共查询到20条相似文献,搜索用时 0 毫秒
1.
2.
植物胁迫信号转导过程中活性氧和促细胞分裂原活化蛋白激酶的调节作用 总被引:4,自引:0,他引:4
以H2O2为代表的活性氧(reactive oxygen species,ROS)和以促细胞分裂原活化蛋白激酶(mitogen-activated protein kinase,MAPK)为代表的蛋白激酶广泛存在于植物细胞并参与各种生理反应过程.生物胁迫条件下,一些MAP激酶特异性地调节氧化猝发(oxidative burst,OXB)和过敏反应(hypersensitive response,HR),水杨酸(salicylic acid,SA)诱导的MAP激酶(SA-induced protein kinase,SIPK)和ROS共同参与系统获得性抗性(systemic acquired resistance,SAR)的建立;SIPK、P38 MAPK等分别与H2O2共同调节臭氧、受伤和渗透胁迫等多种非生物胁迫生理反应.ROS和MAP激酶共同调节植物胁迫信号转导,但其机制尚需进一步的研究. 相似文献
3.
蛋白激酶与植物逆境信号传递途径 总被引:14,自引:0,他引:14
蛋白质的可逆磷酸化是细胞信号识别与转导的重要环节,蛋白激酶主要催化蛋白质的磷酸化作用,植物中已发现并分离了大量蛋白激酶及其基因,它们介导了植物激素和胞外环境信号等引起的多种生理生化反应。文章着重介绍分裂原激活蛋白激酶(MAPK)、钙依赖而钙调素不依赖的蛋白激酶(CDPK)、受体蛋白激酶(RPK)、核糖体蛋白激酶和转录调控蛋白激酶等多种蛋白激酶在植物逆境信号识别与转导中的作用。 相似文献
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.
Kristiina Laanemets Benjamin Brandt Junlin Li Ebe Merilo Yong-Fei Wang Malik M. Keshwani Susan S. Taylor Hannes Kollist Julian I. Schroeder 《Plant physiology》2013,163(2):504-513
Guard cells use compensatory feedback controls to adapt to conditions that produce excessively open stomata.In the past 15 years or more, many mutants that are impaired in stimulus-induced stomatal closing and opening have been identified and functionally characterized in Arabidopsis (Arabidopsis thaliana), leading to a mechanistic understanding of the guard cell signal transduction network. However, evidence has only recently emerged that mutations impairing stomatal closure, in particular those in slow anion channel SLOW ANION CHANNEL-ASSOCIATED1 (SLAC1), unexpectedly also exhibit slowed stomatal opening responses. Results suggest that this compensatory slowing of stomatal opening can be attributed to a calcium-dependent posttranslational down-regulation of stomatal opening mechanisms, including down-regulation of inward K+ channel activity. Here, we discuss this newly emerging stomatal compensatory feedback control model mediated via constitutive enhancement (priming) of intracellular Ca2+ sensitivity of ion channel activity. The CALCIUM-DEPENDENT PROTEIN KINASE6 (CPK6) is strongly activated by physiological Ca2+ elevations and a model is discussed and open questions are raised for cross talk among Ca2+-dependent and Ca2+-independent guard cell signal transduction pathways and Ca2+ sensitivity priming mechanisms.Stomatal pores formed by two guard cells enable CO2 uptake from the atmosphere, but also ensure leaf cooling and provide a pulling force for nutrient uptake from the soil via transpiration. These vitally important processes are inevitably accompanied by water loss through stomata. Stomatal opening and closure is caused by the uptake and release of osmotically active substances and is tightly regulated by signaling pathways that lead to the activation or inactivation of guard cell ion channels and pumps. Potassium ions enter guard cells through the inward-rectifying K+ channels (K+in) during stomatal opening and are released via outward-rectifying K+ channels during stomatal closure (Schroeder et al., 1987; Hosy et al., 2003; Roelfsema and Hedrich 2005). Cytosolic Ca2+, an important second messenger in plants, mediates ion channel regulation, particularly down-regulation of inward-conducting K+in channels and activation of S-type anion channels, thus mediating stomatal closure and inhibiting stomatal opening (Schroeder and Hagiwara, 1989; Dodd et al., 2010; Kim et al., 2010). Stomatal closure is initiated by anion efflux via the slow S-type anion channel SLAC1 (Negi et al., 2008; Vahisalu et al., 2008; Kollist et al., 2011) and the voltage-dependent rapid R-type anion channel QUICK-ACTIVATING ANION CHANNEL1 (Meyer et al. 2010; Sasaki et al., 2010).In recent years, advances have been made toward understanding mechanisms mediating abscisic acid (ABA)-induced stomatal closure (Cutler et al., 2010; Kim et al., 2010; Raghavendra et al., 2010). The core ABA signaling module, consisting of PYR/RCAR (for pyrabactin resistance 1/regulatory components of ABA receptors) receptors, clade A protein phosphatases (PP2Cs), SNF-related protein kinase OPEN STOMATA1 (OST1), and downstream targets, is Ca2+-independent (Ma et al., 2009; Park et al., 2009; Hubbard et al., 2010). However, ABA-induced stomatal closure was reduced to only 30% of the normal stomatal closure response under conditions that inhibited intracellular cytosolic free calcium ([Ca2+]cyt) elevations in Arabidopsis (Siegel et al., 2009), consistent with previous findings in other plants (De Silva et al., 1985; Schwartz, 1985; McAinsh et al., 1991; MacRobbie, 2000). Together these and other studies show the importance of [Ca2+]cyt for a robust ABA-induced stomatal closure. Here, we discuss Ca2+-dependent and Ca2+-independent signaling pathways in guard cells and open questions on how these may work together.Plants carrying mutations in the SLAC1 anion channel have innately more open stomata, and exhibit clear impairments in ABA-, elevated CO2-, Ca2+-, ozone-, air humidity-, darkness-, and hydrogen peroxide-induced stomatal closure (Negi et al., 2008; Vahisalu et al., 2008; Merilo et al., 2013). Recent research, however, unexpectedly revealed that mutations in SLAC1 also down-regulate stomatal opening mechanisms and slow down stomatal opening (Laanemets et al., 2013). 相似文献
6.
受体蛋白激酶是蛋白激酶家族中的重要一类。根据其胞外受体结构域的组成不同 ,植物受体蛋白激酶可划分为不同的类型。近些年的研究发现 ,蛋白激酶是植物发育和抗性反应中重要成分 ,是信号分子的重要受体 ,在信号传导过程中起着重要作用。随着对植物发育过程中信号传导机理认识的不断深入 ,人们有望通过操作植物发育过程向人们需要的方向发展 ,达到控制果实的大小和提高产量的目的。 相似文献
7.
Calcium/calmodulin-mediated signaling contributes in diverse roles in plant growth, development, and response to environmental stimuli.During calcium (Ca2+) signaling, decoding the stimulus-response coupling involves a set of Ca2+ sensor proteins or Ca2+-binding proteins (DeFalco et al., 2010a; Kudla et al., 2010). These proteins usually possess one or more classical helix-loop-helix elongation factor (EF) hand motifs. Three major types of Ca2+-sensor proteins in plants are calmodulin (CaM)/CaM-like proteins, calcium-dependent protein kinases (CDPKs), and calcineurin B-like proteins. As compared with animals, plant genomes encode more diversified Ca2+ sensors; with the exception of canonic CaM, all other types of Ca2+ sensors (CaM-like proteins, CDPKs, and calcineurin B-like proteins) are plant specific. The large population and unique structural composition of Ca2+-binding proteins and the diversity of the target proteins regulated by the Ca2+ sensors reflect the complexity of Ca2+ signaling, which helps plants adapt to the changing environment. This update will be limited primarily to discussions on CaM and CaM-binding proteins and the recent advances in Ca2+/CaM-mediated signaling.CaM is a conserved Ca2+-binding protein found in all eukaryotes. The discovery of CaM can be traced back to the 1970s. An activator of cyclic nucleotide phosphodiesterase was shown to be involved in the regulation of cAMP concentration, which was stimulated by Ca2+ (Kakiuchi and Yamazaki, 1970; Cheung, 1971). The activator was found to bind Ca2+ and was eventually named “calmodulin,” an abbreviation of Ca2+-modulated protein. Since its discovery over 40 years ago, CaM has been regarded as a model Ca2+-binding protein and has been subjected to intensive studies in biochemistry, cell biology, and molecular biology because of its importance in almost all aspects of cellular regulation (Poovaiah and Reddy, 1987, 1993; Bouche et al., 2005; DeFalco et al., 2010a; Du et al., 2011; Reddy et al., 2011b). Disruption or depletion of the single copy of the CaM gene in yeast (Saccharomyces cerevisiae) results in a recessive lethal mutation (Davis et al., 1986), suggesting that CaM has a critical role in eukaryotic cells.The structure of CaM has been well studied, and the prototype of CaM found in all eukaryotes has 149 amino acids with two globular domains, each containing two EF hands connected by a long flexible helix (Meador et al., 1993; Zhang et al., 1995; Yun et al., 2004; Ishida et al., 2009). As more and more genomes are sequenced, it is becoming clear that CaM belongs to a small gene family in plants. In the model plant Arabidopsis (Arabidopsis thaliana), seven CaM genes encode for four highly conserved isoforms (CaM1/4, CaM2/3/5, CaM6, and CaM7) that differ in only one to five amino acid residues. Loss-of-function mutations of individual CaMs indicate that the different CaMs may have overlapping yet different functions. For example, a loss of function in Arabidopsis AtCaM2 affects pollen germination (Landoni et al., 2010). Phenotypic analysis showed that in normal growth conditions, atcam2-2 plants were indistinguishable from the wild type, while genetic analysis showed a reduced transmission of the atcam2-2 allele through the male gametophyte, and in vitro pollen germination revealed a reduced level of germination in comparison with the wild type. However, the atcam3 knockout mutant showed a clear reduction in thermotolerance after heat treatment at 45°C for 50 min (Zhang et al., 2009). Overexpression of AtCaM3 in either the atcam3 knockout or wild-type background significantly rescued or increased the thermotolerance, respectively. Further analysis of individual CaM mutants under different stress conditions should reveal more on the functional significance of individual CaM genes. 相似文献
8.
Wenxiu Ye Daichi Muroyama Shintaro Munemasa Yoshimasa Nakamura Izumi C. Mori Yoshiyuki Murata 《Plant physiology》2013,163(2):591-599
Yeast elicitor (YEL) induces stomatal closure that is mediated by a Ca2+-dependent signaling pathway. A Ca2+-dependent protein kinase, CPK6, positively regulates activation of ion channels in abscisic acid and methyl jasmonate signaling, leading to stomatal closure in Arabidopsis (Arabidopsis thaliana). YEL also inhibits light-induced stomatal opening. However, it remains unknown whether CPK6 is involved in induction by YEL of stomatal closure or in inhibition by YEL of light-induced stomatal opening. In this study, we investigated the roles of CPK6 in induction by YEL of stomatal closure and inhibition by YEL of light-induced stomatal opening in Arabidopsis. Disruption of CPK6 gene impaired induction by YEL of stomatal closure and inhibition by YEL of light-induced stomatal opening. Activation by YEL of nonselective Ca2+-permeable cation channels was impaired in cpk6-2 guard cells, and transient elevations elicited by YEL in cytosolic-free Ca2+ concentration were suppressed in cpk6-2 and cpk6-1 guard cells. YEL activated slow anion channels in wild-type guard cells but not in cpk6-2 or cpk6-1 and inhibited inward-rectifying K+ channels in wild-type guard cells but not in cpk6-2 or cpk6-1. The cpk6-2 and cpk6-1 mutations inhibited YEL-induced hydrogen peroxide accumulation in guard cells and apoplast of rosette leaves but did not affect YEL-induced hydrogen peroxide production in the apoplast of rosette leaves. These results suggest that CPK6 positively functions in induction by YEL of stomatal closure and inhibition by YEL of light-induced stomatal opening in Arabidopsis and is a convergent point of signaling pathways for stomatal closure in response to abiotic and biotic stress.Stomata, formed by pairs of guard cells, play a critical role in regulation of plant CO2 uptake and water loss, thus critically influencing plant growth and water stress responsiveness. Guard cells respond to a variety of abiotic and biotic stimuli, such as light, drought, and pathogen attack (Israelsson et al., 2006; Shimazaki et al., 2007; Melotto et al., 2008).Elicitors derived from microbial surface mimic pathogen attack and induce stomatal closure in various plant species such as Solanum lycopersicum (Lee et al., 1999), Commelina communis (Lee et al., 1999), Hordeum vulgare (Koers et al., 2011), and Arabidopsis (Arabidopsis thaliana; Melotto et al., 2006; Khokon et al., 2010). Yeast elicitor (YEL) induces stomatal closure in Arabidopsis (Klüsener et al., 2002; Khokon et al., 2010; Salam et al., 2013). Our recent studies showed that YEL inhibits light-induced stomatal opening and that protein phosphorylation is involved in induction by YEL of stomatal closure and inhibition by YEL of light-induced stomatal opening (Salam et al., 2013).Cytosolic Ca2+ has long been recognized as a conserved second messenger in stomatal movement (Shimazaki et al., 2007; Roelfsema and Hedrich 2010; Hubbard et al., 2012). Elevation of cytosolic free Ca2+ concentration ([Ca2+]cyt) is triggered by influx of Ca2+ from apoplast and release of Ca2+ from intracellular stores in guard cell signaling (Leckie et al., 1998; Hamilton et al., 2000; Pei et al., 2000; Garcia-Mata et al., 2003; Lemtiri-Chlieh et al., 2003). The influx of Ca2+ is carried by nonselective Ca2+-permeable cation (ICa) channels that are activated by plasma membrane hyperpolarization and H2O2 (Pei et al., 2000; Murata et al., 2001; Kwak et al., 2003). Elevation of [Ca2+]cyt activates slow anion (S-type) channels and down-regulates inward-rectifying potassium (Kin) channels in guard cells (Schroeder and Hagiwara, 1989; Grabov and Blatt, 1999). The activation of S-type channels is a hallmark of stomatal closure, and the suppression of Kin channels is favorable to stomatal closure but not to stomatal opening (Pei et al., 1997; Kwak et al., 2001; Xue et al., 2011; Uraji et al., 2012).YEL induces stomatal closure with extracellular H2O2 production, intracellular H2O2 accumulation, activation of ICa channels, and transient [Ca2+]cyt elevations (Klüsener et al., 2002; Khokon et al., 2010). However, it remains to be clarified whether YEL activates S-type channels and inhibits Kin channels in guard cells.Calcium-dependent protein kinases (CDPKs) are regulators in Ca2+-dependent guard cell signaling (Mori et al., 2006; Zhu et al., 2007; Geiger et al., 2010, 2011; Zou et al., 2010; Munemasa et al., 2011; Brandt et al., 2012; Scherzer et al., 2012). In guard cells, CDPKs regulate activation of S-type and ICa channels and inhibition of Kin channels (Mori et al., 2006; Zou et al., 2010; Munemasa et al., 2011). A CDPK, CPK6, positively regulates activation of S-type channels and ICa channels without affecting H2O2 production in abscisic acid (ABA)- and methyl jasmonate (MeJA)-induced stomatal closure (Mori et al., 2006; Munemasa et al., 2011). CPK6 phosphorylates and activates SLOW ANION CHANNEL-ASSOCIATED1 expressed in Xenopus spp. oocyte (Brandt et al., 2012; Scherzer et al., 2012). These findings underline the role of CPK6 in regulation of ion channel activation and stomatal movement, leading us to test whether CPK6 regulates the induction by YEL of stomatal closure and inhibition by YEL of light-induced stomatal opening.In this study, we investigated activation of S-type channels and inhibition of Kin channels by YEL and roles of CPK6 in induction by YEL of stomatal closure and inhibition by YEL of light-induced stomatal opening. For this purpose, we examined the effects of mutation of CPK6 on induction by YEL of stomatal closure and inhibition by YEL of light-induced stomatal opening, activation of ICa channels, transient [Ca2+]cyt elevations, activation of S-type channels, inhibition of Kin channels, H2O2 production in leaves, and H2O2 accumulation in leaves and guard cells. 相似文献
9.
Biochemical Similarities between Soluble and Membrane-Bound Calcium-Dependent Protein Kinases of Barley 总被引:1,自引:3,他引:1
下载免费PDF全文

The soluble and membrane-bound forms of the calcium-dependent protein kinase from barley leaves (Hordeum vulgare L. cv. Borsoy) have been partially purified and compared. Both forms showed an active polypeptide of 37 kilodaltons on activity gels with incorporated histone as substrate. They eluted from chromatofocusing columns at an identical isoelectric point of pH 4.25 ± 0.2, and also comigrated on several other chromatographic affinity media including Matrex Gel Blue A, histone-agarose, phenyl-Sepharose, and heparin-agarose. Both activities comigrated with chicken ovalbumin during gel filtration through Sephacryl S-200, indicating a native molecular mass of 45 kilodaltons. The activities share a number of enzymatic properties including salt and pH dependence, free calcium stimulation profile, substrate specificity, and Km values. The soluble activity was shown to bind to artificial lipid vesicles. These data suggest strongly that the soluble and membrane-bound calcium-dependent protein kinases from barley are very closely related or even identical. 相似文献
10.
11.
钙依赖蛋白激酶(CDPKs)在植物钙信号转导中的作用 总被引:12,自引:0,他引:12
CDPKs在植物钙信号转导中起重要作用。本文介绍了植物钙信号转导及CDPKs的结构与生化性质,在此基础上,重点总结了CDPKs在植物钙信号转导中的潜在调节作用,包括基因表达、代谢、离子和水分的跨膜运输、细胞骨架的动态变化、气孔运动和生长发育等,并提出了在CDPKs研究中已达成的共识和需要解决的问题。 相似文献
12.
13.
14.
15.
《生物化学与生物物理进展》2006,(6)
A special issue on Calcium Signaling will be published in July issue of Acta Pharmacologica Sinica. From the following selected articles you will see that the approach we take is to give a broad picture on recent developments in the field of calcium sig… 相似文献
16.
Masatsugu Toyota Takuya Furuichi Masahiro Sokabe Hitoshi Tatsumi 《Plant physiology》2013,163(2):543-554
Gravity is a critical environmental factor affecting the morphology and functions of organisms on the Earth. Plants sense changes in the gravity vector (gravistimulation) and regulate their growth direction accordingly. In Arabidopsis (Arabidopsis thaliana) seedlings, gravistimulation, achieved by rotating the specimens under the ambient 1g of the Earth, is known to induce a biphasic (transient and sustained) increase in cytoplasmic calcium concentration ([Ca2+]c). However, the [Ca2+]c increase genuinely caused by gravistimulation has not been identified because gravistimulation is generally accompanied by rotation of specimens on the ground (1g), adding an additional mechanical signal to the treatment. Here, we demonstrate a gravistimulation-specific Ca2+ response in Arabidopsis seedlings by separating rotation from gravistimulation by using the microgravity (less than 10−4g) conditions provided by parabolic flights. Gravistimulation without rotating the specimen caused a sustained [Ca2+]c increase, which corresponds closely to the second sustained [Ca2+]c increase observed in ground experiments. The [Ca2+]c increases were analyzed under a variety of gravity intensities (e.g. 0.5g, 1.5g, or 2g) combined with rapid switching between hypergravity and microgravity, demonstrating that Arabidopsis seedlings possess a very rapid gravity-sensing mechanism linearly transducing a wide range of gravitational changes (0.5g–2g) into Ca2+ signals on a subsecond time scale.Calcium ion (Ca2+) functions as an intracellular second messenger in many signaling pathways in plants (White and Broadley, 2003; Hetherington and Brownlee, 2004; McAinsh and Pittman, 2009; Spalding and Harper, 2011). Endogenous and exogenous signals are spatiotemporally encoded by changing the free cytoplasmic concentration of Ca2+ ([Ca2+]c), which in turn triggers [Ca2+]c-dependent downstream signaling (Sanders et al., 2002; Dodd et al., 2010). A variety of [Ca2+]c increases induced by diverse environmental and developmental stimuli are reported, such as phytohormones (Allen et al., 2000), temperature (Plieth et al., 1999; Dodd et al., 2006), and touch (Knight et al., 1991; Monshausen et al., 2009). The [Ca2+]c increase couples each stimulus and appropriate physiological responses. In the Ca2+ signaling pathways, the stimulus-specific [Ca2+]c pattern (e.g. amplitude and oscillation) provide the critical information for cellular signaling (Scrase-Field and Knight, 2003; Dodd et al., 2010). Therefore, identification of the stimulus-specific [Ca2+]c signature is crucial for an understanding of the intracellular signaling pathways and physiological responses triggered by each stimulus, as shown in the case of cold acclimation (Knight et al., 1996; Knight and Knight, 2000).Plants often exhibit biphasic [Ca2+]c increases in response to environmental stimuli. Thus, slow cooling causes a fast [Ca2+]c transient followed by a second, extended [Ca2+]c increase in Arabidopsis (Arabidopsis thaliana; Plieth et al., 1999; Knight and Knight, 2000). The Ca2+ channel blocker lanthanum (La3+) attenuated the fast transient but not the following increase (Knight and Knight, 2000), suggesting that these two [Ca2+]c peaks have different origins. Similarly, hypoosmotic shock caused a biphasic [Ca2+]c increase in tobacco (Nicotiana tabacum) suspension-culture cells (Takahashi et al., 1997; Cessna et al., 1998). The first [Ca2+]c peak was inhibited by gadolinium (Gd3+), La3+, and the Ca2+ chelator EGTA (Takahashi et al., 1997; Cessna et al., 1998), whereas the second [Ca2+]c increase was inhibited by the intracellular Ca2+ store-depleting agent caffeine but not by EGTA (Cessna et al., 1998). The amplitude of the first [Ca2+]c peak affected the amplitude of the second increase and vice versa (Cessna et al., 1998). These results suggest that even though the two [Ca2+]c peaks originate from different Ca2+ fluxes (e.g. Ca2+ influx through the plasma membrane and Ca2+ release from subcellular stores, respectively), they are closely interrelated, showing the importance of the kinetic and pharmacological analyses of these [Ca2+]c increases.Changes in the gravity vector (gravistimulation) could work as crucial environmental stimuli in plants and are generally achieved by rotating the specimens (e.g. +180°) in ground experiments. Use of Arabidopsis seedlings expressing apoaequorin, a Ca2+-reporting photoprotein (Plieth and Trewavas, 2002; Toyota et al., 2008a), has revealed that gravistimulation induces a biphasic [Ca2+]c increase that may be involved in the sensory pathway for gravity perception/response (Pickard, 2007; Toyota and Gilroy, 2013) and the intracellular distribution of auxin transporters (Benjamins et al., 2003; Zhang et al., 2011). These two Ca2+ changes have different characteristics. The first transient [Ca2+]c increase depends on the rotational velocity but not angle, whereas the second sustained [Ca2+]c increase depends on the rotational angle but not velocity. The first [Ca2+]c transient was inhibited by Gd3+, La3+, and the Ca2+ chelator 1,2-bis(2-aminophenoxy)ethane-N,N,N′,N′-tetraacetic acid but not by ruthenium red (RR), whereas the second sustained [Ca2+]c increase was inhibited by all these chemicals. These results suggest that the first transient and second sustained [Ca2+]c increases are related to the rotational stimulation and the gravistimulation, respectively, and are mediated by distinct molecular mechanisms (Toyota et al., 2008a). However, it has not been demonstrated directly that the second sustained [Ca2+]c increase is induced solely by gravistimulation; it could be influenced by other factors, such as an interaction with the first transient [Ca2+]c increase (Cessna et al., 1998), vibration, and/or deformation of plants during the rotation.To elucidate the genuine Ca2+ signature in response to gravistimulation in plants, we separated rotation and gravistimulation under microgravity (μg; less than 10−4g) conditions provided by parabolic flight (PF). Using this approach, we were able to apply rotation and gravistimulation to plants separately (Fig. 1). When Arabidopsis seedlings were rotated +180° under μg conditions, the [Ca2+]c response to the rotation was transient and almost totally attenuated in a few seconds. Gravistimulation (transition from μg to 1.5g) was then applied to these prerotated specimens at the terminating phase of the PF. This gravistimulation without simultaneous rotation induced a sustained [Ca2+]c increase. The kinetic properties of this sustained [Ca2+]c increase were examined under different gravity intensities (0.5g–2g) and sequences of gravity intensity changes (Fig. 2A). This analysis revealed that gravistimulation-specific Ca2+ response has an almost linear dependency on gravitational acceleration (0.5g–2g) and an extremely rapid responsiveness of less than 1 s.Open in a separate windowFigure 1.Diagram of the experimental procedures for applying separately rotation and gravistimulation to Arabidopsis seedlings. Rotatory stimulation (green arrow) was applied by rotating the seedlings 180° under μg conditions, and 1.5g 180° rotation gravistimulation (blue arrow) was applied to the prerotated seedlings after μg.Open in a separate windowFigure 2.Acceleration, temperature, humidity, and pressure in an aircraft during flight experiments. A, Accelerations along x, y, and z axes in the aircraft during PF. The direction of flight (FWD) and coordinates (x, y, and z) are indicated in the bottom graph. The inset shows an enlargement of the acceleration along the z axis (gravitational acceleration) during μg conditions lasting for approximately 20 s. B, Temperature, humidity, and pressure in the aircraft during PF. Shaded areas in graphs denote the μg condition. 相似文献
17.
18.
19.
Posttranslational modifications (PTMs) of proteins greatly expand proteome diversity, increase functionality, and allow for rapid responses, all at relatively low costs for the cell. PTMs play key roles in plants through their impact on signaling, gene expression, protein stability and interactions, and enzyme kinetics. Following a brief discussion of the experimental and bioinformatics challenges of PTM identification, localization, and quantification (occupancy), a concise overview is provided of the major PTMs and their (potential) functional consequences in plants, with emphasis on plant metabolism. Classic examples that illustrate the regulation of plant metabolic enzymes and pathways by PTMs and their cross talk are summarized. Recent large-scale proteomics studies mapped many PTMs to a wide range of metabolic functions. Unraveling of the PTM code, i.e. a predictive understanding of the (combinatorial) consequences of PTMs, is needed to convert this growing wealth of data into an understanding of plant metabolic regulation.The primary amino acid sequence of proteins is defined by the translated mRNA, often followed by N- or C-terminal cleavages for preprocessing, maturation, and/or activation. Proteins can undergo further reversible or irreversible posttranslational modifications (PTMs) of specific amino acid residues. Proteins are directly responsible for the production of plant metabolites because they act as enzymes or as regulators of enzymes. Ultimately, most proteins in a plant cell can affect plant metabolism (e.g. through effects on plant gene expression, cell fate and development, structural support, transport, etc.). Many metabolic enzymes and their regulators undergo a variety of PTMs, possibly resulting in changes in oligomeric state, stabilization/degradation, and (de)activation (Huber and Hardin, 2004), and PTMs can facilitate the optimization of metabolic flux. However, the direct in vivo consequence of a PTM on a metabolic enzyme or pathway is frequently not very clear, in part because it requires measurements of input and output of the reactions, including flux through the enzyme or pathway. This Update will start out with a short overview on the major PTMs observed for each amino acid residue (PTMs, including determination of the localization within proteins (i.e. the specific residues) and occupancy. Challenges in dealing with multiple PTMs per protein and cross talk between PTMs will be briefly outlined. We then describe the major physiological PTMs observed in plants as well as PTMs that are nonenzymatically induced during sample preparation (PTMs, in particular for enzymes in primary metabolism (Calvin cycle, glycolysis, and respiration) and the C4 shuttle accommodating photosynthesis in C4 plants (PTMs observed in plants
Open in a separate window
Open in a separate window
Open in a separate windowThere are many recent reviews focusing on specific PTMs in plant biology, many of which are cited in this Update. However, the last general review on plant PTMs is from 2010 (Ytterberg and Jensen, 2010); given the enormous progress in PTM research in plants over the last 5 years, a comprehensive overview is overdue. Finally, this Update does not review allosteric regulation by metabolites or other types of metabolic feedback and flux control, even if this is extremely important in the regulation of metabolism and (de)activation of enzymes. Recent reviews for specific pathways, such as isoprenoid metabolism (Kötting et al., 2010; Banerjee and Sharkey, 2014; Rodríguez-Concepción and Boronat, 2015), tetrapyrrole metabolism (Brzezowski et al., 2015), the Calvin-Benson cycle (Michelet et al., 2013), starch metabolism (Kötting et al., 2010; Geigenberger, 2011; Tetlow and Emes, 2014), and photorespiration (Hodges et al., 2013) provide more in-depth discussions of metabolic regulation through various posttranslational mechanisms. Many of the PTMs that have been discovered in the last decade through large-scale proteomics approaches have not yet been integrated in such pathway-specific reviews, because these data are not always easily accessible and because the biological significance of many PTMs is simply not yet understood. We hope that this Update will increase the general awareness of the existence of these PTM data sets, such that their biological significance can be tested and incorporated in metabolic pathways. 相似文献
Amino Acid Residue | Observed Physiological PTM in Plants | PTMs Caused by Sample Preparation |
---|---|---|
Ala (A) | Not known | |
Arg (R) | Methylation, carbonylation | |
Asn (N) | Deamidation, N-linked gycosylation | Deamidation |
Asp (D) | Phosphorylation (in two-component system) | |
Cys (C) | Glutathionylation (SSG), disulfide bonded (S-S), sulfenylation (-SOH), sulfonylation (-SO3H), acylation, lipidation, acetylation, nitrosylation (SNO), methylation, palmitoylation, phosphorylation (rare) | Propionamide |
Glu (E) | Carboxylation, methylation | Pyro-Glu |
Gln (Q) | Deamidation | Deamidation, pyro-Glu |
Gly (G) | N-Myristoylation (N-terminal Gly residue) | |
His (H) | Phosphorylation (infrequent) | Oxidation |
Ile (I) | Not known | |
Leu (L) | Not known | |
Lys (K) | N-ε-Acetylation, methylation, hydroxylation, ubiquitination, sumoylation, deamination, O-glycosylation, carbamylation, carbonylation, formylation | |
Met (M) | (De)formylation, excision (NME), (reversible) oxidation, sulfonation (-SO2), sulfoxation (-SO) | Oxidation, 2-oxidation, formylation, carbamylation |
Phe (F) | Not known | |
Pro (P) | Carbonylation | Oxidation |
Ser (S) | Phosphorylation, O-linked glycosylation, O-linked GlcNAc (O-GlcNAc) | Formylation |
Thr (T) | Phosphorylation, O-linked glycosylation, O-linked GlcNAc (O-GlcNAc), carbonylation | Formylation |
Trp (W) | Glycosylation (C-mannosylation) | Oxidation |
Tyr (Y) | Phosphorylation, nitration | |
Val (V) | Not known | |
Free NH2 of protein N termini | Preprotein processing, Met excision, formylation, pyro-Glu, N-myristoylation, N-acylation (i.e. palmitoylation), N-terminal α-amine acetylation, ubiquitination | Formylation (Met), pyro-Glu (Gln) |
Table II.
Most significant and/or frequent PTMs observed in plantsType of PTM (Reversible, Except if Marked with an Asterisk) | Spontaneous (S; Nonenzymatic) or Enzymatic (E) | Comment on Subcellular Location and Frequency |
---|---|---|
Phosphorylation (Ser, Thr, Tyr, His, Asp) | E | His and Asp phosphorylation have low frequency |
S-Nitrosylation (Cys) and nitration* (Tyr) | S (RNS), but reversal is enzymatic for Cys by thioredoxins | Throughout the cell |
Acetylation (N-terminal α-amine, Lys ε-amine) | E | In mitochondria, very little N-terminal acetylation, but high Lys acetylation; Lys acetylation correlates to [acetyl-CoA] |
Deamidation (Gln, Asn) | S, but reversal of isoAsp is enzymatic by isoAsp methyltransferase | Throughout the cell |
Lipidation (S-acetylation, N-meristoylation*, prenylation*; Cys, Gly, Lys, Trp, N terminal) | E | Not (or rarely) within plastids, mitochondria, peroxisomes |
N-Linked glycosylation (Asp); O linked (Lys, Ser, Thr, Trp) | E | Only proteins passing through the secretory system; O linked in the cell wall |
Ubiquination (Lys, N terminal) | E | Not within plastids, mitochondria, peroxisomes |
Sumoylation (Lys) | E | Not within plastids, mitochondria, peroxisomes |
Carbonylation* (Pro, Lys, Arg, Thr) | S (ROS) | High levels in mitochondria and chloroplast |
Methylation (Arg, Lys, N terminal) | E | Histones (nucleus) and chloroplasts; still underexplored |
Glutathionylation (Cys) | E | High levels in chloroplasts |
Oxidation (Met, Cys) | S (ROS) and E (by PCOs; see Fig. 1B), but reversal is enzymatic by Met sulfoxide reductases, glutaredoxins, and thioredoxins, except if double oxidized | High levels in mitochondria and chloroplast |
Peptidase* (cleavage peptidyl bond) | E | Throughout the cell |
S-Guanylation (Cys) | S (RNS) | Rare; 8-nitro-cGMP is signaling molecule in guard cells |
Formylation (Met) | S, but deformylation is enzymatic by peptide deformylase | All chloroplasts and mitochondria-encoded proteins are synthesized with initiating formylated Met |
Table III.
Regulation by PTMs in plant metabolism and classic examples of well-studied enzymes and pathwaysMany of these enzymes also undergo allosteric regulation through cellular metabolites. GAPDH, Glyceraldehyde-3-phosphate dehydrogenase; PRK, phosphoribulokinase.Process | Enzymes | PTMs, Protein Modifiers, Localization | References |
---|---|---|---|
Calvin-Benson cycle (chloroplasts) | Many enzymes | Oxidoreduction of S-S bonds, reversible nitrosylation, glutathionylation; through ferredoxin/ferredoxin-thioredoxin reductase/thioredoxins (mostly f and m) and glutaredoxins; proteomics studies in Arabidopsis and C. reinhardtii | Michelet et al. (2013) |
Rubisco | Methylation, carbamylation, acetylation, N-terminal processing, oligomerization; classical studies in pea (Pisum sativum), spinach (Spinacia oleracea), and Arabidopsis | Houtz and Portis (2003); Houtz et al. (2008) | |
GAPDH/CP12/PRK supercomplex | Dynamic heterooligomerization through reversible S-S bond formation controlled by thioredoxins | Graciet et al. (2004); Michelet et al. (2013); López-Calcagno et al. (2014) | |
Glycolysis | Cytosolic PEPC | Phosphorylation (S, T), monoubiquitination | O’Leary et al. (2011) |
Photorespiration | Seven enzymes are phosphorylated | Phosphorylation from meta-analysis of public phosphoproteomics data for Arabidopsis; located in chloroplasts, peroxisomes, mitochondria | Hodges et al. (2013) |
Maize glycerate kinase | Redox-regulated S-S bond; thioredoxin f; studied extensively in chloroplasts of C4 maize | Bartsch et al. (2010) | |
Respiration (mitochondria) | Potentially many enzymes, but functional/biochemical consequences are relatively unexplored | Recent studies suggested PTMs for many tricarboxylic acid cycle enzymes, including Lys acetylation and thioredoxin-driven S-S formation; in particular, succinate dehydrogenase and fumarase are inactivated by thioredoxins | Lázaro et al. (2013); Schmidtmann et al. (2014); Daloso et al. (2015) |
PDH | Ser (de)phosphorylation by intrinsic kinase and phosphatase; ammonia and pyruvate control PDH kinase activity; see Figure 1B | Thelen et al. (2000); Tovar-Méndez et al. (2003) | |
C4 cycle (C3 and C4 homologs also involved in glycolysis and/or gluconeogenesis) | Pyruvate orthophosphate dikinase | Phosphorylation by pyruvate orthophosphate dikinase-RP, an S/T bifunctional kinase-phosphatase; in chloroplasts | Chastain et al. (2011); Chen et al. (2014) |
PEPC | Phosphorylation; allosteric regulation by malate and Glc-6-P; in cytosol in mesophyll cells in C4 species (e.g. Panicum maximum); see Figure 1A | Izui et al. (2004); Bailey et al. (2007) | |
PEPC kinase | Ubiquitination resulting in degradation (note also diurnal mRNA levels and linkage to activity level; very low protein level); in cytosol in mesophyll cells in C4 species (e.g. Flaveria spp. and maize) | Agetsuma et al. (2005) | |
PEPC kinase | Phosphorylation in cytosol in bundle sheath cells | Bailey et al. (2007) | |
Starch metabolism (chloroplasts) | ADP-Glc pyrophosphorylase | Redox-regulated disulfide bonds and dynamic oligomerization; thioredoxins; see Figure 1C | Geigenberger et al. (2005); Geigenberger (2011) |
Starch-branching enzyme II | Phosphorylation by Ca2+-dependent protein kinase; P-driven heterooligomerization | Grimaud et al. (2008); Tetlow and Emes (2014) | |
Suc metabolism (cytosol) | SPS (synthesis of Suc) | (De)phosphorylation; SPS kinase and SPS phosphatase; 14-3-3 proteins; cytosol (maize and others) | Huber (2007) |
Suc synthase (breakdown of Suc) | Phosphorylation; Ca2+-dependent protein kinase; correlations to activity, localization, and turnover | Duncan and Huber (2007); Fedosejevs et al. (2014) | |
Photosynthetic electron transport (chloroplast thylakoid membranes) | PSII core and light-harvesting complex proteins | (De)phosphorylation by state-transition kinases (STN7/8) and PP2C phosphatases (PBCP and PPH1/TAP38) | Pesaresi et al. (2011); Tikkanen et al. (2012); Rochaix (2014) |
Nitrogen assimilation | Nitrate reductase | (De)phosphorylation; 14-3-3 proteins | Lillo et al. (2004); Huber (2007) |
20.
《植物生理与分子生物学学报》2014,(6):939-942
Ethylene, a gaseous plant hormone, plays critical roles in plant growth, development, and response to environment. Ethylene-regulated processes are initiated by the elevation of ethylene biosynthesis, which is under tight control by a complex signaling network. An elevated level of ethyl- ene is then perceived by ethylene receptors in local and neighboring cells, which activates signaling pathways that lead to ethylene responses. Different types of tissues/cells have differential capacities in producing ethylene and dif- ferential sensitivity to ethylene, which are crucial to the diverse functions of ethylene in plants. This report high- lights recent advances in our understanding of kinases and phosphatases in ethylene biosynthesis and signaling. 相似文献