studied. Here, we aimed to investigate the possible neuroprotective effects of Homer1a and explore the mechanisms underlying Homer1a activity in NMDA-induced neuronal injury.  相似文献   

19.
To grant faithful chromosome segregation, the spindle assembly checkpoint (SAC) delays mitosis exit until mitotic spindle assembly. An exceedingly prolonged mitosis, however, promotes cell death and by this means antimicrotubule cancer drugs (AMCDs), that impair spindle assembly, are believed to kill cancer cells. Despite malformed spindles, cancer cells can, however, slip through SAC, exit mitosis prematurely and resist killing. We show here that the Fcp1 phosphatase and Wee1, the cyclin B-dependent kinase (cdk) 1 inhibitory kinase, play a role for this slippage/resistance mechanism. During AMCD-induced prolonged mitosis, Fcp1-dependent Wee1 reactivation lowered cdk1 activity, weakening SAC-dependent mitotic arrest and leading to mitosis exit and survival. Conversely, genetic or chemical Wee1 inhibition strengthened the SAC, further extended mitosis, reduced antiapoptotic protein Mcl-1 to a minimum and potentiated killing in several, AMCD-treated cancer cell lines and primary human adult lymphoblastic leukemia cells. Thus, the Fcp1-Wee1-Cdk1 (FWC) axis affects SAC robustness and AMCDs sensitivity.The spindle assembly checkpoint (SAC) delays mitosis exit to coordinate anaphase onset with spindle assembly. To this end, SAC inhibits the ubiquitin ligase Anaphase-Promoting Complex/Cyclosome (APC/C) to prevent degradation of the anaphase inhibitor securin and cyclin B, the major mitotic cyclin B-dependent kinase 1 (cdk1) activator, until spindle assembly.1 However, by yet poorly understood mechanisms, exceedingly prolonging mitosis translates into cell death induction.2, 3, 4, 5, 6, 7 Although mechanistic details are still missing on how activation of cell death pathways is linked to mitosis duration, prolongation of mitosis appears crucial for the ability of antimicrotubule cancer drugs (AMCDs) to kill cancer cells.2, 3, 4, 5, 6, 7 These drugs, targeting microtubules, impede mitotic spindle assembly and delay mitosis exit by chronically activating the SAC. Use of these drugs is limited, however, by toxicity and resistance. A major mechanism for resistance is believed to reside in the ability of cancer cells to slip through the SAC and exit mitosis prematurely despite malformed spindles, thus resisting killing by limiting mitosis duration.2, 3, 4, 5, 6, 7 Under the AMCD treatment, cells either die in mitosis or exit mitosis, slipping through the SAC, without or abnormally dividing.2, 3, 4 Cells that exit mitosis either die at later stages or survive and stop dividing or proliferate, giving rise to resistance.2, 3, 4 Apart from a role for p53, what dictates cell fate is still unknown; however, it appears that the longer mitosis is protracted, the higher the chances for cell death pathway activation are.2, 3, 4, 5, 6, 7Although SAC is not required per se for killing,6 preventing SAC adaptation should improve the efficacy of AMCD by increasing mitosis duration.2, 3, 4, 5, 6, 7 Therefore, further understanding of the mechanisms by which cells override SAC may help to improve the current AMCD therapy. Several kinases are known to activate and sustain SAC, and cdk1 itself appears to be of primary relevance.1, 8, 9 By studying mitosis exit and SAC resolution, we recently reported a role for the Fcp1 phosphatase to bring about cdk1 inactivation.10, 11 Among Fcp1 targets, we identified cyclin degradation pathway components, such as Cdc20, an APC/C co-activator, USP44, a deubiquitinating enzyme, and Wee1.10, 11 Wee1 is a crucial kinase that controls the G2 phase by performing inhibitory phosphorylation of cdk1 at tyr-15 (Y15-cdk1). Wee1 is also in a feedback relationship with cdk1 itself that, in turn, can phosphorylate and inhibit Wee1 in an autoamplification loop to promote the G2-to-M phase transition.12 At mitosis exit, Fcp1 dephosphorylated Wee1 at threonine 239, a cdk1-dependent inhibitory phosphorylation, to dampen down the cdk1 autoamplification loop, and Cdc20 and USP44, to promote APC/C-dependent cyclin B degradation.10, 11, 12 In this study we analysed the Fcp1 relevance in SAC adaptation and AMCD sensitivity.  相似文献   

20.
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