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 31 毫秒
1.
2.
Resistance (R) protein–associated pathways are well known to participate in defense against a variety of microbial pathogens. Salicylic acid (SA) and its associated proteinaceous signaling components, including enhanced disease susceptibility 1 (EDS1), non–race-specific disease resistance 1 (NDR1), phytoalexin deficient 4 (PAD4), senescence associated gene 101 (SAG101), and EDS5, have been identified as components of resistance derived from many R proteins. Here, we show that EDS1 and SA fulfill redundant functions in defense signaling mediated by R proteins, which were thought to function independent of EDS1 and/or SA. Simultaneous mutations in EDS1 and the SA–synthesizing enzyme SID2 compromised hypersensitive response and/or resistance mediated by R proteins that contain coiled coil domains at their N-terminal ends. Furthermore, the expression of R genes and the associated defense signaling induced in response to a reduction in the level of oleic acid were also suppressed by compromising SA biosynthesis in the eds1 mutant background. The functional redundancy with SA was specific to EDS1. Results presented here redefine our understanding of the roles of EDS1 and SA in plant defense.  相似文献   

3.
The α-Hairpinins are a family of plant defense peptides with a common fold presenting two short α-helices stabilized by two invariant S–S-bridges. We have shown previously that substitution of just two amino acid residues in a wheat α-hairpinin Tk-AMP-X2 leads to Tk-hefu-2 that features specific affinity to voltage-gated potassium channels KV1.3. Here, we utilize a combined molecular modeling approach based on molecular dynamics simulations and protein surface topography technique to improve the affinity of Tk-hefu-2 to KV1.3 while preserving its specificity. An important advance of this work compared with our previous studies is transition from the analysis of various physicochemical properties of an isolated toxin molecule to its consideration in complex with its target, a membrane-bound ion channel. As a result, a panel of computationally designed Tk-hefu-2 derivatives was synthesized and tested against KV1.3. The most active mutant Tk-hefu-10 showed a half-maximal inhibitory concentration of ∼150 nM being >10 times more active than Tk-hefu-2 and >200 times more active than the original Tk-hefu. We conclude that α-hairpinins provide an attractive disulfide-stabilized scaffold for the rational design of ion channel inhibitors. Furthermore, the success rate can be considerably increased by the proposed “target-based” iterative strategy of molecular design.  相似文献   

4.
5.
This study was to investigate defense mechanisms on cassava induced by salicylic acid formulation (SA) against anthracnose disease. Our results indicated that the SA could reduce anthracnose severity in cassava plants up to 33.3% under the greenhouse condition. The β-1,3-glucanase and chitinase enzyme activities were significantly increased at 24 hours after inoculation (HAI) and decrease at 48 HAI after Colletotrichum gloeosporioides challenge inoculation, respectively, for cassava treated with SA formulation. Synchrotron radiation–based Fourier-transform infrared microspectroscopy spectra revealed changes of the C=H stretching vibration (3,000–2,800 cm−1), pectin (1,740–1,700 cm−1), amide I protein (1,700–1,600 cm−1), amide II protein (1,600–1,500 cm−1), lignin (1,515 cm−1) as well as mainly C–O–C of polysaccharides (1,300–1,100 cm−1) in the leaf epidermal and mesophyll tissues treated with SA formulations, compared to those treated with fungicide carbendazim and distilled water after the challenged inoculation with C. gloeosporioides. The results indicate that biochemical changes in cassava leaf treated with SA played an important role in the enhancement of structural and chemical defense mechanisms leading to reduced anthracnose severity.  相似文献   

6.
Study reveals how a slowly activating calcium channel is able to control rapid excitation–contraction coupling in skeletal muscle.

Skeletal muscle contraction is initiated by action potentials that depolarize the muscle fiber and trigger the rapid release of Ca2+ from the SR via RYR1 channels. This process of excitation–contraction coupling depends on voltage-gated CaV1.1 channels in the plasma membrane, or sarcolemma, of muscle fibers. But CaV1.1 channels are only slowly activated by changes in the sarcolemma membrane potential, and it is therefore unclear how they are able to trigger the much faster activation of RYR1 channels. In this issue of JGP, Savalli et al. reveal that this paradox can be explained by the fact that each of CaV1.1’s four voltage-sensing domains (VSDs) have distinct biophysical properties (1).Nicoletta Savalli (left), Riccardo Olcese (center), and colleagues reveal the distinct physical properties of the CaV1.1 channel’s four voltage-sensing domains (VSD I–IV, right). VSD-I shows slow activation kinetics and is the main contributor to the opening of CaV1.1. The other VSDs activate much faster and may therefore be coupled to RYR1 to mediate the rapid release of Ca2+ from the SR during skeletal muscle contraction.RYR1 channels have no voltage-sensing machinery of their own and therefore rely on a physical connection to CaV1.1 channels to release Ca2+ and initiate muscle contraction in response to muscle fiber depolarization. But RYR1 channels open ∼25 times faster than CaV1.1 channels. “So, how can these slowly activating CaV1.1 channels trigger the rapid release of Ca2+ from the SR?” asks Riccardo Olcese, a professor at the David Geffen School of Medicine, UCLA.Olcese and colleagues, including Assistant Project Scientist Nicoletta Savalli, suspected that the answer might lie in the fact that, like many other voltage-gated ion channels, CaV1.1 has four VSDs that alter their conformation in response to voltage changes. These domains are similar, but not identical, to each other, potentially enabling them to have distinct biophysical properties and perform distinct functions. Indeed, Olcese and colleagues previously demonstrated that, in the closely related channel CaV1.2, only VSDs II and III are involved in pore opening (2, 3).Savalli et al. used voltage-clamp fluorometry to compare the properties of CaV1.1’s VSDs, expressing the channel in Xenopus oocytes and labeling each of its VSDs in turn with an environmentally sensitive fluorophore to report voltage-dependent changes in their conformation (1). “We found that the four VSDs were very heterogenous in both their kinetics and voltage dependencies,” says Olcese. “VSD-I had very slow kinetics, compatible with the slow activation of the CaV1.1 pore. The other three VSDs had much faster kinetics and could, therefore, be good candidates to be the voltage sensors for RYR1 activation.”Olcese and colleagues confirmed the importance of VSD-I for CaV1.1 activation by analyzing a naturally occurring, charge-neutralizing mutation in this domain, R174W, that is linked to malignant hyperthermia (4). The team found that this mutation reduced the voltage-sensitivity of VSD-I and abolished the ability of CaV1.1 to conduct Ca2+ at physiological membrane potentials, but had no effect on the behavior of the other three VSDs.Finally, Savalli et al. applied their data on both the wild-type and mutant VSDs to an allosteric model of CaV activation (2, 3), which predicted that VSD-I contributes most of the energy required to stabilize the open state of CaV1.1, while the other VSDs contribute little to nothing.Thus, CaV1.1 activation is mainly driven by a single VSD—a mechanism that hasn’t been seen in any other voltage-gated ion channel—leaving the other VSDs free to perform other functions, such as the rapid activation of RYR1. Olcese and colleagues now want to pinpoint exactly which VSD(s) are coupled to RYR1 and determine how they trigger rapid Ca2+ release from the SR.  相似文献   

7.
In an effort to engineer countermeasures for the category B toxin ricin, we produced and characterized a collection of epitopic tagged, heavy chain-only antibody VH domains (VHHs) specific for the ricin enzymatic (RTA) and binding (RTB) subunits. Among the 20 unique ricin-specific VHHs we identified, six had toxin-neutralizing activity: five specific for RTA and one specific for RTB. Three neutralizing RTA-specific VHHs were each linked via a short peptide spacer to the sole neutralizing anti-RTB VHH to create VHH “heterodimers.” As compared with equimolar concentrations of their respective monovalent monomers, all three VHH heterodimers had higher affinities for ricin and, in the case of heterodimer D10/B7, a 6-fold increase in in vitro toxin-neutralizing activity. When passively administered to mice at a 4:1 heterodimer:toxin ratio, D10/B7 conferred 100% survival in response to a 10 × LD50 ricin challenge, whereas a 2:1 heterodimer:toxin ratio conferred 20% survival. However, complete survival was achievable when the low dose of D10/B7 was combined with an IgG1 anti-epitopic tag monoclonal antibody, possibly because decorating the toxin with up to four IgGs promoted serum clearance. The two additional ricin-specific heterodimers, when tested in vivo, provided equal or greater passive protection than D10/B7, thereby warranting further investigation of all three heterodimers as possible therapeutics.  相似文献   

8.
V-ATPase (VoV1) converts the chemical free energy of ATP into an ion-motive force across the cell membrane via mechanical rotation. This energy conversion requires proper interactions between the rotor and stator in VoV1 for tight coupling among chemical reaction, torque generation, and ion transport. We developed an Escherichia coli expression system for Enterococcus hirae VoV1 (EhVoV1) and established a single-molecule rotation assay to measure the torque generated. Recombinant and native EhVoV1 exhibited almost identical dependence of ATP hydrolysis activity on sodium ion and ATP concentrations, indicating their functional equivalence. In a single-molecule rotation assay with a low load probe at high ATP concentration, EhVoV1 only showed the “clear” state without apparent backward steps, whereas EhV1 showed two states, “clear” and “unclear.” Furthermore, EhVoV1 showed slower rotation than EhV1 without the three distinct pauses separated by 120° that were observed in EhV1. When using a large probe, EhVoV1 showed faster rotation than EhV1, and the torque of EhVoV1 estimated from the continuous rotation was nearly double that of EhV1. On the other hand, stepping torque of EhV1 in the clear state was comparable with that of EhVoV1. These results indicate that rotor-stator interactions of the Vo moiety and/or sodium ion transport limit the rotation driven by the V1 moiety, and the rotor-stator interactions in EhVoV1 are stabilized by two peripheral stalks to generate a larger torque than that of isolated EhV1. However, the torque value was substantially lower than that of other rotary ATPases, implying the low energy conversion efficiency of EhVoV1.  相似文献   

9.
D222G/N substitutions in A(H1N1)pdm09 hemagglutinin may be associated with increased binding of viruses causing low respiratory tract infections and human pathogenesis. We assessed the impact of such substitutions on the balance between hemagglutinin binding and neuraminidase cleavage, viral growth and in vivo virulence.Seven viruses with differing polymorphisms at codon 222 (2 with D, 3 G, 1 N and 1 E) were isolated from patients and characterized with regards hemagglutinin binding affinity (Kd) to α-2,6 sialic acid (SAα-2,6) and SAα-2,3 and neuraminidase enzymatic properties (Km, Ki and Vmax). The hemagglutination assay was used to quantitatively assess the balance between hemagglutinin binding and neuraminidase cleavage. Viral growth properties were compared in vitro in MDCK-SIAT1 cells and in vivo in BALB/c mice. Compared with D222 variants, the binding affinity of G222 variants was greater for SAα-2,3 and lower for SAα-2,6, whereas that of both E222 and N222 variants was greater for both SAα-2,3 and SAα-2,6. Mean neuraminidase activity of D222 variants (16.0 nmol/h/106) was higher than that of G222 (1.7 nmol/h/106 viruses) and E/N222 variants (4.4 nmol/h/106 viruses). The hemagglutination assay demonstrated a deviation from functional balance by E222 and N222 variants that displayed strong hemagglutinin binding but weak neuraminidase activity. This deviation impaired viral growth in MDCK-SIAT1 cells but not infectivity in mice. All strains but one exhibited low infectious dose in mice (MID50) and replicated to high titers in the lung; this D222 strain exhibited a ten-fold higher MID50 and replicated to low titers. Hemagglutinin-neuraminidase balance status had a greater impact on viral replication than hemagglutinin affinity strength, at least in vitro, thus emphasizing the importance of an optimal balance for influenza virus fitness. The mouse model is effective in assessing binding to SAα-2,3 but cannot differentiate SAα-2,3- from SAα-2,6- preference, nor estimate the hemagglutinin-neuraminidase balance in A(H1N1)pdm09 strains.  相似文献   

10.
Small recombinant antibody fragments (e.g. scFvs and VHHs), which are highly tissue permeable, are being investigated for antivenom production as conventional antivenoms consisting of IgG or F(ab’)2 antibody fragments do not effectively neutralize venom toxins located in deep tissues. However, antivenoms composed entirely of small antibody fragments may have poor therapeutic efficacy due to their short serum half-lives. To increase serum persistence and maintain tissue penetration, we prepared low and high molecular mass antivenom antibodies. Four llama VHHs were isolated from an immune VHH-displayed phage library and were shown to have high affinity, in the low nM range, for α-cobratoxin (α–Cbtx), the most lethal component of Naja kaouthia venom. Subsequently, our highest affinity VHH (C2) was fused to a human Fc fragment to create a VHH2-Fc antibody that would offer prolonged serum persistence. After in planta (Nicotiana benthamiana) expression and purification, we show that our VHH2-Fc antibody retained high affinity binding to α–Cbtx. Mouse α–Cbtx challenge studies showed that our highest affinity VHHs (C2 and C20) and the VHH2-Fc antibody effectively neutralized lethality induced by α–Cbtx at an antibody:toxin molar ratio as low as ca. 0.75×:1. Further research towards the development of an antivenom therapeutic involving these anti-α-Cbtx VHHs and VHH2-Fc antibody molecules should involve testing them as a combination, to determine whether they maintain tissue penetration capability and low immunogenicity, and whether they exhibit improved serum persistence and therapeutic efficacy.  相似文献   

11.
Alzheimer''s disease (AD) pathology is characterized by loss of memory cognitive and behavioral deterioration. One of the hallmarks of AD is amyloid β (Aβ) plaques in the brain that consists of Aβ oligomers and fibrils. It is accepted that oligomers, particularly dimers, are toxic species that are produced extracellularly and intracellularly in membranes. It is believed that the disruption of membranes by polymorphic Aβ oligomers is the key for the pathology of AD. This is a first study that investigate the effect of polymorphic “α‐helix/random coil” and “fibril‐like” Aβ dimers on 1,2‐dioleoyl‐sn‐glycero‐3‐phosphocholine (DOPC) membrane. It has been found that the DOPC membrane promotes Aβ1–42 “fibril‐like” dimers and impedes Aβ1–42 “α‐helix/random coil” dimers. The N‐termini domains within Aβ1–42 dimers play a role in Aβ aggregation in membrane milieus. In addition, the aromatic π–π interactions (involving residues F19 and F20 in Aβ1–42) are the driving forces for the hydrophobic interactions that initiate the primary nucleation of polymorphic Aβ1–42 dimers within DOPC membrane. Finally, the DOPC bilayer membrane thickness is locally decreased, and it is disrupted by an embedded distinct Aβ1–42 dimer, due to relatively large contacts between Aβ1–42 monomers and the DOPC membrane. This study reveals insights into the molecular mechanisms by which polymorphic early‐stage Aβ1–42 dimers have distinct impacts on DOPC membrane.  相似文献   

12.
CoxD of the α-proteobacterium Oligotropha carboxidovorans is a membrane protein which is involved in the posttranslational biosynthesis of the [CuSMoO2] cluster in the active site of the enzyme CO dehydrogenase. The bacteria synthesize CoxD only in the presence of CO. Recombinant CoxD produced in E. coli K38 pGP1-2/pETMW2 appeared in inclusion bodies from where it was solubilized by urea and refolded by stepwise dilution. Circular dichroism spectroscopy revealed the presence of secondary structural elements in refolded CoxD. CoxD is a P-loop ATPase of the AAA-protein family. Refolded CoxD catalyzed the hydrolysis of MgATP yielding MgADP and inorganic phosphate at a 1∶1∶1 molar ratio. The reaction was inhibited by the slow hydrolysable MgATP-γ-S. GTPase activity of CoxD did not exceed 2% of the ATPase activity. Employing different methods (non linear regression, Hanes and Woolf, Lineweaver-Burk), preparations of CoxD revealed a mean KM value of 0.69±0.14 mM ATP and an apparent Vmax value of 19.3±2.3 nmol ATP hydrolyzed min−1 mg−1. Sucrose density gradient centrifugation and gel filtration showed that refolded CoxD can exist in various multimeric states (2-mer, 4-mer or 6-mer), preferentially as hexamer or dimer. Within weeks the hexamer dissociates into the dimer, a process which can be reversed by MgATP or MgATP-γ-S within hours. Only the hexamers and the dimers exhibited MgATPase activity. Transmission electron microscopy of negatively stained CoxD preparations revealed distinct particles within a size range of 10–16 nm, which further corroborates the oligomeric organization. The 3D structure of CoxD was modeled with the 3D structure of BchI from Rhodobacter capsulatus as template. It has the key elements of an AAA+ domain in the same arrangement and at same positions as in BchI and displays the characteristic inserts of the PS-II-insert clade. Possible functions of CoxD in [CuSMoO2] cluster assembly are discussed.  相似文献   

13.
Experimental and theoretical evidence is presented that supports the theory that the intramolecular charge transfer (ICT) state of peridinin is an evolved state formed via excited-state bond-order reversal and solvent reorganization in polar media. The ICT state evolves in <100 fs and is characterized by a large dipole moment (∼35 D). The charge transfer character involves a shift of electron density within the polyene chain, and it does not involve participation of molecular orbitals localized in either of the β-rings. Charge is moved from the allenic side of the polyene into the furanic ring region and is accompanied by bond-order reversal in the central portion of the polyene chain. The electronic properties of the ICT state are generated via mixing of the “11Bu+” ionic state and the lowest-lying “21Ag” covalent state. The resulting ICT state is primarily 1Bu+-like in character and exhibits not only a large oscillator strength but an unusually large doubly excited character. In most solvents, two populations exist in equilibrium, one with a lowest-lying ICT ionic state and a second with a lowest-lying “21Ag” covalent state. The two populations are separated by a small barrier associated with solvent relaxation and cavity formation.  相似文献   

14.
During ricin intoxication in mammalian cells, ricin''s enzymatic (RTA) and binding (RTB) subunits disassociate in the endoplasmic reticulum. RTA is then translocated into the cytoplasm where, by virtue of its ability to depurinate a conserved residue within the sarcin–ricin loop (SRL) of 28S rRNA, it functions as a ribosome-inactivating protein. It has been proposed that recruitment of RTA to the SRL is facilitated by ribosomal P-stalk proteins, whose C-terminal domains interact with a cavity on RTA normally masked by RTB; however, evidence that this interaction is critical for RTA activity within cells is lacking. Here, we characterized a collection of single-domain antibodies (VHHs) whose epitopes overlap with the P-stalk binding pocket on RTA. The crystal structures of three such VHHs (V9E1, V9F9, and V9B2) in complex with RTA revealed not only occlusion of the ribosomal P-stalk binding pocket but also structural mimicry of C-terminal domain peptides by complementarity-determining region 3. In vitro assays confirmed that these VHHs block RTA–P-stalk peptide interactions and protect ribosomes from depurination. Moreover, when expressed as “intrabodies,” these VHHs rendered cells resistant to ricin intoxication. One VHH (V9F6), whose epitope was structurally determined to be immediately adjacent to the P-stalk binding pocket, was unable to neutralize ricin within cells or protect ribosomes from RTA in vitro. These findings are consistent with the recruitment of RTA to the SRL by ribosomal P-stalk proteins as a requisite event in ricin-induced ribosome inactivation.  相似文献   

15.
Point mutations in pore-lining S6 segments of CaV1.2 shift the voltage dependence of activation into the hyperpolarizing direction and significantly decelerate current activation and deactivation. Here, we analyze theses changes in channel gating in terms of a circular four-state model accounting for an activation R–A–O and a deactivation O–D–R pathway. Transitions between resting-closed (R) and activated-closed (A) states (rate constants x(V) and y(V)) and open (O) and deactivated-open (D) states (u(V) and w(V)) describe voltage-dependent sensor movements. Voltage-independent pore openings and closures during activation (A–O) and deactivation (D–R) are described by rate constants α and β, and γ and δ, respectively. Rate constants were determined for 16-channel constructs assuming that pore mutations in IIS6 do not affect the activating transition of the voltage-sensing machinery (x(V) and y(V)). Estimated model parameters of 15 CaV1.2 constructs well describe the activation and deactivation processes. Voltage dependence of the “pore-releasing” sensor movement ((x(V)) was much weaker than the voltage dependence of “pore-locking” sensor movement (y(V)). Our data suggest that changes in membrane voltage are more efficient in closing than in opening CaV1.2. The model failed to reproduce current kinetics of mutation A780P that was, however, accurately fitted with individually adjusted x(V) and y(V). We speculate that structural changes induced by a proline substitution in this position may disturb the voltage-sensing domain.  相似文献   

16.
We have recently improved the automation of an in-gel digestion system, DigestPro 96, using in situ alkylation of proteins with acrylamide, conducted during one-dimensional (1D) SDS-PAGE. The improved method included the processes of destaining, dehydration, trypsin digestion, and extraction but excluded the reduction and alkylation steps following staining of proteins with CBB. The extracted peptide mixtures were directly loaded onto a micro C18 LC column of the mass spectrometer. The resultant spectra were processed with “Mascot” search engine to estimate the sequence coverage of the bovine serum albumin (BSA). The original method, designed for Laemmli 1D SDS gel applications, consisted of reduction and post-alkylation with iodoacetamide, which produced carboxyamidemethyl (CAM; –S–CH2CONH2) derivatives. The original method also included a desalting step essential for mass spectrometry, especially matrix-assisted laser desorption/ionization time-of-flight mass spectrometry. We compared the original and improved methods using BSA (3 pmol loaded to the gel, one third of digested peptide mixture injected into LC-MS). The original method yielded both CAM and propionicamide (PAM; –S–CH2CH2CONH2) derivatives. The source of PAM derivatives is the unpolymerized acrylamide formed during electrophoresis. The sequence coverage of CAM derivatives of BSA by the original method was 10% with desalting and 19% without desalting. The sequence coverage of PAM derivative by the improved method was 32%. Our results clearly show the advantage of our improved automated in-gel digestion method for in situ PAM alkylated protein with respect to peptide recovery, compared with the original method with CAM post-alkylation.  相似文献   

17.
Dilated cardiomyopathy is a disease of left ventricular dysfunction accompanied by impairment of the β1-adrenergic receptor (β1-AR) signal cascade. The disturbed β1-AR function may be based on an elevated sympathetic tone observed in patients with heart failure. Prolonged adrenergic stimulation may induce metabolic and electrophysiological disturbances in the myocardium, resulting in tachyarrhythmia that leads to the development of heart failure in human and sudden death. Hence, β1-AR is considered as a promising drug target but attempts to develop effective and specific drug against this tempting pharmaceutical target is slowed down due to the lack of 3D structure of Homo sapiens β1-AR (hsβADR1). This study encompasses elucidation of 3D structural and physicochemical properties of hsβADR1 via threading-based homology modeling. Furthermore, the docking performance of several docking programs including Surflex-Dock, FRED, and GOLD were validated by re-docking and cross-docking experiments. GOLD and Surflex-Dock performed best in re-docking and cross docking experiments, respectively. Consequently, Surflex-Dock was used to predict the binding modes of four hsβADR1 agonists. This study provides clear understanding of hsβADR1 structure and its binding mechanism, thus help in providing the remedial solutions of cardiovascular, effective treatment of asthma and other diseases caused by malfunctioning of the target protein.  相似文献   

18.
Auxiliary Ca2+ channel β subunits (CaVβ) regulate cellular Ca2+ signaling by trafficking pore-forming α1 subunits to the membrane and normalizing channel gating. These effects are mediated through a characteristic src homology 3/guanylate kinase (SH3–GK) structural module, a design feature shared in common with the membrane-associated guanylate kinase (MAGUK) family of scaffold proteins. However, the mechanisms by which the CaVβ SH3–GK module regulates multiple Ca2+ channel functions are not well understood. Here, using a split-domain approach, we investigated the role of the interrelationship between CaVβ SH3 and GK domains in defining channel properties. The studies build upon a previously identified split-domain pair that displays a trans SH3–GK interaction, and fully reconstitutes CaVβ effects on channel trafficking, activation gating, and increased open probability (Po). Here, by varying the precise locations used to separate SH3 and GK domains and monitoring subsequent SH3–GK interactions by fluorescence resonance energy transfer (FRET), we identified a particular split-domain pair that displayed a subtly altered configuration of the trans SH3–GK interaction. Remarkably, this pair discriminated between CaVβ trafficking and gating properties: α1C targeting to the membrane was fully reconstituted, whereas shifts in activation gating and increased Po functions were selectively lost. A more extreme case, in which the trans SH3–GK interaction was selectively ablated, yielded a split-domain pair that could reconstitute neither the trafficking nor gating-modulation functions, even though both moieties could independently engage their respective binding sites on the α1C (CaV1.2) subunit. The results reveal that CaVβ SH3 and GK domains function codependently to tune Ca2+ channel trafficking and gating properties, and suggest new paradigms for physiological and therapeutic regulation of Ca2+ channel activity.  相似文献   

19.
Part of the “signature sequence” that defines the voltage-gated proton channel (HV1) is a tryptophan residue adjacent to the second Arg in the S4 transmembrane helix: RxWRxxR, which is perfectly conserved in all high confidence HV1 genes. Replacing Trp207 in human HV1 (hHV1) with Ala, Ser, or Phe facilitated gating, accelerating channel opening by 100-fold, and closing by 30-fold. Mutant channels opened at more negative voltages than wild-type (WT) channels, indicating that in WT channels, Trp favors a closed state. The Arrhenius activation energy, Ea, for channel opening decreased to 22 kcal/mol from 30–38 kcal/mol for WT, confirming that Trp207 establishes the major energy barrier between closed and open hHV1. Cation–π interaction between Trp207 and Arg211 evidently latches the channel closed. Trp207 mutants lost proton selectivity at pHo >8.0. Finally, gating that depends on the transmembrane pH gradient (ΔpH-dependent gating), a universal feature of HV1 that is essential to its biological functions, was compromised. In the WT hHV1, ΔpH-dependent gating is shown to saturate above pHi or pHo 8, consistent with a single pH sensor with alternating access to internal and external solutions. However, saturation occurred independently of ΔpH, indicating the existence of distinct internal and external pH sensors. In Trp207 mutants, ΔpH-dependent gating saturated at lower pHo but not at lower pHi. That Trp207 mutation selectively alters pHo sensing further supports the existence of distinct internal and external pH sensors. Analogous mutations in HV1 from the unicellular species Karlodinium veneficum and Emiliania huxleyi produced generally similar consequences. Saturation of ΔpH-dependent gating occurred at the same pHo and pHi in HV1 of all three species, suggesting that the same or similar group(s) is involved in pH sensing. Therefore, Trp enables four characteristic properties: slow channel opening, highly temperature-dependent gating kinetics, proton selectivity, and ΔpH-dependent gating.  相似文献   

20.
The hypersensitive response (HR) is a cell death phenomenon associated with localized resistance to pathogens. Biphasic patterns in the generation of H2O2, salicylic acid and ethylene have been observed in tobacco during the early stages of the HR. These biphasic models reflect an initial elicitation by pathogen-associated molecular patterns followed by a second phase, induced by pathogen-encoded avirulence gene products. The first phase has been proposed to potentiate the second, to increase the efficacy of plant resistance to disease. This potentiation is comparable to the “priming” of plant defenses which is seen when plants display systemic resistance to disease. The events regulating the generation of the biphasic wave, or priming, remains obscure, however recently we demonstrated a key role for nitric oxide in this process in a HR occurring in tobacco. Here we use laser photoacoustic detection to demonstrate that biphasic ethylene production also occurs during a HR occurring in Arabidopsis. We suggest that ethylene emanation during the HR represents a ready means of visualising biphasic events during the HR and that exploiting the genomic resources offered by this model species will facilitate the development of a mechanistic understanding of potentiating/priming processes.Key words: hypersensitive response, biphasic patterns, potentiation, defense priming, ethylene, ArabidopsisThe Hypersensitive Response (HR) is a cell death process which occurs at the site of attempted pathogen attack and which has been associated with host resistance.1 Much work on the regulation of the HR has indicated the importance of H2O2,2 and NO.3 A feature of H2O2 generation during the HR is its biphasic pattern (Fig. 1A). The first rise reflects elicitation by pathogen-associated molecular patterns (PAMPs)4 and the second reflects the interaction between a pathogen-encoded avirulence (avr) gene product with a plant resistance (R) gene. A key aspect of the first rise is the initiation of salicylic acid (SA) synthesis which potentiates the second rise and hence the potency of plant defense and the HR.5Open in a separate windowFigure 1Patterns of defense signal generation during the Pseudomonas syringae pv. phaseolicola elicited-hypersensitive response in tobacco (Nicotiana tabacum). Generation of (A) H2O2 (●, Mur18); (B) nitric oxide (◇; Mur12 (C) salicylic acid (SA, ■19) and (D) ethylene (○ Mur9) during a HR elicited by Pseudomonas syringae pv. phaseolicola (Psph) in tobacco cv. Samsun NN. In (A) a phase where SA acts to augment the second rise in H2O2—the potentiation phase—is highlighted. The potentiation phase is likely to be similar to defense “priming”.6 Methodological details are contained within the appropriate references. (E) A possible model for biphasic defense signal regulation during the Psph-elicited HR in tobacco. During an initial phase NO and H2O2 act to initiate SA biosynthesis, where SA and NO act to initiate a “H2O2 biphasic switch”. This could initially suppress both SA and the H2O2 generation but subsequently acts to potentiate a second phase of H2O2 generation. This in turn increases SA biosynthesis which could act with NO to initiate the “C2H4 biphasic switch” to potentiate ethylene production. These (and other) signals contribute to initiation of the HR and SAR.This potentiation mechanism appears to be similar to defense priming; when whole plants display systemic resistance to disease as opposed to a localized resistance against pathogens. Priming can be initiated (the “primary stimulus”) following attack with a necrotizing pathogen (leading to “systemic acquired resistance”, SAR) or non-pathogenic rhizosphere bacteria (to confer “induced systemic resistance”, ISR). In the primed state a plant stimulates a range of plant defense genes, produces anti-microbial phytoalexins and deposits cell wall strengthening molecules, but only on imposition of a “secondary stimulus”.6 Such secondary stimuli include SA3 or PAMPs7 and is likely to be mechanistically similar to the potentiation step in the biphasic pattern of H2O2 generation (shaded in Fig. 1A). Accordingly, the two phases in the biphasic wave represent primary and secondary stimuli in priming.Highlighting a similarity between local HR-based events and priming, adds further impetus to efforts aiming to describe the underlying mechanism(s), however both phenomena remain poorly understood. Besides SA, both jasmonates and abscisic acid (ABA) have been shown to prime defenses as have a range of non-plant chemicals, with β-aminobutyric acid (BABA) being perhaps most widely used.6,8 Mutants which fail to exhibit BABA-mediated potentiation were defective in either a cyclin-dependent kinase-like protein, a polyphosphoinositide phosphatase or an ABA biosynthetic enzyme.8We have recently investigated biphasic ethylene production during the HR in tobacco elicited by the nonhost HR-eliciting bacterial pathogen Pseudomonas syringae pv. phaseolicola.9 As with H2O2 generation, this pattern reflected PAMP-and AVR-dependent elicitation events and included a SA-mediated potentiation stage. Crucially, we also showed that NO was a vital component in the SA-potentiation mechanism. When this finding is integrated with our other measurements of defense signal generation in the same host-pathogen system the complexity in the signaling network is revealed (Fig. 1). NO generation (Fig. 1B) appeared to be coincident with the first rise in H2O2 (Fig. 1A) which initiated SA biosynthesis10,11 and together would contribute to the first small, but transient, rise in that hormone (Fig. 1C). In line with established models5 this momentary rise in SA coincides with the potentiation phase (shaded in Fig. 1A) required to augment the second rise in ROS. However, ethylene production seems to be correlated poorly with the patterns of NO, H2O2 and SA (Fig. 1D). Nevertheless, biphasic ethylene production was found to reflect PAMP and AVR-dependent recognition and included a SA-mediated potentiation step.9 Hence, ethylene production could be used as a post-hoc indicator of the potentiation mechanism. Therefore, our discovery that the second wave of ethylene production—a “biphasic switch”—is influenced by NO acting with SA could also be relevant to the H2O2 generation. Significantly, the second phases in both H2O2 and ethylene production occur exactly where SA and NO production coincides; in the case of H2O2 generation 2–4 h post challenge and with ethylene 6 h onwards (Fig. 1E).Thus, ethylene production represents a readily assayable marker to indicate perturbations in the underlying biphasic and possible priming mechanisms. As we have demonstrated, laser photoacoustic detection (LAPD) is a powerful on-line approach to determine in planta ethylene production in tobacco9,12 but any mechanistic investigations would be greatly facilitated if the genetic resources offered by the model species Arabidopsis could be exploited.To address this, Arabidopsis Col-0 rosettes were vacuum infiltrated with either Pseudomonas syringae pv. tomato (Pst) avrRpm1 (HR-eliciting), the virulent Pst strain and the non-HR eliciting and non-virulent Pst hrpA strain. Ethylene production was monitored by LAPD (Fig. 2A). Significantly, Pst avrRpm1 initiated a biphasic pattern of ethylene production whose kinetics were very similar to that seen in tobacco (compare Figs. 2A with with1D).1D). Inoculations with Pst and Pst hrpA only displayed the first PAMP-dependent rise in ethylene production. Thus, these data establish that Arabidopsis can be used to investigate biphasic switch mechanism(s) in ethylene production during the HR and possibly defense priming. When considering such mechanisms, it is relevant to highlight the work of Foschi et al.13 who observed that biphasic activation of a monomeric G protein to cause phase-specific activation of different kinase cascades. Interestingly, ethylene has been noted to initiate biphasic activation of G proteins and kinases in Arabidopsis, although differing in kinetics to the phases seen during the HR.14 Further, plant defense priming has been associated with the increased accumulation of MAP kinase protein.6Open in a separate windowFigure 2Ethylene in the Pseudomonas syringae pv. tomato elicited-hypersensitive response in Arabidopsis thaliana. (A) Ethylene production from 5 week old short day (8 h light 100 µmol.m2.sec−1) grown Arabidopsis rosette leaves which were vacuum infiltrated with bacterial suspensions (2 × 106 colony forming units.ml−1) of Pseudomonas syringae pv. tomato (Pst) strains detected using laser photoacoustic detection (LAPD). Experimental details of the ethylene detection by LAPD are detailed in Mur et al.9 The intercellular spaces in leaves were infiltrated with the HR-eliciting strain Pst avrRpm1, (■), the virulent strain Pst (△) or the non-virulent and non-HR eliciting derivative, Pst hrpA (◇). (B) The appearance of Arabidopsis Col-0 and etr1-1 leaves at various h following injection with 2 × 106 c.f.u.mL−1 with of Pst avrRpm1. (C) Explants (1 cm diameter discs) from Arabidopsis leaf areas infiltrated with suspensions of Pst avrRpm1 were placed in a 1.5 cm diameter well, bathed in 1 mL de-ionized H2O. Changes in the conductivity of the bathing solution, as an indicator of electrolyte leakage from either wild type Col-0 (◆), mutants which were compromised in ethylene signaling; etr1-1 (□), ein2-2 (▲) or which overproduced ethylene; eto2-1 (●) were measured using a conductivity meter. Methodological details are set out in Mur et al.9A further point requires consideration; the role of ethylene as a direct contributor to plant defense.15 The contribution of ethylene to the HR has been disputed,16 but in tobacco we have observed that altered ethylene production influenced the formation of a P. syringae pv. phaseolicola elicited HR.9 In Arabidopsis, cell death in the ethylene receptor mutant etr1-1 following inoculation with Pst avrRpm1 is delayed compared to wild type (Fig. 2B). When electrolyte leakage was used to quantify Pst avrRpm1 cell death, both etr1-1 and the ethylene insensitive signaling mutant ein2-1 exhibited slower death than wild-type but in the ethylene overproducing mutant eto2, cell death was augmented (Fig. 2C). These data indicate that ethylene influences the kinetics of the HR.Taking these data together we suggest that the complexity of signal interaction during the HR or in SAR/ISR could be further dissected by combining the genetic resources of Arabidopsis with measurements of ethylene production using such sensitive approaches as LAPD.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号