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 125 毫秒
1.
《Phytochemistry》1986,26(1):129-134
The B race of the green alga Botryococcus braunii is characterized by the production of large amounts of botryococcenes, i.e. triterpenoid hydrocarbons of general formula CπH2π-109 n= 30–37. The axenic strain used in this work produces botryococcenes ranging from C30 to C34 when fast growth is promoted by air-lift. Sequential extraction of hydrocarbons with solvents showed that botryococcenes accumulate in two distinct sites: externally in the successive outer walls forming a dense matrix and internally, probably in cyctoplasmic inclusions. Moreover, chase experiments after feeding the algae with sodium [1,2-14C]acetate, and feeding experiments with L-[Me-14C]methionine established the existence of an excretory process from the cells towards the matrix. The results of the radio GC analyses of the botryococcenes synthesized during the feeding experiments provided good evidence to show that the C30 botryococcene is the precursor of all the higher hydrocarbons, and that each intermediate botryococcene C31-C33 is the precursor of its next highest homologue. L-Methionine acts as the methyl donor in the methylation process, leading from the C30 precursor to the botryococcene family. The 13C NMR spectra of the botryococcenes produced when the algae were fed with L-[Me-13C]methionine indicate that the methylation takes place on the C30 backbone in positions 37, 16 and 20.  相似文献   

2.
A ‘resting state’ isolate of the hydrocarbon-producing alga Botryococcus braunii photoassimilated sodium [14C]bicarbonate at rates comparable to fast growing algae, such as Chlorella (> 1.50 μg atoms 14C/mg chlorophyll·hr). Early in the reaction (up to several min), most of the radioactivity was associated with water-soluble metabolites. However, labelling of hexane-soluble compounds steadily from ca 3% at 15 sec to over 50% of the total incorporated 14C at 60 min. The purified hexane fraction, which consisted of a series of botryococcenes and squalene, constituted a relatively constant proportion (40–45%) of the total hexane-soluble radioactivity at all but the earliest time points (< 60 sec). This fraction initially consisted almost exclusively of a C30 botryococcene (ca 91%) and squalene (ca 8%); however, small amounts of radioactivity sequentially appeared in the C31, C32 and C34 botryococcenes. The results of pulse-chase experiments implicated the C30 botryococcene as the precursor of the higher homologues; during the chase, loss of radioactivity from the C30 compound was accompanied by a concomitant increase in the labelling of the C31 and C32 compounds. This study provides further evidence that the relatively slow growth of Botryococcus in culture may result, in part, from the diversion of a large proportion of reduced carbon into energetically expensive compounds and that the slower growth rate in the ‘resting state’ cannot be totally attributed to an impaired or intrinsically slow metabolism.  相似文献   

3.
Nine branched hydrocarbons of the botryococcene type (CnH2n-10 30 ? n ? 37) have been isolated from the green alga Botryococcus braunii. Hydrocarbon mixtures were recovered from wild algae collected in fresh water lakes or from the same strains growing in laboratory; they were further separated by reversed-phase, and in some cases by normal phase, HPLC. From chemical investigations, GC/MS analyses, 1H and 13C NMR spectroscopy, the structures of four new botryococcenes (one C33H56, two C34H58 and one C37H64) were elucidated.  相似文献   

4.
When administered in serum, an efficacious therapeutic antibody should be homogeneous to minimize immune reactions or injection site irritation during administration. Monoclonal antibody (mAb) phase separation is one type of inhomogeneity observed in serum, and thus screening potential phase separation of mAbs in serum could guide lead optimization. However, serum contains numerous components, making it difficult to resolve mAb/serum mixtures at a scale amenable to analysis in a discovery setting. To address these challenges, a miniaturized assay was developed that combined confocal microscopy with Raman spectroscopy. The method was examined using CNTO607, a poorly-soluble anti-interleukin-13 human mAb, and CNTO3930, a soluble anti-respiratory syncytial virus humanized mAb. When CNTO607 was diluted into serum above 4.5 mg/mL, phase separation occurred, resulting in droplet formation. Raman spectra of droplet phases in mixtures included bands at 1240 and 1670 cm−1, which are typical of mAb β-sheets, and lacked bands at 1270 and 1655 cm−1, which are typical of α-helices. The continuous phases included bands at 1270 and 1655 cm−1 and lacked those at 1240 and 1670 cm−1. Therefore, CNTO607 appeared to be sequestered within the droplets, while albumin and other α-helix-forming serum proteins remained within the continuous phases. In contrast, CNTO3930 formed only one phase, and its Raman spectra contained bands at 1240, 1670, 1270 and 1655 cm,−1 demonstrating homogeneous distribution of components. Our results indicate that this plate-based method utilizing confocal Raman spectroscopy to probe liquid-liquid phases in mAb/serum mixtures can provide a screen for phase separation of mAb candidates in a discovery setting.  相似文献   

5.
We have obtained by Fourier transformed infra-red (FTIR)-spectroscopy BR-K, BR-L and BR-M difference spectra of bacteriorhodopsin regenerated with isotopically labelled retinals. Thereby, we are able to assign reliably the C14–C15 and C=N stretching vibrations of the various intermediates. The lower C14–C15 stretching vibration frequency in L as compared with 13-cis protonated Schiff base model compounds indicates a 13-cis, 14-s-cis configuration of the retinal in this species. The unusually low C=N stretching vibration in K at 1615 cm−1 indicates less stabilization of the positive charge at the Schiff base by the protein environment. Based on these results, a mechanism is suggested by which the stored light energy is transformed into proton transfers.  相似文献   

6.
A strain Botryococcus braunii Kütz. that produces high levels of branched hydrocarbons (botryococcenes) was grown under different environmental conditions to investigate the relationship between growth and hydrocarbon production. Carbon dioxide concentration had the most significant influence on growth; 0.3% CO2-enriched cultures demonstrated a minimum mass doubling time of ca. 40 h, compared to ca. 6 days for ambient air cultures grown on the same buffered growth medium. The botryococcene fraction, which consisted of 10 identified compounds (CnH2n-10; n = 30–34), usually constituted ca. 25–40% of the culture dry weight under various growth regimes, including nitrogen- and/or phosphate-deficiencies. CO2 enrichment initially favored the production of the lower botryococcenes (C30–C32), whereas relatively slow-growing ambient air cultures accumulated C33 and C34 compounds. Colony color changed in response to different light intensities. High light increased the carotenoid/chlorophyll ratio, which resulted in orange colonies. Cultures exposed to low light intensity appeared green. This change in coloration was reversible over a period of a few days, and at no time were the linear hydrocarbons characteristic of the other form of the alga detected. Ostensible colony color is not, therefore, a reliable indicator of qualitative hydrocarbon content. Sequential solvent extraction experiments indicated that up to ca. 7% of the botryococcene fraction was intracellular and that the remainder was located within the colonial matrix. The internal (cellular) pool principally consisted of C30–C32 botryococcenes, whereas the external (colonial matrix) pool contained >99% of the C33 and C34 compounds, in addition to large amounts of the lower botryococcenes. These results, taken in conjunction with other data, are compatible with the hypothesis that the C30 botryococcene is the precursor, presumably via methylation, of the higher botryococcenes.  相似文献   

7.
The extracellular matrix of the alga Botryococcus braunii, Race B, consists mainly of botryococcenes, which have potential as a hydrocarbon fuel. Botryococcenes are structurally similar to squalene raising the possibility of a common enzyme for the biosynthesis of both. While B. braunii squalene synthase (SS) enzyme activity has been documented, botryococcene synthase (BS) enzyme activity has not been. In the current study, an assay for BS activity has been developed and used to show that many of the assay conditions for BS enzyme activity are similar to those of SS. However, SS enzyme activity is stimulated by Tween 80 while BS enzyme activity is inhibited. Moreover, BS enzyme activity was correlated with the accumulation of botryococcenes during a B. braunii culture growth cycle, which was distinctly different from the profile of SS enzyme activity. While the current results indicate a conservation of enzymological features amongst the BS and SS enzymes, raising the possibility of one enzyme capable of catalyzing both activities, they are also consistent with these two activities arising from separate and distinct enzymes.  相似文献   

8.
Samples of the green colonial alga Botryococcus braunii, collected from various localities, were grown in the laboratory and examined for their hydrocarbon content and morphology. Although few differences appeared between the ultrastructures of the samples, the nature of their hydrocarbons, which remains unchanged at any stage of growth, allows the distinction of two physiological races viz algae producing odd-numbered unbranched alkadienes and trienes (C25C31) (the A race) and those producing polymethylated triterpenes CnH2n-10 (C30C37), the botryococcenes (the B race). In laboratory culture, the hydrocarbon content of these new strains is very high, from 30 to 60% of the dry biomass. For the two races the greatest hydrocarbon productivity takes place during the active growth phase. The important variability observed in botryococcene distribution could originate both from genetic and environmental factors.  相似文献   

9.
There is an urgent need for rapid and reliable methods able to detect melamine in animal feed. In this study, a quick, simple, and sensitive method for the determination of melamine content in animal feed was developed using surface-enhanced Raman spectroscopy on fabricated Ag nanorod (AgNR) array substrates with a one-step sample extraction procedure. The AgNR array substrates washed by HNO3 solvent (10−7 M) and methanol and showed the good stability within 6 months. The Raman shift at △ν = 682 cm−1 was used as the characteristic melamine peak in the calculations. Sufficient linearity was obtained in the 2–200 μg·g−1 range (R2 = 0.926). The limits of detection and quantification were 0.9 and 2 μg·g−1, respectively. The recovery rates were 89.7–93.3%, with coefficients of variation below 2.02%. The method showed good accuracy compared with the tradition GC-MS analysis. This new protocol only need 2 min to fininsh the detection which could be developed for rapid onsite screening of melamine contamination in quality control and market surveillance applications.  相似文献   

10.
Identification of the intermediates and determination of their structures in the reduction of dioxygen to water by cytochrome c oxidase (CcO) are particularly important to understanding both O2 activation and proton pumping by the enzyme. In this work, we report the products of the rapid reaction of O2 with the mixed valence form (CuA2+, heme a3+, heme a32+-CuB1+) of the enzyme. The resonance Raman results show the formation of two ferryl-oxo species with characteristic Fe(IV)=O stretching modes at 790 and 804 cm−1 at the peroxy oxidation level (PM). Density functional theory calculations show that the protein environment of the proximal H-bonded His-411 determines the strength of the distal Fe(IV)=O bond. In contrast to previous proposals, the PM intermediate is also formed in the reaction of Y167F with O2. These results suggest that in the fully reduced enzyme, the proton pumping νFe(IV)=O = 804 cm−1 to νFe(IV)=O = 790 cm−1 transition (P→F, where P is peroxy and F is ferryl) is triggered not only by electron transfer from heme a to heme a3 but also by the formation of the H-bonded form of the His-411-Fe(IV)=O conformer in the proximal site of heme a3. The implications of these results with respect to the role of an O=Fe(IV)-His-411-H-bonded form to the ring A propionate of heme a3-Asp-399-H2O site and, thus, to the exit/output proton channel (H2O) pool during the proton pumping P→F transition are discussed. We propose that the environment proximal to the heme a3 controls the spectroscopic properties of the ferryl intermediates in cytochrome oxidases.  相似文献   

11.
Individual leaves of potato (Solanum tuberosum L. W729R), a C3 plant, were subjected to various irradiances (400-700 nm), CO2 levels, and temperatures in a controlled-environment chamber. As irradiance increased, stomatal and mesophyll resistance exerted a strong and some-what paralleled regulation of photosynthesis as both showed a similar decrease reaching a minimum at about 85 neinsteins·cm−2·sec−1 (about ½ of full sunlight). Also, there was a proportional hyperbolic increase in transpiration and photosynthesis with increasing irradiance up to 85 neinsteins·cm−2·sec−1. These results contrast with many C3 plants that have a near full opening of stomata at much less light than is required for saturation of photosynthesis.  相似文献   

12.
The competition for glucose between Escherichia coli ML30, a typical copiotrophic enterobacterium and Chelatobacter heintzii ATCC29600, an environmentally successful strain, was studied in a carbon-limited culture at low dilution rates. First, as a base for modelling, the kinetic parameters μmax and Ks were determined for growth with glucose. For both strains, μmax was determined in batch culture after different precultivation conditions. In the case of C. heintzii, μmax was virtually independent of precultivation conditions. When inoculated into a glucose-excess batch culture medium from a glucose-limited chemostat run at a dilution rate of 0.075 h−1 C. heintzii grew immediately with a μmax of 0.17±0.03 h−1. After five transfers in batch culture, μmax had increased only slightly to 0.18±0.03 h−1. A different pattern was observed in the case of E. coli. Inoculated from a glucose-limited chemostat at D=0.075 h−1 into glucose-excess batch medium E. coli grew only after an acceleration phase of ∼3.5 h with a μmax of 0.52 h−1. After 120 generations and several transfers into fresh medium, μmax had increased to 0.80±0.03 h−1. For long-term adapted chemostat-cultivated cells, a Ks for glucose of 15 μg l−1 for C. heintzii, and of 35 μg l−1 for E. coli, respectively, was determined in 14C-labelled glucose uptake experiments. In competition experiments, the population dynamics of the mixed culture was determined using specific surface antibodies against C. heintzii and a specific 16S rRNA probe for E. coli. C. heintzii outcompeted E. coli in glucose-limited continuous culture at the low dilution rates of 0.05 and 0.075 h−1. Using the determined pure culture parameter values for Ks and μmax, it was only possible to simulate the population dynamics during competition with an extended form of the Monod model, which includes a finite substrate concentration at zero growth rate (smin). The values estimated for smin were dependent on growth rate; at D=0.05 h−1, it was 12.6 and 0 μg l−1 for E. coli and C. heintzii, respectively. To fit the data at D=0.075 h−1, smin for E. coli had to be raised to 34.9 μg l−1 whereas smin for C. heintzii remained zero. The results of the mathematical simulation suggest that it is not so much the higher Ks value, which is responsible for the unsuccessful competition of E. coli at low residual glucose concentration, but rather the existence of a significant smin.  相似文献   

13.
Channelrhodopsins (ChRs) are light-gated cation channels. After blue-light excitation, the protein undergoes a photocycle with different intermediates. Here, we have recorded transient absorbance changes of ChR2 from Chlamydomonas reinhardtii in the visible and infrared regions with nanosecond time resolution, the latter being accomplished using tunable quantum cascade lasers. Because proton transfer reactions play a key role in channel gating, we determined vibrational as well as kinetic isotope effects (VIEs and KIEs) of carboxylic groups of various key aspartic and glutamic acid residues by monitoring their C=O stretching vibrations in H2O and in D2O. D156 exhibits a substantial KIE (>2) in its deprotonation and reprotonation, which substantiates its role as the internal proton donor to the retinal Schiff base. The unusual VIE of D156, upshifted from 1736 cm−1 to 1738 cm−1 in D2O, was scrutinized by studying the D156E variant. The C=O stretch of E156 shifted down by 8 cm−1 in D2O, providing evidence for the accessibility of the carboxylic group. The C=O stretching band of E90 exhibits a VIE of 9 cm−1 and a KIE of ∼2 for the de- and the reprotonation reactions during the lifetime of the late desensitized state. The KIE of 1 determined in the time range from 20 ns to 5 ms is incompatible with early deprotonation of E90.  相似文献   

14.
The light-harvesting 1 reaction center (LH1-RC) complex from Thermochromatium (Tch.) tepidum exhibits unusual Qy absorption by LH1 bacteriochlorophyll-a (BChl-a) molecules at 915 nm, and the transition energy is finely modulated by the binding of metal cations to the LH1 polypeptides. Here, we demonstrate the metal-dependent interactions between BChl-a and the polypeptides within the intact LH1-RC complexes by near-infrared Raman spectroscopy. The wild-type LH1-RC (B915) exhibited Raman bands for the C3-acetyl and C13-keto CO stretching modes at 1637 and 1675 cm? 1, respectively. The corresponding bands appeared at 1643 and 1673 cm? 1 when Ca2 + was biosynthetically replaced with Sr2 + (B888) or at 1647 and 1669 cm? 1 in the mesophilic counterpart, Allochromatium vinosum. These results indicate the significant difference in the BChl–polypeptide interactions between B915 and B888 and between B915 and the mesophilic counterpart. The removal of the original metal cations from B915 and B888 resulted in marked band shifts of the C3-acetyl/C13-carbonyl νCO modes to ~ 1645/~ 1670 cm? 1, supporting a model in which the metal cations are involved in the fine-tuning of the hydrogen bonding between the BChl-a and LH1-polypeptides. Interestingly, the interaction modes were almost identical between the Ca2 +-depleted B915 and Sr2 +-depleted B888 and between B915 and Ca2 +-substituted B888, despite the significant differences in their LH1 Qy peak positions and the denaturing temperatures, as revealed by differential scanning calorimetry. These results suggest that not only the BChl–polypeptide interactions but some structural origin may be involved in the unusual Qy red-shift and the enhanced thermal stability of the LH1-RC complexes from Tch. tepidum.  相似文献   

15.
Photosynthetic carbon metabolism of a marine grass   总被引:5,自引:4,他引:5       下载免费PDF全文
The δ13C value of a tropical marine grass Thalassia testudinum is −9.04‰. This value is similar to the δ13C value of terrestrial tropical grasses. The δ13C values of the organic acid fraction, the amino acid fraction, the sugar fraction, malic acid, and glucose are: −11.2‰, −13.1‰, −10.1‰, −11.1‰, and −11.5‰, respectively. The δ13C values of malic acid and glucose of Thalassia are similar to the δ13C values of these intermediates in sorghum leaves and attest to the presence of the photosynthetic C4-dicarboxylic acid pathway in this marine grass. The inorganic HCO3 for the growth of the grass fluctuates between −6.7 to −2.7‰ during the day. If CO2 fixation in Thalassia is catalyzed by phosphoenolpyruvate carboxylase (which would result in a −3‰ fractionation between HCO3 and malic acid), the predicted δ13C value for Thalassia would be −9.7 to −5.7‰. This range is close to the observed range of −12.6 to −7.8‰ for Thalassia and agree with the operation of the C4-dicarboxylic acid pathway in this plant. The early products of the fixation of HCO3 in the leaf sections are malic acid and aspartic acid which are similar to the early products of CO2 fixation in C4 terrestrial plants.  相似文献   

16.
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   

17.
Cytoglobin (Cygb) was investigated for its capacity to function as a NO dioxygenase (NOD) in vitro and in hepatocytes. Ascorbate and cytochrome b5 were found to support a high NOD activity. Cygb-NOD activity shows respective Km values for ascorbate, cytochrome b5, NO, and O2 of 0.25 mm, 0.3 μm, 40 nm, and ∼20 μm and achieves a kcat of 0.5 s−1. Ascorbate and cytochrome b5 reduce the oxidized Cygb-NOD intermediate with apparent second order rate constants of 1000 m−1 s−1 and 3 × 106 m−1 s−1, respectively. In rat hepatocytes engineered to express human Cygb, Cygb-NOD activity shows a similar kcat of 1.2 s−1, a Km(NO) of 40 nm, and a kcat/Km(NO) (kNOD) value of 3 × 107 m−1 s−1, demonstrating the efficiency of catalysis. NO inhibits the activity at [NO]/[O2] ratios >1:500 and limits catalytic turnover. The activity is competitively inhibited by CO, is slowly inactivated by cyanide, and is distinct from the microsomal NOD activity. Cygb-NOD provides protection to the NO-sensitive aconitase. The results define the NOD function of Cygb and demonstrate roles for ascorbate and cytochrome b5 as reductants.  相似文献   

18.
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   

19.
We report that two species of mouse-tailed bats (Rhinopoma microphyllum and R. cystops) hibernate for five months during winter in geothermally heated caves with stable high temperature (20°C). While hibernating, these bats do not feed or drink, even on warm nights when other bat species are active. We used thermo-sensitive transmitters to measure the bats’ skin temperature in the natural hibernacula and open flow respirometry to measure torpid metabolic rate at different ambient temperatures (Ta, 16–35°C) and evaporative water loss (EWL) in the laboratory. Bats average skin temperature at the natural hibernacula was 21.7 ± 0.8°C, and no arousals were recorded. Both species reached the lowest metabolic rates around natural hibernacula temperatures (20°C, average of 0.14 ± 0.01 and 0.16 ± 0.04 ml O2 g−1 h−1 for R. microphyllum and R. cystops, respectively) and aroused from torpor when Ta fell below 16°C. During torpor the bats performed long apnoeas (14 ± 1.6 and 16 ± 1.5 min, respectively) and had a very low EWL. We hypothesize that the particular diet of these bats is an adaptation to hibernation at high temperatures and that caves featuring high temperature and humidity during winter enable these species to survive this season on the northern edge of their world distribution.  相似文献   

20.
Extracellular vesicles (EVs) produced by a sulfur-reducing, hyperthermophilic archaeon, “Thermococcus onnurineus” NA1T, were purified and characterized. A maximum of four EV bands, showing buoyant densities between 1.1899 and 1.2828 g cm−3, were observed after CsCl ultracentrifugation. The two major EV bands, B (buoyant density at 25°C [ρ25] = 1.2434 g cm−3) and C (ρ25 = 1.2648 g cm−3), were separately purified and counted using a qNano particle analyzer. These EVs, showing different buoyant densities, were identically spherical in shape, and their sizes varied from 80 to 210 nm in diameter, with 120- and 190-nm sizes predominant. The average size of DNA packaged into EVs was about 14 kb. The DNA of the EVs in band C was sequenced and assembled. Mapping of the T. onnurineus NA1T EV (ToEV) DNA sequences onto the reference genome of the parent archaeon revealed that most genes of T. onnurineus NA1T were packaged into EVs, except for an ∼9.4-kb region from TON_0536 to TON_0544. The absence of this specific region of the genome in the EVs was confirmed from band B of the same culture and from bands B and C purified from a different batch culture. The presence of the 3′-terminal sequence and the absence of the 5′-terminal sequence of TON_0536 were repeatedly confirmed. On the basis of these results, we hypothesize that the unpackaged part of the T. onnurineus NA1T genome might be related to the process that delivers DNA into ToEVs and/or the mechanism generating the ToEVs themselves.  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