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.
Emigh TH 《Genetics》1979,92(1):339-351
The dynamics of a gene in a haploid population can be explained approximately by considering the average reproductive value of the gene. The dynamics of the average reproductive value are similar to those of a gene in a population with nonoverlapping generations with the following modifications: The effective population size, Ne, replaces N; the average mutation rates µ* and ν* replace µ and ν; the average overall selection r*+(T-1)s** replaces s; and time is measured in terms of generations, T. The implications of the average selection coefficient to adaptive life histories are discussed.  相似文献   

2.
The kinetics of accumulation (per milliliter of culture) of the α- and β- subunits, associated with chloroplast-localized ammonium inducible nicotinamide adenine dinucleotide phosphate-specific glutamate dehydrogenase (NADP-GDH) isoenzymes, were measured during a 3 hour induction of synchronized daughter cells of Chlorella sorokiniana in 29 millimolar ammonium medium under photoautotrophic conditions. The β-subunit holoenzyme(s) accumulated in a linear manner for 3 hours without an apparent induction lag. A 40 minute induction lag preceded the accumulation of the α-subunit holoenzyme(s). After 120 minutes, the α-subunit ceased accumulating and thereafter remained at a constant level (i.e. steady state between synthesis and degradation). From pulsechase experiments, using 35SO4 and immunochemical procedures, the rate of synthesis of the α-subunit was shown to be greater than the β-subunit during the first 80 minutes of induction. The α- and β-subunits had different rates of degradation during the induction period (t½ = 50 versus 150 minutes, respectively) and during the deinduction period (t½ = 5 versus 13.5 minutes) after removal of ammonium from the culture. During deinduction, total NADP-GDH activity decreased with a half-time of 9 minutes. Cycloheximide completely inhibited the synthesis and degradation of both subunits. A model for regulation of expression of the NADP-GDH gene was proposed.  相似文献   

3.
4.
A phenomenological theory for contracting muscle based on irreversible thermodynamics and the sliding filament theory is developed. The individual cross bridges, considered as subunits, are viewed as linear energy converters with constant transport coefficients. With this view of the subunits, phenomenological equations applicable to the whole muscle are obtained. The transport coefficients are shown to be a function of a single parameter which is the number of activated cross bridges at any instant. By requiring Hill's force-velocity relation (1) to be satisfied, the response of the muscle is related to the number of activated cross bridges. The resulting theory differs significantly from the theory developed by Caplan (2) and a comparison of the theories is presented. The theory is shown to correlate well with the heat data of Woledge (3) for a tortoise muscle and gives a value of Y (ratio of chemical affinity to enthalpy of reaction) equal to 0.945. The comparison of the theory with Hill's frog muscle data (1) and (4) is also encouraging. In part II of this series, length variations are considered and the resulting theoretical predictions are shown to be consistent with experimental data.  相似文献   

5.
DNA polymerase ν (pol ν) is a low fidelity A-family polymerase with a putative role in interstrand cross-link repair and homologous recombination. We carried out pre-steady-state kinetic analysis to elucidate the kinetic mechanism of this enzyme. We found that the mechanism consists of seven steps, similar that of other A-family polymerases. pol ν binds to DNA with a Kd for DNA of 9.2 nm, with an off-rate constant of 0.013 s−1and an on-rate constant of 14 μm−1 s−1. dNTP binding is rapid with Kd values of 20 and 476 μm for the correct and incorrect dNTP, respectively. Pyrophosphorylation occurs with a Kd value for PPi of 3.7 mm and a maximal rate constant of 11 s−1. Pre-steady-state kinetics, examination of the elemental effect using dNTPαS, and pulse-chase experiments indicate that a rapid phosphodiester bond formation step is flanked by slow conformational changes for both correct and incorrect base pair formation. These experiments in combination with computer simulations indicate that the first conformational change occurs with rate constants of 75 and 20 s−1; rapid phosphodiester bond formation occurs with a Keq of 2.2 and 1.7, and the second conformational change occurs with rate constants of 2.1 and 0.5 s−1, for correct and incorrect base pair formation, respectively. The presence of a mispair does not induce the polymerase to adopt a low catalytic conformation. pol ν catalyzes both correct and mispair formation with high catalytic efficiency.  相似文献   

6.
The depolarization ratios of heme protein Raman lines arising from vibrations of the heme group exhibit significant dependence on the excitation wavelength. From the analysis of this depolarization ratio dispersion, one obtains information about symmetry-lowering distortions δQΓ of the heme group that can be classified in terms of the symmetry races Γ = A1g, B1g, B2g, and A2g in D4h symmetry. The heme-protein interaction can be changed by the protonation of distinct amino acid side chains (i.e., for instance the Bohr groups in hemoglobin derivates), which gives rise to specific static heme distortions for each protonation state. From the Raman dispersion data, it is possible to obtain parameters by fitting to a theoretical expression of the Raman tensor, which provide information on these static distortions and also about the pK values of the involved titrable side chains. We have applied this method to the ν4 (1,355 cm-1) and ν10 (1,620 cm-1) lines of deoxygenated hemoglobin of the fourth component of trout and have measured their depolarization ratio dispersion as a function of pH between 6 and 9. From the pH dependence of the thus derived parameters, we obtain pK values identical to those of the Bohr groups, which were earlier derived from the corresponding O2-binding isotherms. These are pKα1 = pKα2 = 8.5 for the α and pKβ1 = 7.5, pKβ2 = 7.4 for the β chains. We also obtain the specific distortion parameters for each protonation state. As shown in earlier studies, the ν4 mode mainly probes distortions from interactions between the proximal histidine and atoms of the heme core (i.e., the nitrogens and the Cα atoms of the pyrroles). Group theoretical argumentation allows us to relate specific changes of the imidazole geometry as determined by its tilt and azimuthal angle and the iron-out-of-plane displacement to distinct variations of the normal distortions δQΓ derived from the Raman dispersion data. Thus, we found that the pH dependence of the heme distortions δQA1g (totally symmetric) and δQB1g (asymmetric) is caused by variations of the azimuthal rather than the tilt angle of the Fe-His (F8) bond. In contrast to this, the ν10 line mainly monitors changes resulting from the interaction between peripheral substituents of the porphyrin macrocycle (vinyl). From the pH dependence of the parameters, it is possible to separately identify distortions δQΓ affecting the hemes in the α and β chains, respectively. From this, we find that in the α subunit structural changes induced on protonation of the corresponding Bohr groups are mainly transferred via the Fe—Nε bond and give rise to changes in the azimuthal angle. In the β subunit, however, in addition, structural changes of the heme pocket arise, which most probably result from protonation of the imidazole of the COOH-terminal His (HC3 β). This rearranges the net of H bonds between His HC3 β, Ser (F9 β), and Glu (F7 β).  相似文献   

7.
Two sets of iso-1-cytochrome c variants have been prepared with N-terminal insertions of pure polyglutamine, i.e., PolyQ variants, or polyglutamine interrupted with lysine every sixth residue, i.e., Gln-rich variants. The polymer properties of these pure polyGln or Gln-rich sequences have been evaluated using equilibrium and kinetic His-heme loop formation methods for loop sizes ranging from 22 to 46 in 1.5, 3.0, and 6.0 M guanidine hydrochloride (GdnHCl). In 6.0 M GdnHCl, the scaling exponent, ν3, for the pure polyGln sequences, is ∼1.7—significantly less than ν3 ≈ 2.15 for the Gln-rich sequences. The stability of the His-heme loops becomes progressively greater for the pure polyGln sequences relative to the Gln-rich sequences as GdnHCl concentration decreases from 6.0 to 1.5 M. Thus, the context of the sequence effects the polymer properties of Gln repeats even in denaturing concentrations of GdnHCl. Comparison of data for the Gln-rich variants with previous results for Gly-rich and Ala-rich variants shows that ν3 ∼ 2.2 for the Gln-rich, Gly-rich, and Ala-rich sequences in 6.0 M GdnHCl, whereas ν3 remains unchanged at 3.0 M GdnHCl concentration for the Gln-rich and Ala-rich sequences but decreases to ∼1.7 for the Gly-rich sequences. Thus, the polymer properties of Gln-rich and Ala-rich sequences are less sensitive to solvent quality in denaturing solutions of GdnHCl than Gly-rich sequences. Evaluation of Flory’s characteristic ratio, Cn, for the Gln-rich and Ala-rich sequences relative to the Gly-rich sequences shows that Gln-rich sequences are stiffer than Ala-rich sequences at both 3.0 and 6.0 M GdnHCl.  相似文献   

8.
9.
The elastic behavior of the cell wall as a function of the temperature has been studied with particular attention being given to the swelling of egg cells of Strongylocentrotus purpuratus and Crassostrea virginica in different sea water concentrations at different temperatures. It was found that the modulus of elasticity is a nonlinear function of temperature. At about 12-13°C the modulus of elasticity (E) is constant, independent of the stress (σ) and strain (εν) which exist at the cell wall; the membranous material follows Hooke's law, and E ≈ 3 × 107 dyn/cm2 for S. purpuratus and C. virginica. When the temperature is higher or lower than 12-13°C, the modulus of elasticity increases, and the membranous material does not follow Hooke's law, but is almost directly proportional to the stresses existing at the cell wall. On increasing the stress, the function Eσ = E(σ) approaches saturation. The corresponding stress-strain diagrams, σ = σ(εν), and the graphs, Eσ = E(σ) and Eσ = E(t) are given. The cyto-elastic phenomena at the membrane are discussed.  相似文献   

10.
Actin filaments propelled in vitro by groups of skeletal muscle myosin motors exhibit distinct phases of active sliding or arrest, whose occurrence depends on actin length (L) within a range of up to 1.0 μm. Smooth muscle myosin filaments are exponentially distributed with ≈150 nm average length in vivo—suggesting relevance of the L-dependence of myosin group kinetics. Here, we found L-dependent actin arrest and sliding in in vitro motility assays of smooth muscle myosin. We perturbed individual myosin kinetics with varying, physiological concentrations of phosphate (Pi, release associated with main power stroke) and adenosine diphosphate (ADP, release associated with minor mechanical step). Adenosine triphosphate was kept constant at physiological concentration. Increasing [Pi] lowered the fraction of time for which actin was actively sliding, reflected in reduced average sliding velocity (ν) and motile fraction (fmot, fraction of time that filaments are moving); increasing [ADP] increased the fraction of time actively sliding and reduced the velocity while sliding, reflected in reduced ν and increased fmot. We introduced specific Pi and ADP effects on individual myosin kinetics into our recently developed mathematical model of actin propulsion by myosin groups. Simulations matched our experimental observations and described the inhibition of myosin group kinetics. At low [Pi] and [ADP], actin arrest and sliding were reflected by two distinct chemical states of the myosin group. Upon [Pi] increase, the probability of the active state decreased; upon [ADP] increase, the probability of the active state increased, but the active state became increasingly similar to the arrested state.  相似文献   

11.
Strong evidence suggests that the stretching vibration of the bound oxygen can be perturbed by an accidentally degenerate porphyrin ring mode, resulting in two split frequencies. In the Co(II)(TpivPP) (pyridine) 18O2 complex, we demonstrate that the ν(18O—18O) mode, after being shifted from its ν(16O—16O) value at 1,156 cm-1, undergoes a resonance interaction with the 1,080 cm-1 porphyrin mode, giving rise to two lines at 1,067 and 1,089 cm-1. In the O2 complex of Co(II) mesoporphyrin IX-substituted sperm whale myoglobin, we observed a dramatic intensity increase at 1,132 cm-1 upon 16O218O2 substitution, which is due to the reappearance of the 1,132-cm-1 porphyrin mode after the removal of resonance conditions. A decrease in O2 binding affinity, caused by the proximal base tension, corresponds to an increase in the Co—O2 stretching frequency. The ν(Co—O2) at 527 cm-1 for the low affinity Co(II)(TpivPP)(1,2-Me2Im) O2 complex is 11 cm-1 higher than the 516-cm-1 value for the high affinity complex (with N-MeIm replacing 1,2-Me2Im). However, in the corresponding iron complexes the reverse behavior is observed, i.e., the ν(Fe—O2) decreases for the (1,2-Me2Im) complex. There is a 24-cm-1 difference in the Co—O2 stretching frequencies between Co(II)(TpivPP)(N-MeIm)O2 (at 516 cm-1) and oxy meso CoMb (at 540 cm-1), suggesting a protein induced distortion of the Co—O—O linkage. However, the values for ν(Fe—O2) are nearly identical between Fe(II)(TpivPP)(N-MeIm)O2 (at 571 cm-1) and oxy Mb (at 573 cm-1), indicating that O2 binds to myoglobin in the same manner as in the sterically unhindered “picket fence” complex. Evidence is presented that suggests the presence of two dioxygen stretching frequencies due to two different conformers in each of the N-MeIm and 1,2-Me2Im complex of oxy Co(II)(TpivPP).  相似文献   

12.
How Many Processed Pseudogenes Are Accumulated in a Gene Family?   总被引:2,自引:0,他引:2       下载免费PDF全文
James Bruce Walsh 《Genetics》1985,110(2):345-364
A simple kinetic model is developed that describes the accumulation of processed pseudogenes in a functional gene family. Insertion of new pseudogenes occurs at rate ν per gene and is countered by spontaneous deletion (at rate δ per DNA segment) of segments containing processed pseudogenes. If there are k functional genes in a gene family, the equilibrium number of processed pseudogenes is k(ν/δ), and the percentage of functional genes in the gene family at equilibrium is 1/[1 + (ν/δ)]. ν/δ values estimated for five gene families ranged from 1.7 to 15. This fairly narrow range suggests that the rates of formation and deletion of processed pseudogenes may be positively correlated for these families. If δ is sufficiently large relative to the per nucleotide mutation rate µ (δ > 20µ), processed pseudogenes will show high homology with each other, even in the absence of gene conversion between pseudogenes. We argue that formation of processed pseudogenes may share common pathways with transposable elements and retroviruses, creating the potential for correlated responses in the evolution of processed pseudogenes due to direct selection for control of transposable elements and/or retroviruses. Finally, we discuss the nature of the selective forces that may act directly or indirectly to influence the evolution of processed pseudogenes.

Anything produced by evolution is bound to be a bit of a mess—S. Brenner

  相似文献   

13.
Different strains of Thiobacillus ferrooxidans and Thiobacillus thiooxidans were used to catalyze the oxidative dissolution of iron pyrite, FeS2, in nine different coal samples. Kinetic variables and parametric factors that were determined to have a pronounced effect on the rate and extent of oxidative dissolution at a fixed Po2 were: the bacterial strain, the nitrogen/phosphorus molar ratio, the partial pressure of CO2, the coal source, and the total reactive surface area of FeS2. The overall rate of leaching, which exhibited a first-order dependence on the total surface area of FeS2, was analyzed mathematically in terms of the sum of a biochemical rate, ν1, and a chemical rate, ν2. Results of this study show that bacterial desulfurization (90 to 98%) of coal samples which are relatively high in pyritic sulfur can be achieved within a time-frame of 8 to 12 days when pulp densities are ≤20% and particle sizes are ≤74 μm. The most effective strains of T. ferrooxidans were those that were isolated from natural systems, and T. ferrooxidans ATCC 19859 was the most effective pure strain. The most effective nutrient media contained relatively low phosphate concentrations, with an optimal N/P molar ratio of 90:1. These results suggest that minimal nutrient additions may be required for a commercial desulfurization process.  相似文献   

14.
Phospholipids, the major structural components of membranes, can also have functions in regulating signaling pathways in plants under biotic and abiotic stress. The effects of adding phospholipids on the activity of stress-induced calcium dependent protein kinase (CaCDPK1) from chickpea are reported here. Both autophosphorylation as well as phosphorylation of the added substrate were enhanced specifically by phosphatidylcholine and to a lesser extent by phosphatidic acid, but not by phosphatidylethanolamine. Diacylgylerol, the neutral lipid known to activate mammalian PKC, stimulated CaCDPK1 but at higher concentrations. Increase in Vmax of the enzyme activity by these phospholipids significantly decreased the Km indicating that phospholipids enhance the affinity towards its substrate. In the absence of calcium, addition of phospholipids had no effect on the negligible activity of the enzyme. Intrinsic fluorescence intensity of the CaCDPK1 protein was quenched on adding PA and PC. Higher binding affinity was found with PC (K½ = 114 nM) compared to PA (K½ = 335 nM). We also found that the concentration of PA increased in chickpea plants under salt stress. The stimulation by PA and PC suggests regulation of CaCDPK1 by these phospholipids during stress response.  相似文献   

15.
The virulence of Staphylococcus aureus, in both human and animal hosts, is largely influenced by the acquisition of mobile genetic elements (MGEs). Most S. aureus strains carry a variety of MGEs, including three genomic islands (νSaα, νSaβ, νSaγ) that are diverse in virulence gene content but conserved within strain lineages. Although the mobilization of pathogenicity islands, phages and plasmids has been well studied, the mobilization of genomic islands is poorly understood. We previously demonstrated the mobilization of νSaβ by the adjacent temperate bacteriophage ϕSaBov from strain RF122. In this study, we demonstrate that ϕSaBov mediates the mobilization of νSaα and νSaγ, which are located remotely from ϕSaBov, mostly to recipient strains belonging to ST151. Phage DNA sequence analysis revealed that chromosomal DNA excision events from RF122 were highly specific to MGEs, suggesting sequence-specific DNA excision and packaging events rather than generalized transduction by a temperate phage. Disruption of the int gene in ϕSaBov did not affect phage DNA excision, packaging, and integration events. However, disruption of the terL gene completely abolished phage DNA packing events, suggesting that the primary function of temperate phage in the transfer of genomic islands is to allow for phage DNA packaging by TerL and that transducing phage particles are the actual vehicle for transfer. These results extend our understanding of the important role of bacteriophage in the horizontal transfer and evolution of genomic islands in S. aureus.  相似文献   

16.
The β-lactamase from Klebsiella pneumoniae E70 behaved in a similar fashion to the TEM-2 plasmid mediated enzyme on reaction with clavulanic acid. Both enzymes produced two types of enzyme–clavulanate complex, a transiently stable species (t½=4min at pH7.3 and 37°C) and irreversibly inhibited enzyme. In the initial rapid reaction (2.5min) the enzymes partitioned between the transient and irreversible complexes in the ratios 3:1 for TEM-2 β-lactamase and 1:1 for Klebsiella β-lactamase. Biphasic inactivation was observed for both enzymes and the slower second phase was rate limited by the decay of the transiently stable complex. This decay released free enzyme for further reaction with fresh clavulanic acid, the products again partitioning between transiently stable and irreversibly inhibited enzyme. This cycle continued until all the enzyme had been irreversibly inhibited. A 115 molar excess of inhibitor was required to achieve complete inactivation of TEM-2 β-lactamase. Hydrolysis of clavulanic acid with product release appeared to occur with the inhibition reaction, which explained this degree of clavulanic acid turnover. The stoichiometry of the interaction with Klebsiella β-lactamase was not examined. The penicillinase from Proteus mirabilis C889 was rapidly inhibited by low concentrations of clavulanic acid. The major product was a moderately stable complex (t½=40min at pH7.3 and 37°C); the proportion of the enzyme that was irreversibly inactivated was small. The cephalosporinase from Enterobacter cloacae P99 had low affinity for the inhibitor and only reacted with high concentrations of clavulanic acid (k=4.0m−1·s−1) to produce a relatively stable complex (t½=180min at pH7.3 and 37°C). No irreversible inactivation of this enzyme was detected. The rates of decay of the clavulanate–enzyme complexes produced in reactions with Proteus and Enterobacter enzymes were markedly increased at acid pH.  相似文献   

17.
In this paper, the fractional-order generalized Laguerre operational matrices (FGLOM) of fractional derivatives and fractional integration are derived. These operational matrices are used together with spectral tau method for solving linear fractional differential equations (FDEs) of order ν (0 < ν < 1) on the half line. An upper bound of the absolute errors is obtained for the approximate and exact solutions. Fractional-order generalized Laguerre pseudo-spectral approximation is investigated for solving nonlinear initial value problem of fractional order ν. The extension of the fractional-order generalized Laguerre pseudo-spectral method is given to solve systems of FDEs. We present the advantages of using the spectral schemes based on fractional-order generalized Laguerre functions and compare them with other methods. Several numerical examples are implemented for FDEs and systems of FDEs including linear and nonlinear terms. We demonstrate the high accuracy and the efficiency of the proposed techniques.  相似文献   

18.
To the revolution-ellipsoidal and spherical membranous shell (cell mitochondrion) are introduced the equations for the calculation of both the modulus of elasticity (Young's modulus) and the stresses, which exist at the membrane. The existing pressure difference between the inner and outer surface of the membrane is calculated in the dilution of seawater media in the osmotic steady state. The experimental results are obtained by using egg cells of the sea urchin, Strongylocentrotus purpuratus. Up to the specific volume of the egg cell (VE ≈ 35·10-8 cm3) Boyle-van't Hoff's law is valid (defined as the subelastic range) beyond that the elastic stresses exist (elastic range). For the maximum value of the stresses existing at the cell wall one obtains σ ≈ 5.5·106 dyne/cm2 and for the modulus of elasticity E = 1.0·107 dyne/cm2, which is constant when the value of relative strain εν > 15%. The breaking limit by an approximate calculation is σU ≈ 11·106 dyne/cm2. The membrane is assumed to be convoluted and its hypothetical degree of folding was calculated [unk]a = 34%. The results are compared with the values existing in the literature and other types of cells are found to have values of elasticity in the same range as values of the membrane of S. purpuratus. Both compression and cell elastometer methods are criticized and in certain cases results of these methods are considered to belong to the subelastic domain.  相似文献   

19.
Evidence shows that an abnormal deposition of amyloid beta-peptide25–35 (Aβ25–35) was the primary cause of the pathogenesis of Alzheimer’s disease (AD). And the elimination of Aβ25–35 is considered an important target for the treatment of AD. Triptolide (TP), isolated from Tripterygium wilfordii Hook.f. (TWHF), has been shown to possess a broad spectrum of biological profiles, including neurotrophic and neuroprotective effects. In our study investigating the effect and potential mechanism of triptolide on cytotoxicity of differentiated rat pheochromocytoma cell line (the PC12 cell line is often used as a neuronal developmental model) induced by Amyloid-Beta25–35 (Aβ25–35), we used 3-(4, 5-dimethylthiazol-2-yl)-2, 5- diphenyltetrazolium bromide (MTT) assay, flow cytometry, Western blot, and acridine orange staining to detect whether triptolide could inhibit Aβ25–35–induced cell apoptosis. We focused on the potential role of the autophagy pathway in Aβ25–35-treated differentiated PC12 cells. Our experiments show that cell viability is significantly decreased, and the apoptosis increased in Aβ25–35-treated differentiated PC12 cells. Meanwhile, Aβ25–35 treatment increased the expression of microtubule-associated protein light chain 3 II (LC3 II), which indicates an activation of autophagy. However, triptolide could protect differentiated PC12 cells against Aβ25–35-induced cytotoxicity and attenuate Aβ25–35-induced differentiated PC12 cell apoptosis. Triptolide could also suppress the level of autophagy. In order to assess the effect of autophagy on the protective effects of triptolide in differentiated PC12 cells treated with Aβ25–35, we used 3-Methyladenine (3-MA, an autophagy inhibitor) and rapamycin (an autophagy activator). MTT assay showed that 3-MA elevated cell viability compared with the Aβ25–35-treated group and rapamycin inhibits the protection of triptolide. These results suggest that triptolide will repair the neurological damage in AD caused by deposition of Aβ25–35 via the autophagy pathway, all of which may provide an exciting view of the potential application of triptolide or TWHF as a future research for AD.  相似文献   

20.
Botryococcus braunii, B race is a unique green microalga that produces large amounts of liquid hydrocarbons known as botryococcenes that can be used as a fuel for internal combustion engines. The simplest botryococcene (C30) is metabolized by methylation to give intermediates of C31, C32, C33, and C34, with C34 being the predominant botryococcene in some strains. In the present work we have used Raman spectroscopy to characterize the structure of botryococcenes in an attempt to identify and localize botryococcenes within B. braunii cells. The spectral region from 1600–1700 cm−1 showed ν(C=C) stretching bands specific for botryococcenes. Distinct botryococcene Raman bands at 1640 and 1647 cm−1 were assigned to the stretching of the C=C bond in the botryococcene branch and the exomethylene C=C bonds produced by the methylations, respectively. A Raman band at 1670 cm−1 was assigned to the backbone C=C bond stretching. Density function theory calculations were used to determine the Raman spectra of all botryococcenes to compare computed theoretical values with those observed. The analysis showed that the ν(C=C) stretching bands at 1647 and 1670 cm−1 are actually composed of several closely spaced bands arising from the six individual C=C bonds in the molecule. We also used confocal Raman microspectroscopy to map the presence and location of methylated botryococcenes within a colony of B. braunii cells based on the methylation-specific 1647 cm−1 botryococcene Raman shift.  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