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 62 毫秒
1.
The reaction of [Ru(CO)2Cl2]n with bis(2-pyridylmethyl)amine (bpma) in refluxing ethanol followed by anion exchange yields two products: cis,fac-[Ru(bpma)(CO)2Cl]PF6 (1a, 71%) and trans,fac-[Ru(bpma)(CO)2Cl]PF6 (1b, 29%). Reaction of 1a with AgBF4 in acetone, followed by acetonitrile and then anion exchange gave cis,fac-[Ru(bpma)(CO)2(CH3CN)](PF6)2 (2a). In the same way, 1b afforded trans,fac-[Ru(bpma)(CO)2(CH3CN)](PF6)2 (2b). Reaction of depolymerized [Ru(CO)2Cl2]n with bpma in ethanol at room temperature afforded cis,cis-[Ru(η2-bpma)(CO)2Cl2] (3). In refluxing ethanol, 3 was converted to cis,fac-[Ru(bpma)(CO)2Cl]Cl (1a-Cl). Heating 3 in chlorobenzene afforded 1b-Cl, exclusively; heating 3 in ethylene glycol gave mainly 1a-Cl. Heating 1a-Cl in ethanol resulted in no isomerization, but heating in chlorobenzene gave a mixture of 3 and 1b-Cl. Anion exchange for PF6 with 1a-Cl and 1b-Cl afforded 1a and 1b, respectively, whereas anion exchange for BPh4 afforded 1a-BPh4. Compounds 1a, 1b, 2a and 3 have been structurally characterized.  相似文献   

2.
In this study, frequencies of the polymorphic variants of the genes encoding antioxidant enzymes, GSTM1, GSTT1, GSTP1, CAT, GPX1, NQO1, SOD1, and SOD3 were examined in three ethnic groups of healthy subjects from the Republic of Bashkortostan (Russians, Tatars, and Bashkirs). An association of these markers with the development of chronic obstructive pulmonary disease (COPD) was tested. Interethnic differences relative to the distribution of the polymorphic variants of the GSTP1 locus Ile105Val and the NQO1 locus 609C/T were revealed. Relative to the genotype distribution at the Ile105Val locus of the GSTP1 gene, ethnic group of Bashkirs was found to be statistically significantly different from Tatars (χ2 = 8.819; d.f. = 2; P = 0.012). Relative to the genotype frequency distribution pattern at the NQO1 locus 609C/T, the group of Bashkirs differed from Russians (χ2 = 8.913; d.f. = 2; P = 0.012). An association of genotype Val/Val of the GSTP1 Ile105Val locus with the risk of COPD in Russians (χ2 = 5.25; P = 0.022; P cor = 0.044; OR = 4.09), and of the GSTP1 haplotype *D in Tatars, was demonstrated (χ2 = 11.575; P = 0.0014; P cor = 0.0042; OR = 3.178). Genotype TT of the CAT ?262C/T locus marked resistance to the COPD development in Russians (χ2 = 6.82; P = 0.0098; P cor = 0.0196; OR = 0.31; 95%CI, 0.119 to 0.77). The risk for COPD in the ethnic group of Tatars was associated with the CAT haplotype (?262)C/(1167)T (χ2 = 6.038; P = 0.0147; P cor = 0.044; OR = 1.71). Analysis of the NQO1 haplotypes at the 465C/T and 609C/T loci showed that haplotype 465C/609T was associated with COPD in Russians (χ2 = 4.571; P = 0.0328; P cor = 0.01; OR = 1.799). It was demonstrated that Gly allele of the Arg213Gly polymorphic locus of the SOD3 gene marked the risk for COPD in the ethnic group of Tatars (OR = 2.23; 95%CI, 1.22 to 4.1). Thus, GSTP1, CAT, NQO1, and SOD3 polymorphisms play an important role in the development of COPD among the population of Bashkortostan.  相似文献   

3.
There are two close empirical scalings, namely, the T-11 and neo-Alcator ones, that provide correct estimates for the energy confinement time in tokamaks in ohmic heating regimes in the linear part of the dependence τ E (\(\bar n_e \)) in the range of low values of \(\bar n_e \) and 〈ν e * 〉 ≤ 1. The similar character of electron energy confinement in this range, which expands with increasing magnetic field B 0, has stimulated the search for dimensionless parameters and simple physical models that would explain the experimentally observed dependences χ e ~ 1/n e and τ Ee \(\bar n_e \). In 1987, T. Okhawa showed that the experimental data were satisfactorily described by the formula χe = (c 2 pe 2 )ν e /qR, in deriving of which the random spatial leap along the radius r on the electron trajectory was assumed to be the same as that in the coefficient of the poloidal field diffusion, while the repetition rate of these leaps was assumed to be ν e /qR. In 2004, J. Callen took into account the decrease in the fraction of transient electrons with increasing toroidal ratio ? = r/R and corrected the coefficient c 2 pe 2 in Okhawa equation by the factor σ Sp neo . If one takes into account this correction and assumes that the frequency of the stochastic process is equal to the reciprocal of the half-period of rotation of a trapped electron along its banana trajectory, then the resulting expression for χe will coincide with the T-11 scaling: χ e an ∞ ?1.75(T e /A i )0.5/(n e qR) at A i = 1. If the same stochastic process also involves ions, it may result in the opening of the orbit of a trapped ion at the distance ~(c pe )(m i /m e )1/4. In this case, the calculated coefficient of electron and ion diffusion D is close to D an ≈ χ e an /2.  相似文献   

4.

Background

The methanol-regulated AOX1 promoter (PAOX1) is the most widely used promoter in the production of recombinant proteins in the methylotrophic yeast Pichia pastoris. However, as the tight regulation and methanol dependence of PAOX1 restricts its application, it is necessary to develop a flexible induction system to avoid the problems of methanol without losing the advantages of PAOX1. The availability of synthetic biology tools enables researchers to reprogram the cellular behaviour of P. pastoris to achieve this goal.

Results

The characteristics of PAOX1 are highly related to the expression profile of methanol expression regulator 1 (Mxr1). In this study, we applied a biologically inspired strategy to reprogram regulatory networks in P. pastoris. A reprogrammed P. pastoris was constructed by inserting a synthetic positive feedback circuit of Mxr1 driven by a weak AOX2 promoter (PAOX2). This novel approach enhanced PAOX1 efficiency by providing extra Mxr1 and generated switchable Mxr1 expression to allow PAOX1 to be induced under glycerol starvation or carbon-free conditions. Additionally, the inhibitory effect of glycerol on PAOX1 was retained because the synthetic circuit was not activated in response to glycerol. Using green fluorescent protein as a demonstration, this reprogrammed P. pastoris strain displayed stronger fluorescence intensity than non-reprogrammed cells under both methanol induction and glycerol starvation. Moreover, with single-chain variable fragment (scFv) as the model protein, increases in extracellular scFv productivity of 98 and 269% were observed in Mxr1-reprogrammed cells under methanol induction and glycerol starvation, respectively, compared to productivity in non-reprogrammed cells under methanol induction.

Conclusions

We successfully demonstrate that the synthetic positive feedback circuit of Mxr1 enhances recombinant protein production efficiency in P. pastoris and create a methanol-free induction system to eliminate the potential risks of methanol.
  相似文献   

5.
Nitric-oxide synthases (NOSs) are calmodulin-dependent flavoheme enzymes that oxidize l-Arg to nitric oxide (NO) and l-citrulline. Their catalytic behaviors are complex and are determined by their rates of heme reduction (kr), ferric heme-NO dissociation (kd), and ferrous heme-NO oxidation (kox). We found that point mutation (E762N) of a conserved residue on the enzyme''s FMN subdomain caused the NO synthesis activity to double compared with wild type nNOS. However, in the absence of l-Arg, NADPH oxidation rates suggested that electron flux through the heme was slower in E762N nNOS, and this correlated with the mutant having a 60% slower kr. During NO synthesis, little heme-NO complex accumulated in the mutant, compared with ∼50–70% of the wild-type nNOS accumulating as this complex. This suggested that the E762N nNOS is hyperactive because it minimizes buildup of an inactive ferrous heme-NO complex during NO synthesis. Indeed, we found that kox was 2 times faster in the E762N mutant than in wild-type nNOS. The mutational effect on kox was independent of calmodulin. Computer simulation and experimental measures both indicated that the slower kr and faster kox of E762N nNOS combine to lower its apparent Km,O2 for NO synthesis by at least 5-fold, which in turn increases its V/Km value and enables it to be hyperactive in steady-state NO synthesis. Our work underscores how sensitive nNOS activity is to changes in the kox and reveals a novel means for the FMN module or protein-protein interactions to alter nNOS activity.Nitric oxide (NO)2 is a biological mediator that is produced in animals by three NO synthase isozymes (NOS, EC 1.14.13.39): inducible NOS (iNOS), neuronal NOS (nNOS), and endothelial NOS (eNOS) (1, 2). The NOS are modular enzymes composed of an N-terminal oxygenase domain and a C-terminal flavoprotein domain, with a calmodulin (CaM)-binding site connecting the two domains (3). During NO synthesis, the flavoprotein domain transfers NADPH-derived electrons through its FAD and FMN cofactors to a heme located in the oxygenase domain. The FMN-to-heme electron transfer enables heme-dependent oxygen activation and a stepwise conversion of l-Arg to NO and citrulline (4, 5). Heme reduction also requires that CaM be bound to NOS and is rate-limiting for NO biosynthesis (69).NOS enzymes operate under the constraint of having their newly made NO bind to the ferric heme before it can exit the enzyme (10). How this intrinsic heme-NO binding event impacts NOS catalytic cycling is shown in Fig. 1 and has previously been discussed in detail (1013). The l-Arg to NO biosynthetic reaction (FeIII to FeIIINO in Fig. 1) is limited by the rate of ferric heme reduction (kr), because all biosynthetic steps downstream are faster than kr. However, once the ferric heme-NO complex forms at the end of each catalytic cycle, it can either dissociate to release NO into the medium (at a rate kd as shown in Fig. 1) or become reduced by the flavoprotein domain (at a rate kr in Fig. 1; equal to kr) to form the enzyme ferrous heme-NO species (FeIINO), which releases NO very slowly (11, 12). Consequently, two cycles compete during steady-state NO synthesis (Fig. 1); NO dissociation from the ferric heme (kd) is part of a “productive cycle” that releases NO and is essential for NOS bioactivity, whereas reduction of the ferric heme-NO complex (kr′) channels the enzyme into a “futile cycle” that actually represents a NO dioxygenase activity. The rate of futile cycling is also determined by the rate of O2 reaction with the ferrous heme-NO complex (at a rate kox in Fig. 1), which regenerates the ferric enzyme. Surprisingly, NOS enzymes have evolved to have a broad range of kr (varies 40×), kox (varies 15×), and kd (varies 30×) values (Table S1) (12). This causes each NOS to distribute quite differently during steady-state NO synthesis and gives each NOS a unique catalytic profile (12).Open in a separate windowFIGURE 1.Global kinetic model for NOS catalysis. Ferric enzyme reduction (kr) is rate-limiting for the biosynthetic reactions (central linear portion). kcat1 and kcat2 are the conversion rates of the enzyme FeIIO2 species to products in the l-Arg and Nω-hydroxy-l-arginine (NOHA) reactions, respectively. The ferric heme-NO product complex (FeIIINO) can either release NO (kd) or become reduced (kr) to a ferrous heme-NO complex (FeIINO), which reacts with O2 (kox) to regenerate ferric enzyme. Enzyme partitioning and NO release are determined by the relative rates of kr, kox, and kd. This figure is adapted from Ref. 12.The enzyme physical and electronic factors that may set and regulate each of the three kinetic parameters (kr, kox, and kd) in NOS enzymes remain to be fully described. At present, the composition of the NOS flavoprotein domain and CaM appear to be primarily responsible for determining the kr (1417), whereas the composition of the NOS oxygenase domain is presumed to determine the kd and kox (18, 19). Indeed, our recent point mutagenesis study identified a patch of electronegative residues on the FMN subdomain that are required to maintain a normal kr and NO synthesis activity in nNOS, suggesting that subdomain electrostatic interactions are important in the process (20). We found particularly large effects when the negative charge at Glu762 was neutralized via mutation to Asn. Remarkably, the NO synthesis activity of E762N nNOS was double that of wild-type nNOS, despite the mutant displaying a slow kr that was half of wild type. In the current report, we show that the E762N mutation has an additional, unsuspected effect on the kox kinetic parameter of nNOS. How this effect alters distribution of the nNOS enzyme during steady-state catalysis, impacts the apparent Km,O2, and leads to hyperactive NO synthesis is described. Our finding that the nNOS flavoprotein domain can tune a key kinetic parameter that defines the rate of a heme-based reaction in the nNOS oxygenase domain is unusual and suggests a means by which protein-protein interactions could regulate the catalytic behavior of nNOS.  相似文献   

6.
On the basis of the winter bread wheat cultivar Obryi, two independent disomic addition lines BC12F with the chromosome of the E. sibiricus St genome are created. A practical algorithm for determining the probabilities of transmission of the odd chromosome separately through male and female gametes in selfpollination of hemizygous hybrids from the equation p2–(1 + f1f4) × p + f1 = 0 is proposed, where p is the probability of the formation of viable gametes with the considered chromosome and f1 and f4 are the empirical frequencies of the corresponding homozygotes with and without the trait. The probability of transmission of an alien univalent chromosome through pollen (p) is associated with the frequency of its transmission through the egg cell (p) in backcrosses and in self-pollination (1–f4) by the equation p = 1–f4/(1–p). The calculated empirically dependent estimates of the probabilities of transmission of the added chromosome through the egg cell p = 18.7% and through pollen p = 4.3% correspond to the empirical frequencies obtained for backcrosses. The coefficients of the gamete selection V = 0.748 and V = 0.172 are calculated, and the expected segregation for the alien trait controlled by a dominant gene located in the added chromosome is determined—with the trait: without the trait is 0.222: 0.778 in F2; 0.187: 0.813 in equational and 0.043: 0.957 in certational backcrosses.  相似文献   

7.
Two new ligand-containing histidine based on N,N′,N″-tris(N-benzyl-l-histidinyl)tri(2-aminoethyl)amine, L1, namely N,N′,N″-tris[(1S)-2-methoxy-2-oxy-1-(1-benzylimidazol-4-ylmethyl)]nitrilotriacetamide L2 and N,N′,N″-tris{N-benzyl-N-[N-benzyl-N-(N-benzyl-l-histidinyl)-l-histidinyl]-l-histidinyl}tri(2-aminoethyl)amine L3 were prepared. Zinc(II) binding studies by these ligand systems were analyzed by means of potentiometric and 1H NMR titrations in aqueous methanol (33 % v/v). Subsequently their zinc(II) complexes [L1Zn(H2O)](ClO4)2·HClO4 (1), [L2Zn(OH2)](ClO4)2·H2O (2), and ([L3Zn3(H2O)3](ClO4)6·3HClO4·5H2O (3), respectively were synthesized and characterized. The reactivity of the trinuclear complex (3) toward the hydrolysis of the toxic organophosphate parathion was investigated and compared with that of the mononuclear reference complex (1). From the pH dependence of the apparent rate constants, and the deprotonation constant (pKa) of the coordinated water molecules in (1), the active species were confirmed to be {[HL1Zn(OH)]2+/[L1Zn(H2O)]2+} at pH 8.5. The trizinc complex (3) effects hydrolysis of parathion, with three times rate enhancement over the mononuclear (1), indicating that cooperative action of the three zinc centers is limited.  相似文献   

8.
The ligands 1-hydroxymethylpyrazole (hl1), 1-(2-hydroxyethyl)pyrazole (hl2) and 1-(3-hydroxypropyl)pyrazole (hl3) react with [PdCl2(CH3CN)2] to give trans-[PdCl2(hl)2] compounds. Due to a hindered rotation around the Pd-bond, these compounds present two different conformations in solution: anti and syn. The conformation presented depends on the relative disposition of the hydroxyalkylic chains of the two pyrazolic ligands. The present study was carried out on the basis of NMR experiments. The present paper reports the crystal structure of trans-[PdCl2(hl2)2]. The synthesis and characterisation of compounds [Pd(hl)4](BF4)2 (hl = hl1, hl2 and hl3) starting from [Pd(CH3CN)4](BF4)2 and the corresponding chlorocomplexes trans-[PdCl2(hl)2] are also described.  相似文献   

9.
A method is presented for rapid extraction of the total plastoquinone (PQ) pool from Synechocystis sp. strain PCC 6803 cells that preserves the in vivo plastoquinol (PQH2) to -PQ ratio. Cells were rapidly transferred into ice-cold organic solvent for instantaneous extraction of the cellular PQ plus PQH2 content. After high-performance liquid chromatography fractionation of the organic phase extract, the PQH2 content was quantitatively determined via its fluorescence emission at 330 nm. The in-cell PQH2-PQ ratio then followed from comparison of the PQH2 signal in samples as collected and in an identical sample after complete reduction with sodium borohydride. Prior to PQH2 extraction, cells from steady-state chemostat cultures were exposed to a wide range of physiological conditions, including high/low availability of inorganic carbon, and various actinic illumination conditions. Well-characterized electron-transfer inhibitors were used to generate a reduced or an oxidized PQ pool for reference. The in vivo redox state of the PQ pool was correlated with the results of pulse-amplitude modulation-based chlorophyll a fluorescence emission measurements, oxygen exchange rates, and 77 K fluorescence emission spectra. Our results show that the redox state of the PQ pool of Synechocystis sp. strain PCC 6803 is subject to strict homeostatic control (i.e. regulated between narrow limits), in contrast to the more dynamic chlorophyll a fluorescence signal.The photosynthetic apparatus of oxygenic phototrophs consists of two types of photosynthetic reaction centers: PSII and PSI. Both photosystems are connected in series, with electrons flowing from PSII toward PSI through an intermediate electron transfer chain, which comprises the so-called plastoquinone (PQ) pool, plastocyanin and/or cytochrome c553, and the cytochrome b6f complex. The redox potential of the PQ pool is clamped by the relative rates of electron release into and uptake from this pool. Within the PSII complex, electrons are extracted from water at the lumenal side of the thylakoid membrane and transferred to the primary accepting quinone (QA) at the stromal side. The electron is subsequently transferred to a PQ molecule in the secondary accepting quinone (QB) of PSII. The intermediate QB semiquinone, which is formed accordingly, is stable in the QB site for several seconds (Diner et al., 1991; Mitchell, 1993) and subsequently can be reduced to plastoquinol (PQH2). The midpoint potential of QA reduction is approximately −100 mV (Krieger-Liszkay and Rutherford, 1998; Allakhverdiev et al., 2011), whereas the corresponding midpoint potential of the QB semiquinone is close to zero (Nicholls and Ferguson, 2013). PQH2 equilibrates with the PQ pool in the thylakoid membranes, which has a size that is approximately 1 order of magnitude larger than the number of PSII reaction centers (Melis and Brown, 1980; Aoki and Katoh, 1983).PQ is a lipophilic, membrane-bound electron carrier, with a midpoint potential of +80 mV (Okayama, 1976), that can accept two electrons and two protons to form PQH2 (Rich and Bendall, 1980). PQH2 can donate both electrons to the cytochrome b6f complex, one to low-potential cytochrome b6, by which reduced high-potential cytochrome b6 is formed, and one to the cytochrome f moiety on the lumenal side of the thylakoid membrane, where the two protons are released. High-potential cytochrome b6 then donates an electron back to PQ on the stromal side of the membrane, rendering a semiquinone in the PQ-binding pocket on the cytoplasmic face of the b6f complex ready as an acceptor of another electron from PSII, and reduced cytochrome f feeds an electron to a water-soluble electron carrier (i.e. either plastocyanin or cytochrome c553) for subsequent transfer to the reaction center of PSI or to cytochrome c oxidase, respectively (Rich et al., 1991; Geerts et al., 1994; Schubert et al., 1995; Paumann et al., 2004; Mulkidjanian, 2010).Electron transfer through the cytochrome b6f complex proceeds according to the Q-cycle mechanism (Rich et al., 1991). As a result, maximally two protons from the stroma are released into the lumen per electron transferred. This electrochemical proton gradient can be used for the synthesis of ATP by the ATP synthase complex (Walker, 1998). In PSI, another transthylakoid membrane charge separation process is energized by light. Electron transfer within the PSI complex involves iron-sulfur clusters and quinones and leads to the reduction of ferredoxin, the reduced form of which serves as the electron donor for NADPH by the ferredoxin:NADP+ oxidoreductase enzyme (van Thor et al., 1999). The ATP and NADPH generated this way are used for CO2 fixation in a mutual stoichiometry that is close to the stoichiometry at which these two energy-rich compounds are formed at the thylakoid membrane. Normally, this ratio is ATP:NADPH = 3:2 (Behrenfeld et al., 2008).Photosynthetic and respiratory electron transport in cyanobacteria share a single PQ pool (Aoki and Katoh, 1983; Aoki et al., 1983; Matthijs et al., 1984; Scherer, 1990). Respiratory electron transfer provides cells the ability to form ATP in the dark, but this ability is not limited to those conditions. Transfer of electrons into the PQ pool is the result of the joint activity of PSII, respiratory dehydrogenases [in particular those specific for NAD(P)H and succinate], and cyclic electron transport around PSI (Mi et al., 1995; Cooley et al., 2000; Howitt et al., 2001;Yeremenko et al., 2005), whereas oxidation of PQH2 is catalyzed by the PQH2 oxidase, the cytochrome b6f complex, the respiratory cytochrome c oxidase (Nicholls et al., 1992; Pils and Schmetterer, 2001; Berry et al., 2002), and possibly plasma terminal oxidase (Peltier et al., 2010). Multiples of these partial reactions can proceed simultaneously, including respiratory electron transfer during illumination (Schubert et al., 1995), which includes oxygen uptake through a Mehler-like reaction (Helman et al., 2005; Allahverdiyeva et al., 2013).Because of its central location between the two photosystems, the redox state of the PQ pool has been identified as an important parameter that can signal photosynthetic imbalances (Mullineaux and Allen, 1990; Allen, 1995; Ma et al., 2010; Allen et al., 2011). Yet, an accurate estimation of the in vivo redox state of this pool has not been reported in cyanobacteria so far. Instead, the redox state of the PQ pool is widely assumed to be reflected in, or related to, the intensity of the chlorophyll a fluorescence emissions (Prasil et al., 1996; Yang et al., 2001; Gotoh et al., 2010; Houyoux et al., 2011). Imbalance in electron transport through the two photosystems may lead to a loss of excitation energy and, hence, to a loss of chlorophyll a fluorescence emission (Schreiber et al., 1986). Therefore, patterns of chlorophyll a fluorescence (pulse-amplitude modulated [PAM] fluorimetry; Baker, 2008) have widely been adopted for the analysis of (un)balanced photosynthetic electron transfer and, by inference, for indirect recording of the redox state of the PQ pool. However, the multitude of electron transfer pathways in the thylakoid membranes of cyanobacteria (see above) makes it much more complex to explain PAM signals in these organisms than in chloroplasts (Campbell et al., 1998). Additional regulatory mechanisms of nonphotochemical quenching, via the xanthophyll cycle in chloroplasts (Demmig-Adams et al., 2012) and the orange carotenoid protein (Kirilovsky and Kerfeld, 2012) in cyanobacteria, and energy redistribution via state transitions (Allen, 1995; Van Thor et al., 1998) complicate such comparisons even further.Several years ago, an HPLC-based technique was developed for the detection of the redox state of PQH2 in isolated thylakoids (Kruk and Karpinski, 2006), but these results have neither been related to physiological conditions nor to the results of chlorophyll a fluorescence measurements. In this report, we describe an adaptation of this method with elements of a method for estimation of the redox state of the ubiquinone pool in Escherichia coli (Bekker et al., 2007). This modified method allows for reliable measurements of the redox state of the PQ pool of Synechocystis sp. strain PCC 6803 under physiologically relevant conditions. The method uses rapid cell lysis in an organic solvent to arrest all physiological processes, followed by extraction and identification of PQH2 by HPLC separation with fluorescence detection. Next, we manipulated the redox state of the PQ pool with various redox-active agents, with inhibitors of photosynthetic electron flow, and by illumination with light specific for either PSII or PSI. The measured redox state of the PQ pool was then related to the chlorophyll a fluorescence signal and 77 K fluorescence emission spectra of cell samples taken in parallel and to oxygen-exchange rates measured separately. These experiments reveal that, despite highly fluctuating conditions of photosynthetic and respiratory electron flow, a remarkably stable redox state of the PQ pool is maintained. This homeostatically regulated redox state correlates poorly in many of the conditions tested with the more dynamic signal of chlorophyll a fluorescence emission, as measured with PAM fluorimetry. The latter signal only reflects the redox state of QA and not that of the PQ pool.  相似文献   

10.
Two heme-dependent catalase genes were amplified from genomic DNA of Lactobacillus plantarum WCFS1 (KatE1) and Lactobacillus brevis ATCC 367 (KatE2), respectively, and a manganese-containing superoxide dismutase from Lactobacillus casei MCJΔ1 (MnSOD) were cloned into plasmid pELX1, yielding pELX1-KatE1, pELX1-KatE2 and pELX1-MnSOD, then the recombinant plasmids were transferred into L. casei MCJΔ1. The strains of L. casei MCJΔ1/pELX1-KatE1 and L. casei MCJΔ1/pELX1-KatE2 were tolerant at 2 mM H2O2. The survival rates of L. casei MCJΔ1/pELX1-KatE1 and L. casei MCJΔ1/pELX1-KatE2 were 270-fold and 300-fold higher than that of the control strain on a short-term H2O2 exposure, and in aerated condition, the survival cells counts were 146- and 190-fold higher than that of the control strain after 96 h of incubation. Furthermore, L. casei MCJΔ1/pELX1-MnSOD was the best in three recombinants which was superior in the living cell viability during storage when co-storage with Lactobacillus delbrueckii subsp. lactis LBCH-1.  相似文献   

11.
The biogenesis and assembly of photosynthetic multisubunit protein complexes is assisted by a series of nucleus-encoded auxiliary protein factors. In this study, we characterize the dac mutant of Arabidopsis (Arabidopsis thaliana), which shows a severe defect in the accumulation of the cytochrome b6/f complex, and provide evidence suggesting that the efficiency of cytochrome b6/f complex assembly is affected in the mutant. DAC is a thylakoid membrane protein with two predicted transmembrane domains that is conserved from cyanobacteria to vascular plants. Yeast (Saccharomyces cerevisiae) two-hybrid and coimmunoprecipitation analyses revealed a specific interaction between DAC and PetD, a subunit of the cytochrome b6/f complex. However, DAC was found not to be an intrinsic component of the cytochrome b6/f complex. In vivo chloroplast protein labeling experiments showed that the labeling rates of the PetD and cytochrome f proteins were greatly reduced, whereas that of the cytochrome b6 protein remained normal in the dac mutant. DAC appears to be a novel factor involved in the assembly/stabilization of the cytochrome b6/f complex, possibly through interaction with the PetD protein.The cytochrome b6/f (Cyt b6/f) complex is a multisubunit complex that resides in the thylakoid membrane and functions in linear and cyclic electron transport. In the linear process, the complex receives electrons from PSII and transfers them to PSI, a process that is accompanied by the generation of a proton gradient, which is essential for ATP synthesis (Mitchell, 1961; Saraste, 1999). The native form of this complex is present as a dimer with a mass of 310 kD that can be converted into a 140-kD monomer with increasing detergent concentrations (Huang et al., 1994; Breyton et al., 1997; Mosser et al., 1997; Baniulis et al., 2009). In higher plants, the Cyt b6/f monomer contains at least eight subunits: Cyt f, Cyt b6, PetC, PetD, PetM, PetL, PetG, and PetN (Wollman, 2004). PetC and PetM are encoded by nuclear genes, whereas the others are encoded by plastid genes. It has been shown that PetG and PetN are necessary for complex stability in tobacco (Nicotiana tabacum; Schwenkert et al., 2007). By contrast, PetL is not required for the accumulation of other subunits of the Cyt b6/f complex, even though it is involved in the stability and formation of the functional dimer (Bendall et al., 1986; Schwenkert et al., 2007). Inactivation of PetC in Arabidopsis (Arabidopsis thaliana) resulted in significantly reduced amounts of Cyt b6/f subunits and completely blocked linear electron transport, indicating that PetC participates in the formation of the functionally assembled Cyt b6/f complex (Maiwald et al., 2003). In Synechocystis sp. PCC 6803, the PetM subunit has no essential role in Cyt b6/f complex electron transfer or accumulation; however, the absence of this subunit apparently affects the levels of other protein complexes involved in energy transduction (Schneider et al., 2001). In addition to the other proteins, FNR was identified as a subunit of the Cyt b6/f complex isolated from spinach (Spinacia oleracea) thylakoid membranes (Zhang et al., 2001).Previous research has revealed how the Cyt b6/f complex assembles into a functional dimer (Bendall et al., 1986; Lemaire et al., 1986; Kuras and Wollman, 1994). In the Cyt b6/f complex, Cyt b6 and PetD form a mildly protease-resistant subcomplex that serves as a template for the assembly of Cyt f and PetG, producing a protease-resistant cytochrome moiety (Wollman, 2004). The PetC and PetL proteins then participate in the assembly of the functional dimer (Schwenkert et al., 2007). PetD becomes more unstable in the absence of Cyt b6, and the synthesis of Cyt f is greatly reduced when either Cyt b6 or PetD is inactivated, indicating that both Cyt b6 and PetD are prerequisite for the synthesis of Cyt f (Kuras and Wollman, 1994). The reduced synthesis of Cyt f can be explained by the so-called CES (for controlled by epistasy of synthesis) mechanism. It is suggested that, in this mechanism, the synthesis rate of some chloroplast-encoded subunits of photosynthetic protein complexes is regulated by the availability of their assembly partners from the same complexes (Choquet et al., 2001). The mechanism of CES for Cyt f has been studied in detail in Chlamydomonas reinhardtii (Choquet et al., 1998; Choquet and Vallon, 2000). In it, the unassembled Cyt f inhibits its own translation through a negative feedback mechanism, and MCA1 and TCA1 have been demonstrated to be involved in the regulation of Cyt f synthesis (Boulouis et al., 2011).Many studies have focused on understanding the conversion of apocytochrome to holocytochrome via the covalent binding of heme in Cyt f and Cyt b6 during the assembly of Cyt b6/f through the CCS and CCB pathways (Nakamoto et al., 2000; Wollman, 2004; de Vitry, 2011). The CCS pathway was originally discovered in the green alga C. reinhardtii through genetic studies of ccs mutants (for cytochrome c synthesis) that display a specific defect in membrane-bound Cyt f and soluble Cyt c6, two thylakoid lumen-resident c-type cytochromes functioning in photosynthesis (Xie and Merchant, 1998). In the CCS pathway, six loci that include plastid ccsA and nuclear CCS1 to CCS5 have been found in C. reinhardtii (Xie and Merchant, 1998). In these mutants, the apocytochrome is normally synthesized, targeted, and processed, but heme attachment is perturbed. The CCB pathway is involved in the covalent attachment of heme c(i) to Cyt b6 on the stromal side of the thylakoid membranes (Kuras et al., 2007). The ccb mutants show defects in the accumulation of subunits of the Cyt b6/f complex and covalent binding of heme to Cyt b6 (Lyska et al., 2007; Lezhneva et al., 2008). However, heme binding is not a prerequisite for the assembly of Cyt b6 into the Cyt b6/f complex, although the fully formed Cyt b6/f showed an increased sensitivity to protease (Saint-Marcoux et al., 2009).The assembly of the Cyt b6/f complex is a multistep process, and current studies have shown that the covalent binding of heme to Cyt f and Cyt b6 is highly regulated. Thus, it is reasonable to speculate that, similar to the other photosynthetic protein complexes (Mulo et al., 2008; Nixon et al., 2010; Rochaix, 2011), the assembly of the Cyt b6/f complex is also assisted by many nucleus-encoded factors. In this study, we characterized an Arabidopsis protein, DAC (for defective accumulation of Cyt b6/f complex), that seems to be involved in the assembly of the Cyt b6/f complex. In addition, we provide evidence that DAC interacts directly with PetD before it assembles within the Cyt b6/f complex.  相似文献   

12.
The evolution rate v(t) varies among diverse biosystems, but a general theory can be formulated when the dynamics of the biosystem stater x = x(t) = (x1, x2, x m ) T is considered in the m-dimensional space of states. A mathematical approach is proposed for evaluating such processes and describes the processes in terms of particular chaos of the statistical distribution functions f(x). In the case of complex multicomponent systems with a high dimension number m (m ?1) of the phase space of states, we propose using pairwise comparison matrices of samples x(t) when homeostasis is constant and calculating the parameters of quasiattractors. The Glensdorff–Prigogine thermodynamic approach to estimating evolution is inefficient in assessing the third-type systems, while it is applicable and the Prigogine theorem works at the level of molecular systems. Alterations in the state of the human neuromuscular system were found to lead to chaotic changes in the statistical functions f(x) in tremor recording samples, while quasiattractor parameters demonstrate a certain regularity.  相似文献   

13.
The formation of a peaked bell-shaped profile of the electron density n e (r) in the T-11M tokamak (B t=1 T, R/a = 0.7/0.2 m, I p = 100 kA, t shot ≤ 300 ms, Li and C limiters) was observed in Li experiments carried out in the near-plateau collisionality regime (the collisionality parameter at one-half of the minor radius was v* ≥ 0.5) under the conditions of low hydrogen recycling and intense hydrogen influx from the plasma edge. It is well known that peaked n e (r) profiles are observed in collisionless regimes at v* values as low as 10?1–10?2 or in impurity-contaminated discharges, in which this effect can be attributed to the impurity accumulation on the plasma column axis. Moreover, a bell-shaped n e (r) profile in discharges with low n e can result from the ionization of hydrogen atoms at the column axis, where they arrive from the plasma edge due to cascade charge-exchange. In quasi-steady lithium discharges in T-11M, however, peaked n e (r) profiles were observed at a relatively high central electron density n e (0) and relatively high collision frequency, such that the influence of impurities on the n e (r) profile could be ignored (Z eff = 1.1±0.1). To explain this effect, one has to assume that the pinching of hydrogen ions in T-11M is anomalous. The lower estimate of the observed pinch velocity is 4 ± 1 m/s, which is three to five times higher than the velocity of the neoclassical (Ware) pinch, characteristic of these conditions. The work is devoted to the experimental study of this effect.  相似文献   

14.
15.
Last-century climate change has led to variable increases of the intrinsic water-use efficiency (Wi; the ratio of net CO2 assimilation to stomatal conductance for water vapor) of trees and C3 grassland ecosystems, but the causes of the variability are not well understood. Here, we address putative drivers underlying variable Wi responses in a wide range of grassland communities. Wi was estimated from carbon isotope discrimination in archived herbage samples from 16 contrasting fertilizer treatments in the Park Grass Experiment, Rothamsted, England, for the 1915 to 1929 and 1995 to 2009 periods. Changes in Wi were analyzed in relation to nitrogen input, soil pH, species richness, and functional group composition. Treatments included liming as well as phosphorus and potassium additions with or without ammonium or nitrate fertilizer applications at three levels. Wi increased between 11% and 25% (P < 0.001) in the different treatments between the two periods. None of the fertilizers had a direct effect on the change of Wi (ΔWi). However, soil pH (P < 0.05), species richness (P < 0.01), and percentage grass content (P < 0.01) were significantly related to ΔWi. Grass-dominated, species-poor plots on acidic soils showed the largest ΔWi (+14.7 μmol mol−1). The ΔWi response of these acidic plots was probably related to drought effects resulting from aluminum toxicity on root growth. Our results from the Park Grass Experiment show that Wi in grassland communities consistently increased over a wide range of nutrient inputs, soil pH, and plant community compositions during the last century.The intrinsic water-use efficiency (Wi) of plants is controlled by photosynthetic carbon assimilation and stomatal conductance via the leaf-level coupling of CO2 and water fluxes. A general, but variable, increase of Wi under rising atmospheric CO2 has been observed in long-term studies (Peñuelas et al., 2011; Franks et al., 2013; Saurer et al., 2014), but little is known about other environmental or ecosystem factors, which may interact with the effect of increasing CO2 on Wi. An improved understanding of putative interactive mechanisms is important because changes in Wi may have significant effects on the global terrestrial carbon and water cycles (Gedney et al., 2006; Betts et al., 2007). This study explores the interactive effects of the increase in atmospheric CO2 (observed over the last century), nutrient loading, and soil pH together with other related effects on plant species richness and functional group composition on the coupling of plant CO2 and water fluxes in a seminatural grassland in southeastern England.Wi is a leaf-level efficiency that has also been termed potential water-use efficiency or physiological water-use efficiency, as it excludes the direct influence of vapor pressure deficit (VPD), a parameter determined by environmental conditions, on leaf-level water-use efficiency (Farquhar et al., 1989; Franks et al., 2013). Wi reports the relationship between net CO2 assimilation rate (An) and stomatal conductance for water vapor (gH2O):(1)According to the first law of Fick, An can be given as the product of the stomatal conductance for CO2 (gCO2) and the concentration gradient between the atmosphere (ca) and the leaf internal gas space (ci): An= gCO2 (caci). Using gCO2 (caci) instead of An in Equation 1, replacement of gH2O/gCO2 by the numerical value of gH2O/gCO2 (1.6) and rearrangement yields the following alternative expression of Wi:(2)Equation 2 reveals that past changes of Wi must have been controlled by two parameters: the change of ca and the concurrent change of 1 – ci/ca, the relative gradient for CO2 diffusion into the leaf (Franks et al., 2013). A change in the relative gradient is determined by the changes in An relative to gH2O, as leaves respond to changing ca and other environmental factors. In particular, Equation 2 shows that any variation in the climate change response of Wi is determined by the ci/ca response, if the comparison is made for vegetation at the same location and in the same period of time.Studies with C3 vegetation, including trees/forests and C3 grasslands, have revealed a general increase of Wi in the last century (Bert et al., 1997; Duquesnay et al., 1998; Feng, 1999; Arneth et al., 2002; Saurer et al., 2004; Barbosa et al., 2010; Köhler et al., 2010; Andreu-Hayles et al., 2011). In many cases, ci/ca, estimated by 13C discrimination (Farquhar et al., 1989), varied relatively little. Indeed, it has been suggested, based on theoretical grounds and empirical evidence from studies over geological/evolutionary to short time scales, that adaptive feedback responses will tend to maintain ci/ca approximately constant (Ehleringer and Cerling, 1995; Franks et al., 2013), as plants optimize carbon gain with respect to water loss (Cowan and Farquhar, 1977). Yet, ci/ca-dependent variation in the Wi response to climate change has also been noted (Peñuelas et al., 2011; Köhler et al., 2012) over the last century, indicating that additional factors, perhaps including other global change drivers, can modify the Wi response over this time scale, at least transiently. A meta-analysis by Peñuelas et al. (2011) reports ci/ca-dependent increases of Wi for different forests between 6% and 36% from the early 1960s to 2000s. A recent study by Saurer et al. (2014) on European forest trees found increases in Wi ranging from 1% to 53% during the last century. The strongest increase of Wi was recorded in regions where summer soil-water availability decreased in the last century. For different grassland communities, the ci/ca-dependent increases of Wi varied between 13% and 28% at one site (Köhler et al., 2012) from 1915 to 2009. Evidently, such variation can have important repercussions for the coupling of terrestrial CO2 and water fluxes. Yet, little is known about the mechanism(s) underlying the variation.At the Park Grass Experiment (PGE) at Rothamsted, England, Köhler et al. (2012) observed a nitrogen supply-dependent enhancement of the Wi response on plots receiving nitrate fertilizer and maintained at a near-neutral soil pH by liming. However, the actual relationship between nitrogen supply and Wi response did not hold when the unlimed control (soil pH approximately 5.2) was included in the comparison. Remarkably, however, there was a significant positive relationship between the grass content of the community and the Wi response of the experimental plots in the investigation. These results suggested that the effect of nutrient supply on the Wi response of the grassland communities was indirect, perhaps working via effects on soil pH and/or vegetation composition (plant species richness or functional group composition).The PGE provides a unique opportunity to study century-scale variation in the ci/ca-dependent variation of Wi for a wide range of diverse grassland communities. Much of the extant ecosystem-scale variability of plant species richness and soil pH in temperate grasslands of Europe (Ceulemans et al., 2014) is included in the range of plot-scale plant species richness and soil pH at the PGE (which is reported in this investigation). The different long-term applications of fertilizer and lime over the past century have resulted in substantial changes in soil pH, species richness, and grass content on the experimental plots, but in most cases, within-plot changes over the study period considered here (1915–2009) were comparatively small (Crawley et al., 2005; Silvertown et al., 2006). All experimental plots are located at the same site and are exposed to the same weather conditions. Consequently, trends in climate as a direct driver for differences in Wi between plots can be ruled out.Here, we explore putative mechanisms underlying eventual ci/ca-dependent variation of Wi during the last century at the PGE by, first, quantifying the sustained effect of a wide range of contrasting fertilizer treatments (n = 16) on the change of Wi during the last century and, second, analyzing the relationships between the observed Wi response of treatments and the respective nutrient status, soil pH, plant species richness, and plant functional group composition of the grassland communities.  相似文献   

16.
17.
The classical definition of mesophyll conductance (g m) represents an apparent parameter (g m,app) as it places (photo)respired CO2 at the same compartment where the carboxylation by Rubisco takes place. Recently, Tholen and co-workers developed a framework, in which g m better describes a physical diffusional parameter (g m,dif). They partitioned mesophyll resistance (r m,dif = 1/g m,dif) into two components, cell wall and plasmalemma resistance (r wp) and chloroplast resistance (r ch), and showed that g m,app is sensitive to the ratio of photorespiratory (F) and respiratory (R d) CO2 release to net CO2 uptake (A): g m,app = g m,dif/[1?+?ω(F?+?R d)/A], where ω is the fraction of r ch in r m,dif. We herein extend the framework further by considering various scenarios for the intracellular arrangement of chloroplasts and mitochondria. We show that the formula of Tholen et al. implies either that mitochondria, where (photo)respired CO2 is released, locate between the plasmalemma and the chloroplast continuum or that CO2 in the cytosol is completely mixed. However, the model of Tholen et al. is still valid if ω is replaced by ω(1?σ), where σ is the fraction of (photo)respired CO2 that experiences r ch (in addition to r wp and stomatal resistance) if this CO2 is to escape from being refixed. Therefore, responses of g m,app to (F?+?R d)/A lie somewhere between no sensitivity in the classical method (σ =1) and high sensitivity in the model of Tholen et al. (σ =0).  相似文献   

18.
19.
20.
Photosynthetic parameters including net photosynthetic rate (PN), transpiration rate (E), water-use efficiency (WUE), and stomatal conductance (gs) were studied in indoor C3 plants Philodendron domesticum (Pd), Dracaena fragans (Df), Peperomia obtussifolia (Po), Chlorophytum comosum (Cc), and in a CAM plant, Sansevieria trifasciata (St), exposed to various low temperatures (0, 5, 10, 15, 20, and 25°C). All studied plants survived up to 0°C, but only St and Cc endured, while other plants wilted, when the temperature increased back to room temperature (25°C). The PN declined rapidly with the decrease of temperature in all studied plants. St showed the maximum PN of 11.9 μmol m?2 s?1 at 25°C followed by Cc, Po, Pd, and Df. E also followed a trend almost similar to that of PN. St showed minimum E (0.1 mmol m?2 s?1) as compared to other studied C3 plants at 25°C. The E decreased up to ≈4-fold at 5 and 0°C. Furthermore, a considerable decline in WUE was observed under cold stress in all C3 plants, while St showed maximum WUE. Similarly, the gs also declined gradually with the decrease in the temperature in all plants. Among C3 plants, Pd and Po showed the maximum gs of 0.07 mol m?2 s?1 at 25°C followed by Df and Cc. However, St showed the minimum gs that further decreased up to ~4-fold at 0°C. In addition, the content of photosynthetic pigments [chlorophyll a, b, (a+b), and carotenoids] was varying in all studied plants at 0°C. Our findings clearly indicated the best photosynthetic potential of St compared to other studied plants. This species might be recommended for improving air quality in high-altitude closed environments.  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