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 577 毫秒
1.
Two simple experiments measuring the 13C linewidths ν1/2 and spin–lattice relaxation times T1 of each of the signals in the spectrum of trilinolein indicate that the ν1/2 and T1 values are consistent with the different degrees of motional freedom expected for the various 13C nuclei. However, for each chain, the ν1/2 and T1 measurements indicate a small reversal in mobility at C-10 relative to C-9 before motional freedom again steadily increases on each chain starting at C-11. The T1 experiment allows unambiguous assignments of the C-8 signal and C-14 signal, which differ by only 0.010 ppm. Measurements of 13C ν1/2 and T1 values on tripalmitin provide secure assignments for the C-5 and C-6 signals, for which conflicting assignments have been reported. The T1 measurements also show that among the tightly clustered C-8 through C-12 signals, the C-11 signals are the most downfield, while the C-12 signals are the most upfield, again contrary to a previous report. Similar measurements of 13C ν1/2 and T1 values on other triacylglycerols or related compounds may prove equally useful in making chemical shift assignments and detecting any discontinuities in motional freedom along a chain. The benefits and possible limitations of ultrahigh field NMR for studying triacylglycerols and related compounds are discussed.  相似文献   

2.
Evidence is presented suggesting, for the first time, that the protein foldability metric σ = (Tθ − Tf) / Tθ, where Tθ and Tf are, respectively, the collapse and folding transition temperatures, could be used also to measure the foldability of RNA sequences. These results provide further evidence of similarities between the folding energy landscapes of proteins and RNA. The importance of σ is discussed in the context of the in silico design of rapidly foldable RNA sequences.  相似文献   

3.
Deuterium magnetic resonance (2H-NMR) and Raman spectroscopy are used to investigate order and fluidity at the terminal methyl position in 16-d3, 16′-d3 dipalmitoylphosphatidylcholine (16-d6 DPPC) multibilayers. These methods reveal substantial motion and disorder in the gel phase, 5–10°C below the gel-liquid crystal phase transition temperature (Tm). The phase transition is sensed in the 2H-NMR spectrum as a reduction in the quadrupole splitting from 14 kHz to 3 kHz. In contrast, the Raman parameter used to characterize the CD3 vibrations is quite insensitive to the melting process, although an analogous parameter does sense disordering at Tm at the 10 and 10′ position in 10-d2, 10′-d2 DPPC. The difference in the response of the NMR and Raman parameters may arise because the vibrational spectrum of the CD3 group is inhomogenously broadened and is therefore quite sensitive to alterations in the local environment around the methyl group. In contrast, the NMR quadrupole splitting is sensitive to both local motion of the methyl group and, near Tm, to motions of the CD2 group induced by transgauche isomerizations further up the chain. The difficulties that arise when results from different spectroscopic techniques are compared are demonstrated.  相似文献   

4.
An enzymatic method has been used to isolate, for the first time, polymeric suberin from the bark of Quercus suber L. or cork. This was achieved by solvent extraction (dichloromethane, ethanol and water), followed by a step-by-step enzymatic treatment with cellulase, hemicellulase and pectinase, and a final extraction with dioxane/water. The progress of suberin isolation was monitored by Fourier transform infrared spectroscopy using a photoacoustic cell (FTIR-PAS). The material obtained (polymeric suberin (PS)) was characterised by solid-state and liquid-state nuclear magnetic resonance, FTIR-PAS and vapour pressure osmometry, and compared with the suberin fraction obtained by alkaline depolymerisation (depolymerised suberin (DS)). The results showed that PS is an aliphatic polyester of saturated and unsaturated fatty acids, with an average molecular weight (M(w)) of 2050 g mol(-1). Although this fraction represents only 10% of the whole suberin of cork, its polymeric nature gives valuable information about the native form of the polymer. DS was found to have an average M(w) of 750 g mol(-1) and to comprise a significant amount of acidic and alcoholic short aliphatic chains.  相似文献   

5.
Daily circadian rhythms of body temperature (Tb) and oxygen consumption (VO2) were measured in two murid species, which occupy extremely different habitats in Israel. The golden spiny mouse (Acomys mssalus) is a diurnal murid distributed in arid and hot parts of the great Syrio-African Rift Valley, while the broad-toothed field mouse (Apodeinns mystacinus) is a nocturnal species that inhabits the Mediterranean woodlands. In both species, the daily rhythms of Tb and VO2 are entrained by the photoperiod. Under laboratory experimental conditions (ambient temperature Ta = 33oC and photoperiod regime of 12L: 12D), Acomys russatus exhibits a tendency towards a nocturnal activity pattern, compared to the diurnal activity displayed by this species under natural conditions. Under the same photoperiod regime and at Ta = 28oC, Apodemus mystacinus displays nocturnal activity, as observed under natural conditions. The maximal values of Tb were recorded in Acomys russatus at midnight (23:50 h), while the maximal values of VO2 were recorded at the beginning of the dark period (18:20 h). In Apodemus mystacinus, the maximal values of Tb and VO2 were recorded at 23:40 and 20:00 h, respectively. The ecophysiological significance of these results is discussed further.  相似文献   

6.
The long-term dynamics of an amoeboid cell shape were studied using Physarum polycephalum plasmodia with various sizes. Cell shape varied oscillatorily in a multiple periodic manner. The organism periodically elongated with period of T7 = 10 h, branched with T6 = 4 h, became uneven with T5 = 30 min and T4 = 10 min, and blew up with T3 = 1.5 min. Tiny plasmodia changed shape much faster with T3 = 1.3 min, T2 = 24 s and T1 = 3.3 s simultaneously. The plasmodial cytoskeleton also showed periodic pattern formation with T6, T5 and T3. Periods of all known oscillatory phenomena in this organism correspond to some of the periods for the above seven rhythms, and the following geometric progression holds among the periods: Ti + 1/Ti = 7 and Ti + 2/Ti + 1 = 3, where i = 1, 3, 5. Thus, multiple oscillations in the plasmodium are organized globally.  相似文献   

7.

1. 1.|In 15 conscious Pekin ducks, 40 “warm sensitive” hypothalamic neurons were identified according to their discharge rates at 40°C Thy (F40), local temperature coefficients (Δ/ΔT) and Q10.

2. 2.|Q10 and either F40 or ΔFT were little or not related.

3. 3.|A positive correlation between F40 and ΔFT was observed which was particularly close (r = 0.94 and 0.96) when the neurons were classified according to their Q10 of <2 and >2.

4. 4.|The results suggest that neurons with positive temperature coefficients in the duck's hypothalamus mostly exhibit linear to exponential temperature-discharge relationships.

5. 5.|This is an contrast to observations on mammalian hypothalamic thermosensitive neurons and may relate to the absence of the thermosensory function in the duck's rostral brainstem.

Author Keywords: Neuronal thermosensitivity; hypothalamic thermosensory function; Temperature and synaptic transmission; avian thermoregulation; mammalian thermoregulation  相似文献   


8.
Keith A. Rose  Alan Bearden 《BBA》1980,593(2):342-352
Electron paramagnetic resonance (EPR) power saturation and saturation recovery methods have been used to determine the spin lattice, T1, and spin-spin, T2, relaxation times of P-700+ reaction-center chlorophyll in Photosystem I of plant chloroplasts for 10 K T 100 K. T1 was 200 μs at 100 K and increased to 900 μs at 10 K. T2 was 40 ns at 40 K and increased to 100 ns at 10 K. T1 for 40 K T 100 K is inversely proportional to temperature, which is evidence of a direct-lattice relaxation process. At T = 20 K, T1 deviates from the 1/T dependence, indicating a cross relaxation process with an unidentified paramagnetic species. The individual effects of ascorbate and ferricyanide on T1 of P-700+ were examined: T1 of P-700+ was not affected by adding 10 mM ascorbate to digitonin-treated chloroplast fragments (D144 fragments). The P-700+ relaxation time in broken chloroplasts treated with 10 mM ferricyanide was 4-times shorter than in the untreated control at 40 K. Ferricyanide appears to be relaxing the P-700+ indirectly to the lattice by a cross-relaxation process. The possibility of dipolar-spin broadening of P-700+ due to either the iron-sulfur center A or plastocyanin was examined by determining the spin-packet linewidth for P-700+ when center A and plastocyanin were in either the reduced or oxidized states. Neither reduced center A nor oxidized plastocyanin was capable of broadening the spin-packet linewidth of the P-700+ signal. The absence of diplolar broadening indicates that both center A and plastocyanin are located at a distance at least 3.0 nm from the P-700+ reaction center chlorophyll. This evidence supports previous hypotheses that the electron donor and acceptor to P-700 are situated on opposite sides of the chloroplast membrane. It is also shown that the ratio of photo-oxidized P-700 to photoreduced centers A and B at low temperature is 2 : 1 if P-700 is monitored at a nonsaturating microwave power.  相似文献   

9.
The effects of β-glucan (BG) prepared from spent brewer’s yeast on gelatinization and retrogradation of rice starch (RS) were investigated as functions of mixing ratio and of storage time. Results of rapid visco-analysis (RVA) indicated that addition of BG increased the peak, breakdown, setback, and final viscosities, but decreased the pasting temperatures of the rice starch/β-glucan (RS/BG) mixtures. Differential scanning calorimetry (DSC) data demonstrated an increase in onset (To), peak (Tp), and conclusion (Tc) temperatures and a decrease in gelatinization enthalpy (ΔH1) with increasing BG concentration. Storage of the mixed gels at 4 °C resulted in a decrease in To, Tp, Tc, and melting enthalpy (ΔH2). The retrogradation ratio (ΔH2H1) and the phase transition temperature range (Tc − To) of the mixed gels increased with storage time, but this effect was reduced by the addition of BG. BG addition also slowed the syneresis of the mixed gels. Results of dynamic viscoelasticity measurement indicated that the addition of BG promoted RS retrogradation at the beginning and then retarded it during longer storage times. The added BG also retarded the development of gel hardness during refrigerated storage of the RS/BG mixed gels.  相似文献   

10.
Yield stress of 6% (w/w) waxy maize (WXM), cross-linked waxy maize (CLWM), and cold water swelling (CWS) starches in xanthan gum dispersions: 0%, 0.35%, 0.50%, 0.70%, and 1.0% was measured with the vane method at an apparent shear rate of 0.05 s−1. The intrinsic viscosity of the xanthan gum was determined to be: 112.3 dL/g in distilled water at 25 °C. Values of the static (σ0s) and dynamic (σ0d) yield stress of each dispersion were measured before and after breaking down its structure under continuous shear, respectively. The WXM and CWS starches exhibited synergistic behavior, whereas the CLWM starch showed antagonistic effect with xanthan gum. The difference (σ0s − σ0d) was the stress required to break the inter-particle bonding (σb). The contributions of the viscous (σv) and network (σn) components were estimated from an energy balance model. In general, values of σb of the starch–xanthan gum dispersions decreased and those of σn increased with increase in xanthan gum concentration.  相似文献   

11.
A general 3-D dynamic model for men's and women's discus flight is presented including precession of spin angular momentum induced by aerodynamic pitching moment. Dependence of pitching moment coefficient on angle of attack is estimated from experiment. Numerical integration of 11 equations of motion for nominal release speed v0=25 m/s and axial spin p0=42 rad/s also requires 3 other release conditions; initial discus flight path angle β0, pitch attitude θ0, and roll angle φ0. Optimal values for these release conditions are calculated iteratively to maximize range and are similar for both men and women. The optimal men's trajectory and range R=69.39 m is produced by the strategy β0=38.4°, θ0=30.7°, and φ0=54.4°. Initial angular velocities except spin are chosen to minimize wobble but an optimal initial spin rate p0=25.2 rad/s exists that also maximizes range. Optimal 3-D range exceeds that predicted by 2-D models because, although angle of attack and lift are negative initially, 3-D motion allows advantageous orientation of lift later in flight, with tilt of the axis of symmetry from vertical becoming much smaller at landing. Optimal strategies are discontinuous with wind speed, resulting in slicing and kiting strategies in large head and tail winds, respectively. Sensitivity of optimal range is largest to initial β0 and least to φ0. Present calculations do not account for dependence of initial release angle or spin on release velocity or among other release conditions.  相似文献   

12.
R M Santos  E Rojas 《FEBS letters》1987,220(2):342-346
The effects of forskolin on electrical coupling among pancreatic β-cells were studied. Two microelectrodes were used to measure membrane potentials simultaneously in pairs of islet β-cells. Intracellular injection of a current pulse (ΔI) elicited a membrane response ΔV1 in the injected cell and also a response ΔV2 in a nearby β-cell confirming the existence of cell-to-cell electrical coupling among islet β-cells. In the presence of glucose (7 mM), application of forskolin evoked a transient depolarization of the membrane and electrical activity suggesting that the drug induced a partial inhibition of the β-cell membrane K+ conductance. Concomitant with this depolarization of the membrane there was a marked decrease in β-cell input resistance (ΔV2/ΔI) suggesting that exposure to forskolin enhanced intercellular coupling. Direct measurements of the coupling ratio ΔV2/ΔV1 provided further support to the idea that forskolin enhances electrical coupling among islet cells. Indeed, application of forskolin reversibly increased the coupling ratio. These results suggest that cAMP might be involved in the modulation of electrical coupling among islet β-cells.  相似文献   

13.
This paper describes a 13C solid state NMR study of hydrated powders and gels of locust bean gum galactomannan-LBG and Konjac glucomannan-KGM. Changes in relative spectral intensities, cross-polarization dynamics (TCH, T1ρH) and relaxation times (T1C, T1H, T2H) show that hydration (0–90%) of LBG powders increases the 108 Hz frequency molecular motions, probably reflecting the enhanced motion of non-aggregating segments and chain ends. Slower motions (104–105 Hz) are enhanced only slightly at 90% hydration. LBG gel shows higher spatial distinction between aggregated and non-aggregated segments than the hydrated powder and relaxation times indicate higher mobility for galactose-ramified segments, compared to linear mannose segments. While the dynamics of KGM hydration is similar to that of LBG, i.e. mainly affecting fast 108 Hz motions, the gel is significantly more rigid. Both spectra and relaxation times show that glucose residues in KGM gel are particularly hindered, probably due to their preferential involvement in chain aggregation.  相似文献   

14.
Gay Goodman  John S. Leigh  Jr. 《BBA》1987,890(3):360-367
The electron-spin relaxation rates of the two species of cytochrome a3+3-azide found in the azide compound of bovine-heart cytochrome oxidase were measured by progressive microwave saturation at T = 10 K. It has been shown previously that Cyt a+33-azide gives rise to two distinct EPR resonances, depending upon the oxidation state of Cyt a. When Cyt a is ferrous, Cyt a3+3-azide has g = 2.88, 2.19 and 1.64; upon oxidation of Cyt a, the a3+3-azide g-values become g = 2.77, 2.18, and 1.74 (Goodman, G. (1984) J. Biol. Chem. 259, 15094–15099). The relaxation effect of Cyt a on Cyt a3 could be measured as the difference in microwave field saturation parameter H1/2 between the g = 2.77 and g = 2.88 species. For each signal the spin-lattice relaxation time T1 was determined from H1/2 using the transverse relaxation time T2. The value of T2 at 10 K was extrapolated from a plot of line-width vs. temperature at higher temperature. The dipolar contribution to T1 was related to the Cyt a-Cyt a3 spin-spin distance utilizing available information on the relative orientation of Cyt a3-azide and Cyt a (Erecinska, M., Wilson, D.F. and Blasie, J.K. (1979) Biochim. Biophys. Acta 545, 352–364). By taking into account the relaxation parameters for both gx and gz components of the Cyt a3-azide g-tensor, the angle between the gz components of the Cyt a and Cyt a3g-tensors was determined to be between 0 and 18°, and the Cyt a-Cyt a3 spin-spin distance was found to be 19 ± 8 Å.  相似文献   

15.
Conductance and relaxations of gelatin films in glassy and rubbery states   总被引:1,自引:0,他引:1  
The dielectric constant, ′, and the dielectric loss, ″, for gelatin films were measured in the glassy and rubbery states over a frequency range from 20 Hz to 10 MHz; ′ and ″ were transformed into M* formalism (M*=1/(′−i″)=M′+iM″; i, the imaginary unit). The peak of ″ was masked probably due to dc conduction, but the peak of M″, e.g. the conductivity relaxation, for the gelatin used was observed. By fitting the M″ data to the Havriliak–Negami type equation, the relaxation time, τHN, was evaluated. The value of the activation energy, Eτ, evaluated from an Arrhenius plot of 1/τHN, agreed well with that of Eσ evaluated from the DC conductivity σ0 both in the glassy and rubbery states, indicating that the conductivity relaxation observed for the gelatin films was ascribed to ionic conduction. The value of the activation energy in the glassy state was larger than that in the rubbery state.  相似文献   

16.

1. 1.|Body temperatures (Tb) and contaneous evaporative water loss rates (CWL) were measured in tree frogs (Hyla cinerea) and toads (Bufo valliceps) exposed to cyclical ramp changes in water vapor density (WVD) between 7.5 and 9.8 gm−3 (1 cycle h−1 at an air temperature of 27.0°C.

2. 2.|CWL was 3.3 times greater in toads than in tree frogs.

3. 3.|Tb in toads cycled directly with WVd; WVD accounted for 98% of the variation in toad Tb.

4. 4.|Tb in tree frogs was independent of WVD, probably due to changes in skin resistance to water loss.

Author Keywords: Body temperature; evaporative water loss; skin resistance; water vapor density; relative humidity; Anura; Hyla cinerea; Bufo valliceps  相似文献   


17.
The effect of hydration on the molecular dynamics of soft wheat gluten was investigated by solid state NMR. For this purpose, we recorded static and MAS 1H spectra and SPE, CP, and other selective 13C spectra under MAS and dipolar decoupling conditions on samples of dry and H2O and D2O hydrated gluten. Measurements of carbon-proton CP times and several relaxation times (proton T1, T and T2, and carbon T1) were also performed. The combination of these techniques allowed both site-specific and domain-averaged motional information to be obtained in different characteristic frequency ranges. Domains with different structural and dynamic behaviour were identified and the changes induced by hydration on the dynamics of different domains could be monitored. The proton spin diffusion process was exploited to get information on the degree of mixing among different gluten domains. The results are consistent with the “loop and train” model proposed for hydrated gluten.  相似文献   

18.

1. 1. Seven thermal conditions were imposed on male sitting subjects (slightly clothed: 0.6 clo).

2. 2. A thermal mannikin was also used to determine the exact operative temperature, T0.

3. 3. Conditions were: uniform (UN: all parameters at 24.5°C, air velocity at 0.15 ms−1), heated ceiling (HC at 45°C), heated floor (HF at 34°C), cold floor (CF at 14°C), two conditions of one cold wall at 6°C (CW1 and CW2 respectively with and without air temperature compensation) and increased air velocity (AV at 0.4 ms−1).

4. 4. Local skin temperatures and answers to questionnaires were obtained.

5. 5. Skin temperature variations were affected by conditions and slight T0 changes.

6. 6. Comfort judgments were fairly well related to T0, especially when expressed as differences between actual non-uniform environment and the uniform one.

7. 7. It is concluded that, in case of non-uniform environments close to thermoneutral zone, thermal comfort or discomfort reflects the climate alterations better than the thermal sensation does.

Author Keywords: Skin temperature; thermal sensation; comfort; climate heterogeneity  相似文献   


19.
Cell wall polysaccharide interactions in maize bran   总被引:7,自引:0,他引:7  
Sequential extractions with alkali have been carried out in order to study the nature of linkages which hold heteroxylans in maize bran cell walls. Treatment with 0.5 sodium hydroxide at 30 °C for 2 h released all the phenolic acids (p-coumaric, ferulic, and diferulic) but extracted only ˜30% of heteroxylans (S1); further treatment with 1.5 potassium hydroxide at 100 °C for 2 h released the remaining heteroxylans (S2). The heteroxylans from S1 and S2 had a similar neutral sugar composition and structure, but their weight average molecular weights were 270 kDa (Mw/Mn = 2) and 370 kDa (Mw/Mn = 2.8), respectively. Proteins (5%) are found with polysaccharides from S1 and S2, with different amino acid composition. The results suggest that covalent linkages through phenolic acids are only partly responsible for the associations of maize bran heteroxylans in the cell wall and that linkages to structural cell wall proteins were probably the main cause of heteroxylan insolubility.  相似文献   

20.
《植物生态学报》2017,41(4):396
Aims Stem CO2 efflux (Es) is an important component of annual carbon budget in forest ecosystems, but how biotic and environmental factors regulate seasonal and inter-specific variations in Es is poorly understood. The objectives of this study were: (1) to compare seasonal dynamics in Es for four temperate coniferous tree species in northeastern China, including Korean pine (Pinus koraiensis), Korean spruce (Picea koraiensis), Mongolian pine (Pinus sylvestris var. mongolica), and Dahurian larch (Larix gmelinii); and (2) to explore factors driving the inter-specific variability in Es during the growing and non-growing seasons.
Methods Ten to twelve trees for each tree species were sampled for Es and stem temperature at 1 cm depth beneath the bark (Ts) measurements in situ with an infrared gas analyzer (LI-6400 IRGA) and a digital thermometer, respectively, from July to October 2013 and March to July 2014. The daily stem circumference increment (Si), sapwood nitrogen concentration ([N]), and related environmental factors were monitored simultaneously.
Important findings The temporal variation in Es for the four tree species overall followed the changes in Ts throughout the study period, with the maxima occurring in the summer months (late May to early July) characterized by higher temperature and more rapid stem growth and the minima in spring (late March to April) or autumn (October) having lower temperature. Ts accounted for 42%-91% and 56%-89% of variations in Es during the growing (May to September) and non-growing (other months) seasons, respectively. Furthermore, apart from Ts, we also found significant regression relationships between Es and Si, relative air humidity and [N] during the growing season, but their forms and correlation coefficients were species-dependent. These results indicated that Ts was the dominant environmental factor affecting seasonal variations in Es, but the magnitude of the effect varied with tree species and growth rhythm. Mean Es for each of the four tree species was significantly higher in the growing season than in the non-growing season, whereas within the season there were also significant differences in mean Es among the tree species (all p < 0.05). The temperature sensitivity of Es (Q10 value) did not differ significantly among the tree species during the growing season, ranging from 1.64 for Dahurian larch to 2.09 for Mongolian pine, but did differ during the non-growing season which varied from 1.80 for Korean pine to 3.14 for Dahurian larch. Moreover, Korean spruce, Mongolian pine and Dahurian larch had significantly greater Q10 values in the non-growing season than in the growing season (p < 0.05). These findings suggested that the differences of the response of Es to temperature change for different tree species were mainly from the non-growing season. Because the seasonality and inter-specific variability in Es for these temperate coniferous tree species were primarily controlled by multiple factors such as temperature, we conclude that using a single annual temperature response curve to estimate the annual Es may lead to more uncertainty.  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