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.
A dispersion of melting temperatures at pH 5.3 for individual residues of the BBL protein domain has been adduced as evidence for barrier-free downhill folding. Other members of the peripheral subunit domain family fold cooperatively at pH 7. To search for possible causes of anomalies in BBL's denaturation behavior, we measured the pH titration of individual residues by heteronuclear NMR. At 298 K, the pKa of His142 was close to that of free histidine at 6.47 ± 0.04, while that of the more buried His166 was highly perturbed at 5.39 ± 0.02. Protonation of His166 is thus energetically unfavorable and destabilizes the protein by ∼ 1.5 kcal/mol. Changes in Cα secondary shifts at pH 5.3 showed a decrease in helicity of the C-terminus of helix 2, where His166 is located, which was accompanied by a measured decrease of 1.1 ± 0.2 kcal/mol in stability from pH 7 to 5.3. Protonation of His166 perturbs, therefore, the structure of BBL. Only ∼ 1% of the structurally perturbed state will be present at the biologically relevant pH 7.6. Experiments at pH 5.3 report on a near-equal mixture of the two different native states. Further, at this pH, small changes of pH and pKa induced by changes in temperature will have near-maximal effects on pH-dependent conformational equilibria and on propagation of experimental error. Accordingly, conventional barrier-limited folding predicts some dispersion of measured thermal unfolding curves of individual residues at pH 5.3.  相似文献   

2.
α-Amylase from Sorghum bicolor, is reversibly unfolded by chemical denaturants at pH 7.0 in 50 mM Hepes containing 13.6 mM calcium and 15 mM DTT. The isothermal equilibrium unfolding at 27 °C is characterized by two state transition with ΔG (H2O) of 16.5 kJ mol−1 and 22 kJ mol−1, respectively, at pH 4.8 and pH 7.0 for GuHCl and ΔG (H2O) of 25.2 kJ mol−1 at pH 4.8 for urea. The conformational stability indicators such as the change in excess heat capacity (ΔCp), the unfolding enthalpy (Hg) and the temperature at ΔG = 0 (Tg) are 17.9 ± 0.7 kJ mol−1 K−1, 501.2 ± 18.2 kJ mol1 and 337.3 ± 6.9 K at pH 4.8 and 14.3 ± 0.5 kJ mol−1 K−1, 509.3 ± 21.7 kJ mol−1 and 345.4 ± 4.8 K at pH 7.0, respectively. The reactivity of the conserved cysteine residues, during unfolding, indicates that unfolding starts from the ‘B’ domain of the enzyme. The oxidation of cysteine residues, during unfolding, can be prevented by the addition of DTT. The conserved cysteine residues are essential for enzyme activity but not for the secondary and tertiary fold acquired during refolding of the denatured enzyme. The pH dependent stability described by ΔG (H2O) and the effect of salt on urea induced unfolding confirm the role of electrostatic interactions in enzyme stability.  相似文献   

3.
Saccharomyces cerevisiae Atm1p has been cloned, over-expressed and purified from a yeast expression system. The sequence includes both the soluble ATPase and transmembrane-spanning domains. With the introduction of an N-terminal Kozak sequence and a C-terminal (His)6-tag, a yield of 1 mg of Atm1p was obtained from 3 g wet yeast cells, which is comparable to other membrane-associated proteins isolated from eukaryotic expression systems. The ATPase activity of Atm1p is sensitive to sodium vanadate, a P-type ATPase inhibitor, with an IC50 of 4 μM. MgADP is a product inhibitor for Atm1p with an IC50 of 0.9 mM. The Michaelis–Menten constants Vmax, KM and kcat of Atm1p were measured as 8.7 ± 0.3 μM/min, 107 ± 16 μM and 1.24 ± 0.06 min− 1, respectively. A plot of ATPase activity versus concentration of Atm1p exhibits a nonlinear relationship, suggesting an allosteric response and an important role for the transmembrane domain in mediating both ATP hydrolysis and MgADP release. The metal dependence of Atm1p ATPase activity demonstrated a reactivity order of Mg2+ > Mn2+ > Co2+, while each divalent ion was found to be inhibitory at higher concentrations. The activation and inhibitory effect of phospholipids suggest that formation of a lipid–micelle complex is important for enzymatic activity and stability. Structural analysis of Atm1p by CD spectroscopy suggested a similarity of secondary structure to that found for other members of this ABC protein family.  相似文献   

4.
To understand the role of His and Glu in the catalytic activity of Bacillus licheniformis α-amylase (BLA), His235 was replaced with Glu. The mutant enzyme, H235E, was characterized in terms of its mode of action using labeled and unlabeled maltooctaose (Glc8). H235E predominantly produced maltotridecaose (Glc13) from Glc8, exhibiting high substrate transglycosylation activity, with Km = 0.38 mM and kcat/Km = 20.58 mM−1 s−1 for hydrolysis, and Km2 = 18.38 mM and kcat2/Km2 = 2.57 mM−1 s−1 for transglycosylation, while the wild-type BLA exhibited high hydrolysis activity exclusively. Glu235—located on a wide open groove near subsite +1—is likely involved in transglycosylation via formation of an α-1,4-glycosidic linkage and may recognize and stabilize the non-reducing end glucose of the acceptor molecule.  相似文献   

5.
Cytochromes c6 and f react by three et mechanisms under similar conditions. We report temperature and viscosity effects on the protein docking and kinetics of 3Zncyt c6 + cyt f(III) → Zncyt c6+ + cyt f(II). At 0.5-40.0 °C, this reaction occurs within the persistent (associated) diprotein complex with the rate constant kpr and within the transient (collision) complex with the rate constant ktr. The viscosity independence of kpr, the donor-acceptor coupling Hab = (0.5 ± 0.1) cm−1, and reorganizational energy λ = (2.14 ± 0.02) eV indicate true et within the persistent complex. The viscosity dependence of ktr and a break at 30 °C in the Eyring plot for ktr reveal mechanisms within the transient complex that are reversibly switched by temperature change. Kramers protein friction parameters differ much for the reactions below (σ = 0.3 ± 0.1, δ = 0.85 ± 0.07) and above (σ = 4.0 ± 0.9, δ = 0.40 ± 0.06) 30 °C. The transient complex(es) undergo(es) coupled et below ca. 30 °C and gated et above ca. 30 °C. Brownian dynamics simulations reveal two broad, dynamic ensembles of configurations “bridged” by few intermediate configurations through which the interconversion presumably occurs.  相似文献   

6.
Low concentrations of urea and GuHCl (2 M) enhanced the activity of endoglucanase (EC 3.1.2.4) from Aspergillus aculeatus by 2.3- and 1.9-fold, respectively. The Km values for controls, in the presence of 2 M urea and GuHCl, were found to be 2.4 ± 0.2 × 10−8 mol L−1, 1.4 ± 0.2 × 10−8 mol L−1, and 1.6 ± 0.2 × 10−8 mol L−1, respectively. The dissociation constant (Kd) showed changes in the affinity of the enzyme for the substrate with increases in the Kcat suggesting an increased turnover number in the presence of urea and GuHCl. Fluorescence studies showed changes in the microenvironment of the protein. The increase in the activity of this intermediate state was due to conformational changes accompanied by increased flexibility at the active site.  相似文献   

7.
The ocean is a nutritionally heterogeneous environment. For feeding larval forms, food variability has significant consequences for growth and later recruitment success. In this study, the physiological and biochemical responses to a range of different food concentrations (unfed, 4, 20, and 40 algal cells μl− 1) were examined in larvae of the asteroid, Asterina miniata. Measurements of growth, protein synthesis rates, and the energetic cost of protein synthesis were made. Under conditions of rapid growth, protein comprised a larger percent (66%) of a larva's organic biomass compared to similar-aged, slower-growing larvae (26%). Larvae fed at the highest food concentration tested (40 algal cells μl− 1) had a protein depositional efficiency of 80% (± 16%), a value 3-fold higher than larvae fed 20 algal cells μl− 1 (28% ± 11%). Also, faster-growing larvae required 3-fold less energy per unit mass of protein growth. Larvae fed 40 algal cells μl− 1 deposited protein at a respiratory cost of 65 ± 11 pmol O2 h− 1 (μg protein)− 1; larvae fed 20 algal cells μl− 1 had a cost of 192 ± 47 pmol O2 h− 1 (μg protein)− 1. While there were differences in the cost to deposit protein (i.e., protein growth, the balance of synthesis and degradation), there were no differences in the energetic cost of protein synthesis for all food concentrations tested. The energetic cost of protein synthesis was fixed at 13.8 (± 0.92) Joules (mg protein synthesized)− 1 and was independent of developmental stage, growth rates, and large changes (58-fold) in protein synthesis rates. A major conclusion from this study is that larvae grown in high-food environments not only grew faster, but did so for considerably less energy. Defining the complex relationships of food availability and metabolic efficiency will provide more accurate predictions of larval growth under variable food conditions in the ocean.  相似文献   

8.
DNA bending can be promoted by reducing the net negative electrostatic potential around phosphates on one face of the DNA, such that electrostatic repulsion among phosphates on the opposite face drives bending toward the less negative surface. To provide the first assessment of energetic contribution to DNA bending when electrostatic asymmetry is induced by a site-specific DNA binding protein, we manipulated the electrostatics in the EcoRV endonuclease-DNA complex by mutation of cationic side chains that contact DNA phosphates and/or by replacement of a selected phosphate in each strand with uncharged methylphosphonate. Reducing the net negative charge at two symmetrically located phosphates on the concave DNA face contributes − 2.3 kcal mol 1 to − 0.9 kcal mol 1 (depending on position) to complex formation. In contrast, reducing negative charge on the opposing convex face produces a penalty of + 1.3 kcal mol 1. Förster resonance energy transfer experiments show that the extent of axial DNA bending (about 50°) is little affected in modified complexes, implying that modification affects the energetic cost but not the extent of DNA bending. Kinetic studies show that the favorable effects of induced electrostatic asymmetry on equilibrium binding derive primarily from a reduced rate of complex dissociation, suggesting stabilization of the specific complex between protein and markedly bent DNA. A smaller increase in the association rate may suggest that the DNA in the initial encounter complex is mildly bent. The data imply that protein-induced electrostatic asymmetry makes a significant contribution to DNA bending but is not itself sufficient to drive full bending in the specific EcoRV-DNA complex.  相似文献   

9.
10.
Sublevel structure of the 8S7/2 electronic ground state of anionic bis(phthalocyaninato)gadolinium(III) has been determined by simulation analysis of an ESR spectrum in frozen solution. The simplex multidimensional minimization algorithm was employed to find the zero-field-splitting parameter set giving the minimum RMS error from the observed spectrum. The parameter set {B20,B40,B60} for the potential of D4d symmetry has been determined to be ±{(1.54 ± 0.01)× 10−2 cm−1, (0.9 ± 0.1)× 10−4 cm−1, (−0.6 ± 0.9)× 10−6 cm−1}. The energy difference between the lowest and highest sublevels has been found to be about 0.5 cm−1.  相似文献   

11.
The most extensively studied ficins have been isolated from the latex of Ficus glabrata and Ficus carica. However the proteases (ficins) from other species are less known. The purification and characterization of a protease from the latex of Ficus racemosa is reported. The enzyme purified to homogeneity is a single polypeptide chain of molecular weight of 44,500 ± 500 Da as determined by MALDI-TOF. The enzyme exhibited a broad spectrum of pH optima between pH 4.5-6.5 and showed maximum activity at 60 ± 0.5 °C. The enzyme activity was completely inhibited by pepstatin-A indicating that the purified enzyme is an aspartic protease. Far-UV circular dichroic spectra revealed that the purified enzyme contains predominantly β-structures. The purified protease is thermostable. The apparent Tm, (mid point of thermal inactivation) was found to be 70 ± 0.5 °C. Thermal inactivation was found to follow first order kinetics at pH 5.5. Activation energy (Ea) was found to be 44.0 ± 0.3 kcal mol−1. The activation enthalpy (ΔH), free energy change (ΔG) and entropy (ΔS) were estimated to be 43 ± 4 kcal mol−1, −26 ± 3 kcal mol−1 and 204 ± 10 cal mol−1 K−1, respectively. Its enzymatic specificity studied using oxidized B chain of insulin indicates that the protease preferably hydrolyzed peptide bonds C-terminal to glutamate, leucine and phenylalanine (at P1 position). The broad specificity, pH optima and elevated thermal stability indicate the protease is distinct from other known ficins and would find applications in many sectors for its unique properties.  相似文献   

12.
Poliovirus RNA replication is directed by a replication complex on the rosette-like arrangement of membranous vesicles. Proteins derived from the p3 region of the polioviral genome, such as 3D, 3AB, and 3B (VPg), play key roles in the formation and function of the replication complex. In the present study, by using an acceptor photobleaching protocol for fluorescence resonance energy transfer (FRET) imaging, we visualized the interactions of 3D, 3AB, and VPg in living cells. The interaction of 3AB-VPg was determined by live cell FRET analysis. Quantitative analyses showed that the FRET efficiencies of 3AB-3D, VPg-3D, and 3AB-VPg were 3.9 ± 0.4% (n = 36), 4.5 ± 0.4% (n = 39), and 8.3 ± 0.6% (n = 44), respectively, in the cell cytoplasm where viral replication complexes are formed and function. Poliovirus infection enhanced the protein interactions of VPg-3D and 3AB-3D, with FRET efficiencies in the virus-infected cells of 10.7 ± 1.1% (n = 39) and 9.0 ± 0.9% (n = 37), respectively. This method of live cell analysis of protein interactions in the poliovirus RNA replication complex lays the foundation for further understanding of the real-time process of poliovirus RNA replication.  相似文献   

13.
The thermophilic bacterium Thermus aquaticus is a well-known source of Taq polymerase. Here, we studied the structure and dynamics of the T. aquaticus cold-shock protein (Ta-Csp) to better understand its thermostability using NMR spectroscopy. We found that Ta-Csp has a five-stranded β-barrel structure with five salt bridges which are important for more rigid structure and a higher melting temperature (76 °C) of Ta-Csp compared to mesophilic and psychrophilic Csps. Microsecond to millisecond time scale exchange processes occur only at the β1–β2 surface region of the nucleic acid binding site with an average conformational exchange rate constant of 674 s−1. The results imply that thermophilic Ta-Csp has a more rigid structure and may not need high structural flexibility to accommodate nucleic acids upon cold shock compared to its mesophile and psychrophile counterparts.  相似文献   

14.
Dynamics and function of proteins are governed by the structural and energetic properties of the different states they adopt and the barriers separating them. In earlier work, native-state triplet–triplet energy transfer (TTET) on the villin headpiece subdomain (HP35) revealed an equilibrium between a locked native state and an unlocked native state, which are structurally similar but have different dynamic properties. The locked state is restricted to low amplitude motions, whereas the unlocked state shows increased conformational flexibility and undergoes local unfolding reactions. This classified the unlocked state as a dry molten globule (DMG), which was proposed to represent an expanded native state with loosened side-chain interactions and a solvent-shielded core. To test whether the unlocked state of HP35 is actually expanded compared to the locked state, we performed high-pressure TTET measurements. Increasing pressure shifts the equilibrium from the locked toward the unlocked state, with a small negative reaction volume for unlocking (ΔV0 = − 1.6 ± 0.5 cm3/mol). Therefore, rather than being expanded, the unlocked state represents an alternatively packed, compact state, demonstrating that native proteins can exist in several compact folded states, an observation with implications for protein function. The transition state for unlocking/locking, in contrast, has a largely increased volume relative to the locked and unlocked state, with respective activation volumes of 7.1 ± 0.4 cm3/mol and 8.7 ± 0.9 cm3/mol, indicating an expansion of the protein during the locking/unlocking transition. The presented results demonstrate the existence of both compact, low-energy and expanded, high-energy DMGs, prompting a broader definition of this state.  相似文献   

15.
The bovine milk lipocalin, β-Lactoglobulin (β-LG), has been associated with the binding and transport of small hydrophobic and amphiphilic compounds, whereby it is proposed to increase their bioavailability. We have studied the binding of the fluorescent phospholipid-derivative, NBD-didecanoylphosphatidylethanolamine (NBD-diC10PE) to β-LG by following the increase in amphiphile fluorescence upon binding to the protein using established methods. The equilibrium association constant, KB, was (1.2 ± 0.2) × 106 M− 1 at 25 °C, pH 7.4 and I = 0.15 M. Dependence of KB on pH and on the monomer-dimer equilibrium of β-LG gave insight on the nature of the binding site which is proposed to be the hydrophobic calyx formed by the β-barrel in the protein. The monomer-dimer equilibrium of β-LG was re-assessed using fluorescence anisotropy of Tryptophan. The equilibrium constant for dimerization, KD, was (7.0 ± 1.5) × 105 M− 1 at 25 °C, pH 7.4, and 0.15 M ionic strength. The exchange of NBD-diC10PE between β-LG and POPC lipid bilayers was followed by the change in NBD fluorescence. β-LG was shown to be a catalyst of phospholipid exchange between lipid bilayers, the mechanism possibly involving adsorption of the protein at the bilayer surface.  相似文献   

16.
The factors controlling the stability, folding, and dynamics of integral membrane proteins are not fully understood. The high stability of the membrane protein bacteriorhodopsin (bR), an archetypal member of the rhodopsin photoreceptor family, has been ascribed to its covalently bound retinal cofactor. We investigate here the role of this cofactor in the thermodynamic stability and folding kinetics of bR. Multiple spectroscopic probes were used to determine the kinetics and energetics of protein folding in mixed lipid/detergent micelles in the presence and absence of retinal. The presence of retinal increases extrapolated values for the overall unfolding free energy from 6.3 ± 0.4 kcal mol− 1 to 23.4 ± 1.5 kcal mol− 1 at zero denaturant, suggesting that the cofactor contributes 17.1 kcal mol− 1 towards the overall stability of bR. In addition, the cooperativity of equilibrium unfolding curves is markedly reduced in the absence of retinal with overall m-values decreasing from 31.0 ± 2.0 kcal mol− 1 to 10.9 ± 1.0 kcal mol− 1, indicating that the folded state of the apoprotein is less compact than the equivalent for the holoprotein. This change in the denaturant response means that the difference in the unfolding free energy at a denaturant concentration midway between the two unfolding curves is only ca 3-6 kcal mol− 1. Kinetic data show that the decrease in stability upon removal of retinal is associated with an increase in the apparent intrinsic rate constant of unfolding, kuH2O, from ~1 × 10− 16 s− 1 to ~1 × 10− 4 s− 1 at 25 °C. This correlates with a decrease in the unfolding activation energy by 16.3 kcal mol− 1 in the apoprotein, extrapolated to zero SDS. These results suggest that changes in bR stability induced by retinal binding are mediated solely by changes in the activation barrier for unfolding. The results are consistent with a model in which bR is kinetically stabilized via a very slow rate of unfolding arising from protein-retinal interactions that increase the rigidity and compactness of the polypeptide chain.  相似文献   

17.
Despite the widespread presence of the globin fold in most living organisms, only eukaryotic globins have been employed as model proteins in folding/stability studies so far. This work introduces the first thermodynamic and kinetic characterization of a prokaryotic globin, that is, the apo form of the heme-binding domain of flavohemoglobin (apoHmpH) from Escherichia coli. This bacterial globin has a widely different sequence but nearly identical structure to its eukaryotic analogues. We show that apoHmpH is a well-folded monomeric protein with moderate stability at room temperature [apparent ΔG°UN(w) = − 3.1 ± 0.3 kcal mol− 1; mUN = − 1.7 kcal mol− 1 M− 1] and predominant α-helical structure. Remarkably, apoHmpH is the fastest-folding globin known to date, as it refolds about 4- to 16-fold more rapidly than its eukaryotic analogues (e.g., sperm whale apomyoglobin and soybean apoleghemoglobin), populating a compact kinetic intermediate (βI = 0.9 ± 0.2) with significant helical content. Additionally, the single Trp120 (located in the native H helix) becomes locked into a fully native-like environment within 6 ms, suggesting that this residue and its closest spatial neighbors complete their folding at ultrafast (submillisecond) speed. In summary, apoHmpH is a bacterial globin that shares the general folding scheme (i.e., a rapid burst phase followed by slower rate-determining phases) of its eukaryotic analogues but displays an overall faster folding and a kinetic intermediate with some fully native-like traits. This study supports the view that the general folding features of bacterial and eukaryotic globins are preserved through evolution while kinetic details differ.  相似文献   

18.
Production of electricity from samples obtained during anaerobic digestion of grass silage was examined using single-chamber air-cathode mediator-less microbial fuel cells (MFCs). The samples were obtained from anaerobic reactors at start-up conditions after 3 and 10 days of operation under psychrophilic (15 °C) and mesophilic (37 °C) temperatures. Electricity was directly produced from all samples at a concentration of 1500 mg COD L−1. Power density obtained from the samples, as a sole carbon source, ranged from 56 ± 3 W m−3 to 31 ± 1 W m−3 for the mesophilic and psychrophilic samples, respectively. Coulombic efficiencies ranged from 18 ± 1% to 12 ± 1% for the same samples. The relationship between the maximum voltage output and initial COD concentration appeared to follow saturation kinetics at the external resistance of 217 Ω. Chemical oxygen demand (COD) removal was over 90% and total phenolics removal was in the range of 30-75% for all samples tested, with a standard amount of 60 mg L−1 total phenolics removed for every sample. Our results indicate that generating electricity from solution samples of anaerobic reactors utilizing grass silage is possible, opening the possibility for combination of anaerobic digestion with MFC technology for energy generation.  相似文献   

19.
The bindings of biogenic polyamines spermine (spm), spermidine (spmd) and synthetic polyamines 3,7,11,15-tetrazaheptadecane·4HCl (BE-333) and 3,7,11,15,19-pentazahenicosane·5HCl (BE-3333) with β-lactoglobulin (β-LG) were determined in aqueous solution. FTIR, UV-vis, CD and fluorescence spectroscopic methods as well as molecular modeling were used to determine the polyamine binding sites and the effect of polyamine complexation on protein stability and secondary structure. Structural analysis showed that polyamines bind β-LG via both hydrophilic and hydrophobic contacts. Stronger polyamine-protein complexes formed with synthetic polyamines than biogenic polyamines, with overall binding constants of Kspm-β-LG = 3.2(±0.6) × 104 M−1, Kspmd-β-LG = 1.8(±0.5) × 104 M−1, KBE-333-β-LG = 5.8(±0.3) × 104 M−1 and KBE-3333-β-LG = 6.2(±0.05) × 104 M−1. Molecular modeling showed the participation of several amino acids in the polyamine complexes with the following order of polyamine-protein binding affinity: BE-3333 > BE-333 > spermine > spermidine, which correlates with their positively charged amino group content. Alteration of protein conformation was observed with a reduction of β-sheet from 57% (free protein) to 55-51%, and a major increase of turn structure from 13% (free protein) to ∼21% in the polyamine-β-LG complexes, indicating a partial protein unfolding.  相似文献   

20.
The Iberian Peninsula encompasses more than 80% of the species richness of European aquatic ranunculi. The floristic diversity of the phytocoenosis characterised by aquatic Ranunculus and the main physical–chemical factors of the water were studied in 43 localities of the central Iberian Peninsula. Four aquatic Ranunculus communities are found in most of the aquatic environments. These are species-poor and have an uneven distribution: three species of Batrachium are heterophyllous and their communities are distributed in different aquatic ecosystems on silicated substrates; one species is homophyllous and its community occurs in various aquatic ecosystems with carbonated waters. In the Mediterranean climate, Ranunculus species are present in different habitats, as shown by the results of all the statistical analyses. Ranunculus trichophyllus communities occur in base-rich waters with a high buffering capacity (2273.44 ± 794.57 mg CaCO3 L−1) and a high concentration of cations (Ca2+, 121 ± 33.12 mg L−1; Mg2+, 71.64 ± 82.77 mg L−1), nitrates (2.89 ± 4.80 mg L−1), ammonium (2.19 ± 1.36 mg L−1) and sulphates (216.25 ± 218.54 mg L−1). Ranunculus penicillatus communities are found in flowing waters with a high concentration of phosphates (0.48 ± 0.6 mg L−1) and intermediate buffering capacity (683.66 ± 446.76 mg CaCO3 L−1). Both Ranunculus pseudofluitans and Ranunculus peltatus communities grow in waters with low buffering capacity (R. pseudofluitans, 385.91 ± 209.2 mg CaCO3 L−1; R. peltatus, 263.3 ± 180.36 mg CaCO3 L−1), and a low concentration of cations (R. pseudofluitans: Ca2+, 12.57 ± 9.42 mg L−1; Mg2+, 3.42 ± 1.67 mg L−1; R. peltatus: Ca2+, 15 ± 18.26 mg L−1; Mg2+, 6.26 ± 8.89 mg L−1) and nutrients (R. pseudofluitans: nitrates, 0.23 ± 0.2 mg L−1; phosphates, 0.09 ± 0.1 mg L−1; R. peltatus: nitrates, 0.19 ± 0.21 mg L−1; phosphates, 0.09 ± 0.12 mg L−1); the first in flowing waters, the latter in still waters.  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