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 15 毫秒
1.
Methanol biofiltration using methylotrophic microorganisms has been previously reported by various authors. In a previous study, a modified strain of Pichia pastoris was tested for the ability to produce endochitinase (Ech42) when coupled with methanol vapor biodegradation in batch tests. The next challenge was to validate the process in a continuous system. Thus, in the present study, a biofilter packed with perlite and inoculated with P. pastoris transformed with the plasmid pPIC-ech42 was used for methanol vapor biofiltration and the continuous production of recombinant endochitinase (Ech42) for 60 days. The maximum elimination capacity (EC) of methanol obtained was 1320 g m?3 h?1 at a loading rate of 1465 g m?3 h?1. The extracellular protein production rate in the leachate was 2360 μg h?1 with a chitinase enzymatic activity of 123 U L?1. The protein content on the biofilm samples was negligible, indicating the effectiveness of the overall process and of P. pastoris to excrete proteins. The carbon balance indicated that 81% of the consumed methanol was mineralized and 5.8% was incorporated into biomass. The results of this study and the economic balance underscore the promising application of linking methanol vapor biofiltration to the continuous production of recombinant proteins.  相似文献   

2.
The psychrotrophic Sanguibacter antarcticus KOPRI 21702T, isolated from Antarctic seawater, produced a cold-adapted chitinolytic enzyme that is a new 55 kDa family 18 chitinase (Chi21702). Chi21702 exhibited high activities toward pNP-(GlcNAc)2 and pNP-(GlcNAc)3 with no activity for pNP-GlcNAc, indicating that it prefers chitin chains longer than dimers, just as endochitinases do. A mixture of GlcNAc and GlcNAc2 was produced as a main product by Chi21702 activity from chitin oligosaccharides and swollen chitin, while less GlcNAc3 was produced. These results show that Chi21702 has an endochitinase activity, randomly hydrolyzing chitin at internal sites. Chi21702 displayed chitinase activity at 0–40 °C (optimal temperature of 37 °C), maintained its activity at pH 4–11 (optimal pH of 7.6). Interestingly, Chi21702 exhibited relative activities of 40% and 60% at 0 and 10 °C, respectively, in comparison to 100% at 37 °C, which is higher than those of the previously characterized, cold-adapted, chitinases from bacterial strains.  相似文献   

3.
Substrate inhibition hinders chitinolytic β-N-acetyl-d-hexosaminidases in producing N-acetyl-d-glucosamine (GlcNAc), the valuable chemical widely applied in medical and food industries. Here we focused on a promising chitinolytic enzyme, OfHex1 from the insect, Ostrinia furnacalis. By structural analysis of OfHex1, five residues nearby the active pocket including V327, E328, Y471, V484 and W490 were chosen and nine mutants including V327G, E328Q, E328A, Y471V, V484R, W490A, W490H, V327G/V484R/W490A and V327G/Y471V/W490H were constructed and recombinantly expressed in Pichia pastoris. The best-performing mutant, W490A, obtained by a higher yield of 5 mg/L, did not show substrate inhibition even when 5 mM of the substrates, (GlcNAc)2–4, were applied. The kcat/Km values for (GlcNAc)2–4 are 239.8, 111.3 and 79.8 s?1 mM?1, respectively. Besides, the pH stability of the mutant ranges from pH 4 to 11 and the thermal stability is up to 50 °C. This work suggests the W490A mutant might be an ideal biocatalyst for GlcNAc production from chitin.  相似文献   

4.
《Process Biochemistry》2010,45(1):75-80
Xylanase is an important industrial enzyme. In this research, to improve the thermostability and biochemical properties of a xylanase from Aspergillus niger F19, five arginine substitutions and a disulfide bond were introduced by site-directed mutagenesis. The wild-type gene xylB and the mutant gene xylCX8 were expressed in Pichia pastoris. Compare to those of the wild-type enzyme, the optimal reaction temperature for the mutant enzyme increased from 45 °C to 50 °C, the half-life of the mutant enzyme extended from 10 min to 180 min, and the specific activity increased from 2127 U/mg to 3330 U/mg. However, the Vmax and Km of the mutant xylanase decreased. The enzyme activity in broth obtained from shake flask cultures could be induced to 1850 U/mL in 7 days, which is higher than results reported previously. Furthermore, the highest achievable enzyme activity was 7340 U/mL from 140 g/L of biomass in a 3 L fermentor used in our study.  相似文献   

5.
A highly chitinolytic strain Penicillium ochrochloron MTCC 517 was procured from MTCC, Chandigarh, India. Culture medium supplemented with 1% chitin was found to be suitable for maximum production of chitinase. Purification of extracellular chitinase was done from the culture medium by organic solvent precipitation and DEAE-cellulose column chromatography. The chitinase was purified 6.92-fold with 29.9% yield. Molecular mass of purified chitinase was found to be 64 kDa by SDS-PAGE. The chitinase showed optimum temperature 40 °C and pH 7.0. The enzyme activity was completely inhibited by Hg2+, Zn2+, K+ and NH4+. The enzyme kinetic study of purified chitinase revealed the following characteristics, such as apparent Km 1.3 mg ml?1, Vmax 5.523 × 10?5 moles l?1 min?1 and Kcat 2.37 s?1 and catalytic efficiency 1.82 s?1 M?1. The enzyme hydrolyzed colloidal chitin, glycol chitin, chitosan, glycol chitosan, N,N′-diacetylchitobiose, p-nitrophenyl N-acetyl-β-d-glucosaminide and 4-methylumbelliferyl N-acetyl-β-d-glucosaminide. The chitinase of P. ochrochloron MTCC 517 is an exoenzyme, which gives N-acetylglucosamine as the main hydrolyzate after hydrolysis of colloidal chitin. Protoplasts with high regeneration capacity were obtained from Aspergillus niger using chitinase from P. ochrochloron MTCC 517. Since it also showed antifungal activity, P. ochrochloron MTCC 517 seems to be a promising biocontrol agent.  相似文献   

6.
To investigate the contribution of solvent environments to the enzymatic function of Escherichia coli dihydrofolate reductase (DHFR), the salt-, pH-, and pressure-dependence of the enzymatic function of the wild-type protein were compared with those of the active-site mutant D27E in relation to their structure and stability. The salt concentration-dependence of enzymatic activity indicated that inorganic cations bound to and inhibited the activity of wild-type DHFR at neutral pH. The BaCl2 concentration-dependence of the 1H–15N HSQC spectra of the wild-type DHFR–folate binary complex showed that the cation-binding site was located adjacent to the Met20 loop. The insensitivity of the D27E mutant to univalent cations, the decreased optimal pH for its enzymatic activity, and the increased Km and Kd values for its substrate dihydrofolate suggested that the substrate-binding cleft of the mutant was slightly opened to expose the active-site side chain to the solvent. The marginally increased fluorescence intensity and decreased volume change due to unfolding of the mutant also supported this structural change or the modified cavity and hydration. Surprisingly, the enzymatic activity of the mutant increased with pressurization up to 250 MPa together with negative activation volumes of ? 4.0 or ? 4.8 mL/mol, depending on the solvent system, while that of the wild-type was decreased and had positive activation volumes of 6.1 or 7.7 mL/mol. These results clearly indicate that the insertion of a single methylene at the active site could substantially change the enzymatic reaction mechanism of DHFR, and solvent environments play important roles in the function of this enzyme.  相似文献   

7.
The enzymatic decolorization process of manganese peroxidase (MnP) is a complex system, which is greatly affected by the concentrations of H2O2, Mn2+, dye and enzyme. This work aimed to study these factors and investigate the combined interactions between them by applying response surface methodology (RSM) for decolorization of Congo red with MnP from Schizophyllum sp. F17, meanwhile conventional one-factor-at-a-time analysis was carried out. Through the one-factor-at-a-time analysis the optimized H2O2, Mn2+, Congo red and MnP extract was 0.2 mM, 0.5 mM, 50 mg/l and 0.8 ml, respectively, and the maximum decolorization attained under such conditions was 24.2%. Response surface analysis was conducted through Box–Behnken design and a second-order polynomial model (R2 = 0.8565) was generated to describe the combined effect and the interactions quantificationally. ANOVA analysis indicated that the interactions between H2O2 and MnP, between dye and MnP were significant; the optimum condition through RSM was found to be 0.35 mM H2O2, 0.5 mM Mn2+, 75 mg/l Congo red and 1.4 ml MnP extract, for maximum decolorization of 30.8%.  相似文献   

8.
Docosahexaenoic acid (DHA) is an important polyunsatured fatty acid (PUFA) which can be purified from tuna fish oil fatty acids by selective enzymatic esterification. The present paper investigates the kinetic study for selective esterification of tuna fish oil fatty acids with butanol catalyzed by Rhizopus oryzae lipase (ROL) in biphasic solvent system. Under the most suitable reaction conditions, 76.2% esterification was achieved in 24 h. Different kinetic models for esterification given by Segel [1], Oliveira et al. [2], Gogoi et al. [3], and Kraai et al. [4] were tested for fitting the esterification data and the model given by Oliveira et al. [2] was found to be most suitable. The model given by Prazeres et al. [5] for hydrolysis was also tested for esterification and the model with second order product inhibition was found to provide better match between the predicted and experimental values than that of model by Oliveira et al. [2]. The kinetic model was fitted using MATLAB® to determine the best kinetic parameters. The average value of kinetic constants using the model given by Prazeres et al. were estimated as Km = 23.6 μmoles FFA/ml, Ki1 = 4.6 × 10−5 μmoles FFA/mg enzyme h, Ki2 = 0.0062 μmoles FFA/mg enzyme h and K2 = 149.5 μmoles FFA/mg enzyme h.  相似文献   

9.
Crude glycerol from the biodiesel industry was used as carbon source for high cell density fed-batch cultivation of Pichia pastoris aiming at producing a chitin–glucan complex (CGC). More than 100 g L?1 biomass was obtained in less than 48 h. The yield of biomass on a glycerol basis was 0.55 g g?1 during the batch phase and 0.63 g g?1 during the fed-batch phase. The chitin–glucan complex was recovered from the yeast cell wall by hot alkaline extraction. CGC content in the cell wall was found to be relatively constant throughout the cultivation (18–26%) with a volumetric productivity of 1.28 g L?1 h?1 at the end of the fed-batch phase. The molar ratio of chitin:β-glucan in the extracted biopolymer was 16:84, close to other CGC extracted from Aspergillus biomass. The extracted polymer was characterized by Differential Scanning Calorimetry (DCS) and solid-state Nuclear Magnetic Resonance (NMR) spectroscopy and compared with commercial biopolymers, namely, crab shell chitin and/or chitosan, algal β-glucan (laminarin) and fungal chitin–glucan complex (kiOsmetine).  相似文献   

10.
A putative laccase gene was cloned from Shigella dysenteriae W202 and expressed in Escherichia coli as a soluble fusion protein with high yield. The purified product (Wlac) was characterized as the CueO-like laccase from E. coli, a monomer of molecular mass 55 kDa, with a maximum activity of 24.4 U/mg (Km = 0.086) and a pH optimum of 2.5, in a standard assay using ABTS (2,2′-azino-di(3-ethyl-benzthiazoline-6-sulfonate) as the substrate. Activity was stable at 0–25 °C but inhibited above 40 °C. Purified Wlac was completely inhibited by 200 mM EDTA and partially by 32 mM SDS, 50 mM NaN3 and 60 mM thioglycolic acid. Activity was stimulated by Cu2+; other metal ions had only slight or negative effects. Two mutated variants, WlacS and WlacD, were obtained by substituting Glu 106 with Phe 106, and adding a deletion of an α-helix domain (from Leu 351 to Gly 378). WlacS had a 2.2-fold (52.9 U/mg) and WlacD a 3.5-fold (85.1 U/mg) higher enzyme activity than the wild-type laccase and WlacD showed greater thermostability at higher temperatures. Sce VMA intein-associated fusion proteins maintained ~80% of total enzyme activity. Thus, deletion and site-directed mutagenesis of laccases are capable of promoting both enzymatic activity and thermostability.  相似文献   

11.
《Process Biochemistry》2007,42(9):1319-1325
A two-step enzymatic resolution process for production of (R)- and (S)-glycidyl butyrate was investigated and the lipases were screened. The first step involved a hydrolysis of (R,S)-glycidyl butyrate catalyzed by porcine pancreatic lipase (S-favored) with an E of 21 for production of (R)-glycidyl butyrate (13.2 mmol, 98% ee, 36% yield) under the optimal conditions (pH 7.4, 30 °C, 30 mg/ml CTAB). Then, the recovered (R)-enriched glycidol (19.8 mmol, 65% ee, 56% yield) was used for transesterification catalyzed by Novozym 435 (R-favored) with an E of 69 to obtain (S)-glycidyl butyrate (15.1 mmol, 98% ee, 42% yield) under the optimum conditions (aW = 0.24, n-heptane, 80 min).  相似文献   

12.
Sorbitol, one of the main by-products of growth on high sucrose concentrations, is catalyzed by glucose-fructose oxidoreductase (GFOR, EC 1.1.99.28) in Zymomonas mobilis, which decreases the ethanol yield. In this study, an unmarked gfo mutant from Z. mobilis ZM4 was constructed using a site-specific FLP recombinase, and growth and ethanol production were evaluated with or without the addition of sorbitol to the media. The inactivation of gfo had contrasting effects in different substrates, especially at high concentrations. The maximum specific growth rate (μm) and theoretical ethanol yield value (Ym) increased from 0.065 h−1 and 60.56% to 0.094 h−1 and 83.87% in 342 g/L sucrose, respectively. Conversely, in 200 g/L glucose, gfo inactivation decreased μm and Ym from 0.15 h−1 and 89.85% to 0.10 h−1 and 67.59%, respectively, and prolonged the lag period from 16 h to 40 h. The addition of sorbitol slightly accelerated growth and sucrose hydrolysis by the gfo mutant in 342 g/L sucrose; however, addition of sorbitol restored the μm and Ym of the gfo mutant in 200 g/L glucose to 0.14 h−1 and 82.50%, respectively. Inactivation of gfo had a small effect on fructose utilization, and a positive one on mixture of glucose and fructose similar to that on sucrose. These results provide further understanding of the osmoregulation mechanisms in Z. mobilis and may help to exploit the biotechnological applications of this industrially important bacterium.  相似文献   

13.
The present study is the first report demonstrating that ionic liquids consisting of cholinium cations and linear carboxylate anions ([Ch][CA] ILs) can be used for pretreatment of lignocellulosic materials to enhance subsequent enzymatic saccharification. Six variants of [Ch][CA] ILs were systematically prepared by combining cholinium cations with linear monocarboxylate anions ([CnH2n+1–COO], n = 0–2) or dicarboxylate anions ([HOOC–CnH2n+1–COO], n = 0–2). These [Ch][CA] ILs were analyzed for their toxicity to yeast cell growth and their ability to pretreat kenaf powder for subsequent enzymatic saccharification. When assayed against yeast growth, the EC50 for choline acetate ([Ch][OAc]) was 510 mM, almost one order of magnitude higher than that for 1-ethyl-3-methylimidazolium acetate ([Emim][OAc]). The cellulose saccharification ratio after pretreatment at 110 °C for 16 h with [Ch][OAc] (100.6%) was almost comparable with that after pretreatment with [Emim][OAc]. Therefore, [Ch][OAc] is a biocompatible alternative to [Emim][OAc] for lignocellulosic material pretreatment.  相似文献   

14.
Different crude microbial proteases were applied for chitin extraction from shrimp shells. A Box–Behnken design with three variables and three levels was applied in order to approach the prediction of optimal enzyme/substrate ratio, temperature and incubation time on the deproteinization degree with Bacillus mojavensis A21 crude protease. These optimal conditions were: an enzyme/substrate ratio of 7.75 U/mg, a temperature of 60 °C and an incubation time of 6 h allowing to predict 94 ± 4% deproteinization. Experimentally, in these optimized conditions, a deproteinization degree of 88 ± 5% was obtained in good agreement with the prediction and larger than values generally given in literature. The deproteinized shells were then demineralized to obtain chitin which was converted to chitosan by deacetylation and its antibacterial activity against different bacteria was investigated. Results showed that chitosan dissolved at 50 mg/ml markedly inhibited the growth of most Gram-negative and Gram-positive bacteria tested.  相似文献   

15.
Chitin based polyurethane elastomers (PUEs) were synthesized by step growth polymerization techniques using poly (ε-caprolactone) (PCL), 4, 4′- diphenylmethane diisocyanate (MDI) and blends of chitin and 1,4-butanne diol (BDO). The conventional spectroscopic characterization of the samples with FT-IR, 1H NMR and 13C NMR were in accordance with proposed PUEs structure. The crystalline behavior of the synthesized polymers were investigated by X-ray diffraction (XRD), differential scanning calorimetery (DSC), optical microscopic technique and loss tangent curves (tan δ peaks). Results showed that crystallinity of the synthesized PUEs samples was affected by varying the chitin contents used as chain extender. The contents of chitin favors the formation of more ordered structure, as higher peak intensities were obtained from the PU extended with chitin than 1,4-butane diol (BDO). X-ray diffraction experiments results correlates with optical microscopy findings. The higher ΔH value; 41.57 (J g?1) was found in the samples extended with chitin than BDO (31.32 J g?1).  相似文献   

16.
Organophosphorus compounds (OPs) are widely used as pesticides because of their ability to inhibit the activity of acetylcholinesterase (AChE) in the nervous system. Thus, AChE is generally used as a biosensor for pesticide detection. Due to the instability of AChE a more stable enzyme would be desirable for robust applications. We investigated the sensitivity of a thermostable carboxylesterase from the archaeon Archaeoglobus fulgidus (AFEST) to seven selected OPs. The IC50 of dichlorvos against AFEST (50.8 ± 2.6 nM) was 10-fold lower than that of the commercially obtained AChE, indicating that AFEST had higher sensitivity. Its sensitivity for the other OPs was lower than AChE. To enhance the sensitivity of AFEST to OPs, site-directed mutations were introduced in the cap domain of AFEST. The sensitivity of mutant N44S/S48V was enhanced toward all seven OPs compared to the wild-type and was higher than AChE for four OPs, including paraoxon (3.3 ± 0.01 nM), dichlorvos (28.0 ± 0.6 nM), profenofos (43.0 ± 1.0 nM), and diazinon (3.0 ± 0.2 nM). The half-lives of AFEST and the mutant N44S/S48V at 37 °C were over 15 d. The interactions between the enzymes and select OPs were investigated by molecular docking. The results demonstrated that AFEST and the mutant N44S/S48V have the potential to be biosensor for OP detection.  相似文献   

17.
《Aquatic Botany》2005,81(4):326-342
The effects of NH4+ or NO3 on growth, resource allocation and nitrogen (N) uptake kinetics of two common helophytes Phragmites australis (Cav.) Trin. ex Steudel and Glyceria maxima (Hartm.) Holmb. were studied in semi steady-state hydroponic cultures. At a steady-state nitrogen availability of 34 μM the growth rate of Phragmites was not affected by the N form (mean RGR = 35.4 mg g−1 d−1), whereas the growth rate of Glyceria was 16% higher in NH4+-N cultures than in NO3-N cultures (mean = 66.7 and 57.4 mg g−1 d−1 of NH4+ and NO3 treated plants, respectively). Phragmites and Glyceria had higher S/R ratio in NH4+ cultures than in NO3 cultures, 123.5 and 129.7%, respectively.Species differed in the nitrogen utilisation. In Glyceria, the relative tissue N content was higher than in Phragmites and was increased in NH4+ treated plants by 16%. The tissue NH4+ concentration (mean = 1.6 μmol g fresh wt−1) was not affected by N treatment, whereas NO3 contents were higher in NO3 (mean = 1.5 μmol g fresh wt−1) than in NH4+ (mean = 0.4 μmol g fresh wt−1) treated plants. In Phragmites, NH4+ (mean = 1.6 μmol g fresh wt−1) and NO3 (mean = 0.2 μmol g fresh wt−1) contents were not affected by the N regime. Species did not differ in NH4+ (mean = 56.5 μmol g−1 root dry wt h−1) and NO3 (mean = 34.5 μmol g−1 root dry wt h−1) maximum uptake rates (Vmax), and Vmax for NH4+ uptake was not affected by N treatment. The uptake rate of NO3 was low in NH4+ treated plants, and an induction phase for NO3 was observed in NH4+ treated Phragmites but not in Glyceria. Phragmites had low Km (mean = 4.5 μM) and high affinity (10.3 l g−1 root dry wt h−1) for both ions compared to Glyceria (Km = 6.3 μM, affinity = 8.0 l g−1 root dry wt h−1). The results showed different plasticity of Phragmites and Glyceria toward N source. The positive response to NH4+-N source may participates in the observed success of Glyceria at NH4+ rich sites, although other factors have to be considered. Higher plasticity of Phragmites toward low nutrient availability may favour this species at oligotrophic sites.  相似文献   

18.
Mutations in the second EF-hand (D61N, D63N, D65N, and E72A) of S100B were used to study its Ca2 + binding and dynamic properties in the absence and presence of a bound target, TRTK-12. With D63NS100B as an exception (D63NKD = 50 ± 9 μM), Ca2 + binding to EF2-hand mutants were reduced by more than 8-fold in the absence of TRTK-12 (D61NKD = 412 ± 67 μM, D65NKD = 968 ± 171 μM, and E72AKD = 471 ± 133 μM), when compared to wild-type protein (WTKD = 56 ± 9 μM). For the TRTK-12 complexes, the Ca2 +-binding affinity to wild type (WT + TRTKKD = 12 ± 10 μM) and the EF2 mutants was increased by 5- to 14-fold versus in the absence of target (D61N + TRTKKD = 29 ± 1.2 μM, D63N + TRTKKD = 10 ± 2.2 μM, D65N + TRTKKD = 73 ± 4.4 μM, and E72A + TRTKKD = 18 ± 3.7 μM). In addition, Rex, as measured using relaxation dispersion for side‐chain 15N resonances of Asn63 (D63NS100B), was reduced upon TRTK-12 binding when measured by NMR. Likewise, backbone motions on multiple timescales (picoseconds to milliseconds) throughout wild type, D61NS100B, D63NS100B, and D65NS100B were lowered upon binding TRTK-12. However, the X-ray structures of Ca2 +-bound (2.0 Å) and TRTK-bound (1.2 Å) D63NS100B showed no change in Ca2 + coordination; thus, these and analogous structural data for the wild-type protein could not be used to explain how target binding increased Ca2 +-binding affinity in solution. Therefore, a model for how S100B–TRTK‐12 complex formation increases Ca2 + binding is discussed, which considers changes in protein dynamics upon binding the target TRTK-12.  相似文献   

19.
This study described the synthesis and in vitro evaluation of eight new derivatives of uridine as antifungal agents and inhibitors of chitin synthase. Dimeric uridinyl derivatives synthesized by us did not exhibit significant activity. One of the studied monomeric derivative, 5′-(N-succinyl)-5′-amino-5′-deoxyuridine methyl ester (compound 7) showed activities against several fungal strains (MIC range 0.06–1.00 mg/mL) and inhibited chitin synthase from Saccharomyces cerevisiae (IC50 = 0.8 mM). Moreover compound 7 exhibited synergistic interaction with caspofungin against Candida albicans (FIC index = 0.28).  相似文献   

20.
《Aquatic Botany》2007,87(2):104-110
A large-scale mesocosm (sixteen 500 L tanks) experiment was conducted to investigate the effects of hypersalinity (45–65 psu), porewater sulfide (2–6 mM) and nighttime water column hypoxia (5–3 mg L−1) on the tropical seagrass Thalassia testudinum Banks ex König. We examined stressor effects on growth, shoot survival, tissue sulfur (S0, TS, δ34S) and leaf quantum efficiencies, as well as, porewater sulfides (∑TSpw) and mesocosm water column O2 dynamics. Sulfide was injected into intact seagrass cores of T. testudinum exposing below-ground tissues to 2, 4, and 6 mM S2−, but rapid oxidation resulted in ∑TSpw < 1.5 mM. Hypersalinity at 65 psu lowered sulfide oxidation and significantly affected plant growth rates and quantum efficiencies (Fv/Fm < 0.70). The most depleted rhizome δ34S signatures were also observed at 65 psu, suggesting increased sulfide exposure. Hypoxia did not influence ∑TSpw and plant growth, but strengthened the hypersalinity response and decreased rhizome S0, indicating less efficient oxidation of ∑TSpw. Following nighttime hypoxia treatments, ecosystem level metabolism responded to salinity treatments. When O2 levels were reduced to 5 and 4 mg L−1, daytime O2 levels recovered to approximately 6 mg L−1; however, this recovery was more limited when O2 levels were lowered to 3 mg L−1. Subsequent to O2 reductions to 3 mg O2 L−1, nighttime O2 levels rose in the 35 and 45 psu tanks, stayed the same in the 55 psu tanks, and declined in the 65 psu tanks. Thus, hypersalinity at 65 psu affects T. testudinum's oxidizing capacity and places subtle demands on the positive O2 balance at an ecosystem level. This O2 demand may influence T. testudinum die-off events, particularly after periods of high temperature and salinity. We hypothesize that the interaction between hypersalinity and sulfide toxicity in T. testudinum is their synergistic effect on the critical O2 balance of the plant.  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