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.
Highly oriented fibers of Li-, Na-, K-, and CsDNA were prepared with a previously developed wet spinning method. The procedure gave a large number of equivalent fiber bundle samples (reference length, L0, typically = 12–15 cm) for systematic measurements of the fiber length L in ethanol–water solutions, using a simple mechanochemical set up. The decrease in relative length L/L0 with increasing ethanol concentration at room temperature gave evidence for the B-A transition centered at 76% (v/v) ethanol for NaDNA fibers and at 80 and 84% ethanol for K- and CsDNA fibers. A smaller decrease in L/L0 of LiDNA fibers was attributed to the B-C transition centered at 80% ethanol. In a second type of experiment with DNA fibers in ethanol–water solutions, the heat-induced helix–coil transition, or melting, revealed itself in a marked contraction of the DNA fibers. The melting temperature Tm, decreased linearly with increasing ethanol concentration for fibers in the B-DNA ethanol concentration region. In the B-A transition region, Na- and KDNA fibers showed a local maximum in Tm. On further increase of the ethanol concentration, the A-DXA region followed with an even steeper linear decrease in Tm. The dependence on the identity of the counterion is discussed with reference to the model for groove binding of cations in B-DNA developed by Skuratovskii and co-workers and to the results from Raman studies of the interhelical bonds in A-DNA performed by Lindsay and co-workers. An attempt to apply the theory of Chogovadze and Frank-Kamenetskii on DNA melting in the B-A transition region to the curves failed. However, for Na- and KDNA the Tm dependence in and around the A-B transition region could be expressed as a weighted mean value of Tm of A- and B-DNA. On further increase of the ethanol concentration, above 84% ethanol for LiDNA and above about 90% ethanol for Na-, K-, and CsDNA, a drastic change occurred. Tm increased and a few percentages higher ethanol concentrations were found to stabilize the DNA fibers so that they did not melt at all, not even at the upper temperature limit of the experiments (~ 80°C). This is interpreted as being due to the strong aggregation induced by these high ethanol concentrations and to the formation of P-DNA. Many features of the results are compatible with the counterion–water affinity model. In another series of measurements, Tm of DNA fibers in 75% ethanol was measured at various salt concentrations. No salt effect was observed (with the exception of LiDNA at low salt concentrations). This result is supported by calculations within the Poisson–Boltzmann cylindrical cell model. © 1994 John Wiley & Sons, Inc.  相似文献   

2.
The refractive indices, water content, and volume of wet-spun films of Na-hyaluronate have been measured as a function of relative humidity (rh). These data are used with the Lorentz-Lorenz formula to determine the optical polarizabilities of Na-hyaluronate parallel and perpendicular to the helical axis. The analysis reveals a drop in the optical polarizabilities of approximately 20% between 80 and 88% rh, indicating a phase transition.  相似文献   

3.
《Free radical research》2013,47(5):413-421
Abstract

Esterification by β-apo-14’-carotenoic acid was found to have opposite effects on antioxidant activity of quercetin (at B4’, B3’ hydroxyl) as of daidzein (at A7 hydroxyl) in phosphatidylcholine liposomes. The daidzein ester had increased activity, while quercetin had a significant decreased activity. Quantum mechanical calculations using density function theory (DFT) indicate a modest decrease in bond dissociation enthalpy, BDE, for (weakest) hydrogen–oxygen phenolic bond in daidzein from 368.4 kJ·mol? 1 to 367.7 kJ·mol? 1 compared to a significant increase in quercetin from 329.5 kJ·mol? 1 to 356.6 kJ·mol? 1 upon derivatization. These opposite changes in tendency for hydrogen atom transfer from phenolic groups to lipid radicals combined with an increase in A-to-B dihedral angle from 0.0° to 36.4° and in dipole moment from 0.40 D to 6.01 D for quercetin upon derivatization, while less significant for daidzein (36.4°–36.7° and 3.26 D–7.87 D, respectively), together provide a rationale for the opposite effect of esterification on antioxidation.  相似文献   

4.
Abstract

The double helical structure of the self-complementary DNA-RNA-DNA hybrid d(CG)r(CG) d(CG) was studied in solution by 500 MHz 1H-NMR spectroscopy. The non-exchangeable base protons and the (deoxy)ribose H1′, H2′ and H2″ protons were unambiguously assigned using 2D-J-correlated (COSY) and 2D-NOE (NOESY) spectroscopy techniques. A general strategy for the sequential assignment of 1H-NMR spectra of (double) helical DNA and RNA fragments by means of 2D-NMR methods is presented.

Conformational analysis of the sugar rings of d(CG)r(CG)d(CG) at 300 K shows that the central ribonucleotide part of the helix adopts an A-type double helical conformation. The 5′- and 3′-terminal deoxyribose base pairs, however, take up the normal DNA-type conformation. The A-to-B transition in this molecule involves only one (deoxyribose) base pair. It is shown that this A-to-B conformational transition can only be accomodated by two specific sugar pucker combinations for the junction base pair, i.e. N·S (C3′-endo-C2′-endo, 60%, where the pucker given first is that assigned to the junction nucleotide residue of the strand running 5′ → 3′ from A-RNA to B-DNA) and S·S (C2′-endo-C2′-endo, 40%).  相似文献   

5.
Pei‐Kun Yang 《Biopolymers》2014,101(8):861-870
To explore the effect of an external electrostatic field (EEF) on the stability of protein conformations, the molecular dynamic modeling approach was applied to evaluate the effect of an EEF along the x or y direction on a water cluster containing a parallel or antiparallel β sheet structure. The β sheet structure contained two strands with a (Gly)3 sequence separated by a distance d along the x direction. The mean forces between the two strands along the x direction were computed from the trajectories of molecular dynamics simulations. In the absence of the EEF, the forces between the two strands in vacuum were repulsive and attractive in the parallel and antiparallel β sheet structures, respectively. In contrast, the mean forces between the two strands in water were attractive in both the parallel and antiparallel β sheet structures. This is because the electric interactions between the two strands were shielded by water, and the hydrophobic effect dominated the interaction between the two strands. When an EEF >50 MV/cm was applied to the water cluster, the attractive force between the two strands in the parallel and antiparallel β sheet structures decreased and increased, respectively. Further, the binding affinity between the two strands in the parallel and antiparallel β sheet structures also decreased and increased, respectively. This is because the large EEF leads to dielectric saturation, and consequently reduces the effects of the dielectric shielding and hydrophobic interactions. © 2014 Wiley Periodicals, Inc. Biopolymers 101: 861–870, 2014.  相似文献   

6.
Gelatin, soy protein, lysozyme, succinyl-casein and succinyl-egg albumin were allowed to react with methyl linoleate (ML) at a relative humidity (RH) of 0% at 50°C for 7 days (protein: ML = 1:1). Gel filtration indicated that only gelatin was extensively fragmented. The gelatin was then incubated with ML under various conditions, and changes in the molecular sizes, the gel forming abilities and the chemical characteristics were investigated. The fragmentation of gelatin was increased by decreasing the RH and with the increase in the ratio of ML to protein. The melting point of gel in heating and cooling gelatin was decreased by increasing the fragmentation. The contents of amide and carbonyl groups increased and that of amino group decreased as the reaction progressed at RH 0%, but no change in C-terminal amino acids was observed. Following the reaction at RH 0%, many kinds of amino acid residues of gelatin were damaged, although in our previous paper [Matoba et al.,Agric. Biol. Chem. , 46, 979 (1982)] such was not detected in casein and egg albumin. From the above results, we conclude that gelatin is susceptible to fragmentation by reaction with oxidizing lipids and one possible mechanism of the degradation may be the –N–C– scission of peptide bonds as proposed by Zirlin and Karel [J. Food Sci., 34, 160 (1969)]. Complex reactions other than this scission may also occur.  相似文献   

7.
We have firstly demonstrated the renaturation process of dissociated single chains of lentinan (s‐LNT) and the variable conformations of the renatured LNT (r‐LNT). The results from ultrasensitive differential scanning calorimetry and circular dichroism revealed that the variable structures including perfect triple helix, defective triple helix containing duplex segment, and single chains occurred in the renaturation of s‐LNT, depending on the renaturation time, solvent composition, molecular weight, and the mode of renaturation. When water was added into s‐LNT/dimethylsulfoxide (DMSO) to reach 95% (v/v), the classic low‐temperature intra‐triple‐helical conformational transition at ~10°C (T1) appeared within 4 h, indicative of a rapid reconstruction of triple helical structure. Besides, one newly endothermic peak at ~43°C (T2) simultaneously occurred, which was first ascribed to the melting of duplex segment in the imperfect triplex. The duplex stretches disappeared when DMSO reached 50%, in which single chains coexisted with triplex. Moreover, the duplex segment disappeared by slowly dropping water into s‐LNT/DMSO. This work suggested that the structure of r‐LNT could be controllable, and provided important information for their successful development and application in polymer and life science. © 2012 Wiley Periodicals, Inc. Biopolymers 97:988–997, 2012.  相似文献   

8.
Sol–gel transition of gelatin was analyzed as a multisite stoichiometric reaction of a gelatin molecule with water and solute molecules. The equilibrium sol–gel transition temperature, Tt, was estimated from the average of gelation and melting temperature measured by differential scanning calorimetry. From Tt and the melting enthalpy, ΔHsol, the equilibrium sol‐to‐gel ratio was estimated by the van't Hoff equation. The reciprocal form of the Wyman–Tanford equation, which describes the sol‐to‐gel ratio as a function of water activity, was successfully applied to obtain a good linear relationship. From this analysis, the role of water activity on the sol–gel transition of gelatin was clearly explained and the contributions of hydration and solute binding to gelatin molecules were separately discussed in sol–gel transition. The general solution for the free energy for gel‐stabilization in various solutions was obtained as a simple function of solute concentration. © 2015 Wiley Periodicals, Inc. Biopolymers 103: 685–691, 2015.  相似文献   

9.
Klebsormidium crenulatum (Kütz.) Lokhorst (Klebsormidiophyceae, Streptophyta) isolated from an alpine soil in Tyrol, Austria, was experimentally exposed to desiccation under various relative air humidities (RH 5, 75, and >95%, ambient air 55%–60%). The effects on the structure and ultrastructure of K. crenulatum after 1, 4, or 7 d of desiccation at 5, 75, and >95% RH were investigated. The cross walls were deformed to an undulated shape, and the cell diameter was reduced to ~60% of the control. Regardless of the RH applied, in all cases the cytoplasm appeared denser compared to that of liquid‐culture‐grown cells. Electron‐dense particles with diameters of 0.4 μm–0.8 μm were observed in the cytoplasm, likely representing lipid droplets. The chloroplasts of desiccated samples contained a large number of plastoglobules. The number and appearance of mitochondria were not visibly altered, as also verified by 3,3′ dihexyloxacarbocyanine iodine (DIOC6) staining. The amphiphilic styryl dye FM 1‐43 resulted in staining of the plasma membrane in cells from liquid culture. In 7 d desiccated samples, a marked fluorescence is seen in ~40%–50% of the cells, which were dead. Actin microfilaments (MFs) were drastically disrupted after desiccation; only dotlike actin batches remained. These results demonstrate that flexibility of the cell walls and maintenance of the key organelles play a key role in the tolerance of desiccation stress in K. crenulatum.  相似文献   

10.
The role of dissolved oxygen as a principal electron acceptor for microbial metabolism was investigated within Fe(III)‐oxide microbial mats that form in acidic geothermal springs of Yellowstone National Park (USA). Specific goals of the study were to measure and model dissolved oxygen profiles within high‐temperature (65–75°C) acidic (pH = 2.7–3.8) Fe(III)‐oxide microbial mats, and correlate the abundance of aerobic, iron‐oxidizing Metallosphaera yellowstonensis organisms and mRNA gene expression levels to Fe(II)‐oxidizing habitats shown to consume oxygen. In situ oxygen microprofiles were obtained perpendicular to the direction of convective flow across the aqueous phase/Fe(III)‐oxide microbial mat interface using oxygen microsensors. Dissolved oxygen concentrations dropped from ~ 50–60 μM in the bulk‐fluid/mat surface to below detection (< 0.3 μM) at a depth of ~ 700 μm (~ 10% of the total mat depth). Net areal oxygen fluxes into the microbial mats were estimated to range from 1.4–1.6 × 10?4 μmol cm?2 s?1. Dimensionless parameters were used to model dissolved oxygen profiles and establish that mass transfer rates limit the oxygen consumption. A zone of higher dissolved oxygen at the mat surface promotes Fe(III)‐oxide biomineralization, which was supported using molecular analysis of Metallosphaera yellowstonensis 16S rRNA gene copy numbers and mRNA expression of haem Cu oxidases (FoxA) associated with Fe(II)‐oxidation.  相似文献   

11.
Studies on biology of Oligonychus mangiferus (Rahman and Sapra) at combination of eight constant temperatures and relative humidities (RHs) viz., 7.0°C with 85% RH, 10°C with 80% RH, 15.0°C with 75% RH, 23.0°C with 70% RH, 31.0°C with 65% RH, 34.0°C with 65% RH, 36.0°C with 60% RH and 40.0°C with 55% RH revealed that the optimal condition for the development of these mites are 15.0–31.0°C and 65–75% RH. The highest temperature and the lowest RH accelerated the rate of development and induced more reproduction of O. mangiferus. Its population also multiplied 30.81 times in a generation time of 27.36 days at 31.0°C and 65% RH, while the same population only increased 7.46 times in a generation time of 48.07 days at 15.0°C and 75% RH. Fecundity was highest at 31.0°C and 65% RH with 46.43 eggs per female. The highest intrinsic rate of natural increase was observed at 31.0°C as 0.125 per day.  相似文献   

12.
For mimicking macromolecular crowding of DNA quadruplexes, various crowding agents have been used, typically PEG, with quadruplexes of micromolar strand concentrations. Thermal and thermodynamic stabilities of these quadruplexes increased with the concentration of the agents, the rise depended on the crowder used. A different phenomenon was observed, and is presented in this article, when the crowder was the quadruplex itself. With DNA strand concentrations ranging from 3 µM to 9 mM, the thermostability did not change up to ~2 mM, above which it increased, indicating that the unfolding quadruplex units were not monomolecular above ~2 mM. The results are explained by self‐association of the G‐quadruplexes above this concentration. The ΔGo37 values, evaluated only below 2 mM, did not become more negative, as with the non‐DNA crowders, instead, slightly increased. Folding topology changed from antiparallel to hybrid above 2 mM, and then to parallel quadruplexes at high, 6–9 mM strand concentrations. In this range, the concentration of the DNA phosphate anions approached the concentration of the K+ counterions used. Volume exclusion is assumed to promote the topological changes of quadruplexes toward the parallel, and the decreased screening of anions could affect their stability. © 2013 Wiley Periodicals, Inc. Biopolymers 101: 428–438, 2014.  相似文献   

13.
Density increments (?ρ/?c2μ in solutions of NaDNA in NaCl and CsDNA in CsCl were determined over a wide range of salt concentrations; calf thymus DNA, fragmented by sonic irradiation to a molecular weight of 4–6 × 105 was used. The partial specific volume v?2° of NaDNA at 25°C was found to ho 0.500 ml/g in water, and that of CsDNA 0.440 ml/g. Both values increase with increasing NaCl and CsCl concentration. Refractive index increments under various experimental conditions were also determined. The relevance of the density increments (at constant, chemical potential of diffusible solutes) to equilibrium sedimentation in a density gradient and the evaluation of molecular weights is discussed. Distribution coefficients of diffusible components, sometimes referred to as preferential solvation or net hydration, were derived from the density increments and partial volumes and compared with direct experimental results, whenever available, from membrane distribution and isopiestic distillation. The thermo-dynamic significance of the distribution coefficients as well as possible interpretations in terms of specific molecular mechanisms are considered.  相似文献   

14.
Vaucheria (Xanthophyceae) exhibited cruciform polarotropism when they were grown under polarized white or blue light for several days. The coexistence of two groups of branches growing perpendicular and parallel to the electric vector (E-vector) resulted in cruciform polarotropic orientation. Such polarotropic bending was, however, not detected within 24 hr. As the fluence rate of polarized white or blue light increased, parallel orientation to the E-vector became dominant. Polarized red light produced exclusively perpendicular polarotropism. This shift in pattern was much obvious in V. terrestris sensu G?tz than V. sessllis and V. dichotoma. Since the photoperception is restricted to the tip of the apical dome and since this region receives maximum photons when the E-vector is at a right angle to the cell axis, Vaucheria becomes oriented normal to the E-vector as far as the fluence rate is optimum. The direction of growth is expected to change into parallel to the E-vector when the fluence rate is supraoptimum. The perpendicular (normal) and parallel polarotropism of Vaucheria, thus, correspond to positive and negative phototropism, respectively. Orientation of photoreceptor molecules is suggested to be predominantly parallel to the surface of the apical dome. Received 14 June 1999/ Accepted in revised form 19 November 1999  相似文献   

15.
The objectives of this study were to investigate the moisture-induced protein aggregation of whey protein powders and to elucidate the relationship of protein stability with respect to water content and glass transition. Three whey protein powder types were studied: whey protein isolate (WPI), whey protein hydrolysates (WPH), and beta-lactoglobulin (BLG). The water sorption isotherms were determined at 23 and 45°C, and they fit the Guggenheim–Andersson–DeBoer (GAB) model well. Glass transition was determined by differential scanning calorimeter (DSC). The heat capacity changes of WPI and BLG during glass transition were small (0.1 to 0.2 Jg−1 °C−1), and the glass transition temperature (T g) could not be detected for all samples. An increase in water content in the range of 7 to 16% caused a decrease in T g from 119 down to 75°C for WPI, and a decrease from 93 to 47°C for WPH. Protein aggregation after 2 weeks’ storage was measured by the increase in insoluble aggregates and change in soluble protein fractions. For WPI and BLG, no protein aggregation was observed over the range of 0 to 85% RH, whereas for WPH, ∼50% of proteins became insoluble after storage at 23°C and 85% RH or at 45°C and ≥73% RH, caused mainly by the formation of intermolecular disulfide bonds. This suggests that, at increased water content, a decrease in the T g of whey protein powders results in a dramatic increase in the mobility of protein molecules, leading to protein aggregation in short-term storage.  相似文献   

16.
San  Phyu Phyu  Tuda  Midori  Takagi  Masami 《BioControl》2021,66(4):497-510

The predatory mite Amblyseius swirskii (Athias-Henriot) (Acari: Phytoseiidae) is currently used as an efficient biological control agent of thrips, whiteflies and spider mites, which are economically damaging pests of ornamental plants and vegetable crops grown in greenhouses and fields worldwide. Currently, the effects of relative humidity (RH) and water availability on the optimal growth of A. swirskii are unknown. Here, we test the combined effects of different levels of RH (33%, 53%, 73% and 92%) and water availability on the development and reproduction of male and female A. swirskii feeding on the dried fruit mite, Carpoglyphus lactis (Linnaeus). While eggs failed to hatch at 33% RH, the survival rates of the immature stages at?≥?53% RH increased solely in response to water availability and not due to changes in RH. Regarding growth and development, low RH extended the egg–adult duration and pre-oviposition period. We also found that the negative effects of low RH on fecundity were partially or completely eliminated when drinking water was available. For the life table parameters, the highest values of net reproductive rate (R0) and intrinsic rate of natural increase (r) were achieved at the highest RH and when drinking water was available. Overall, water availability mitigated the negative effect of low RH on female reproduction, and female development was more sensitive to water availability than male development. Lastly, a comparison of similar research on A. swirskii suggested that water availability and RH are more influential on r than food source or temperature.

  相似文献   

17.
Abstract. The thermal preferences of Alaskozetes antarcticus (Acari, Cryptostigmata) and Cryptopygus antarcticus (Collembola, Isotomidae) were investigated over 6 h within a temperature gradient (?3 to +13 °C), under 100% relative humidity (RH) conditions. After 10 days of acclimation at ?2 or +11 °C, individual supercooling points (SCP) and thermopreferences were assessed, and compared with animals maintained for 10 days under fluctuating field conditions (?6 to +7 °C). Acclimation at ?2 °C lowered the mean SCP of both A. antarcticus (?24.2 ± 9.1) and C. antarcticus (?14.7 ± 7.7) compared to field samples (?19.0 ± 9.0 and ?10.7 ± 5.2, respectively). Acclimation at +11 °C increased A. antarcticus mean SCP values (?13.0 ± 8.5) relative to field samples, whereas those of C. antarcticus again decreased (?16.7 ± 9.1). Mites acclimated under field conditions or at +11 °C selected temperatures between ?3 and +1 °C. After acclimation at ?2 °C, both species preferred +1 to +5 °C. Cryptopygus antarcticus maintained under field conditions preferred +5 to +9 °C, whereas individuals acclimated at +11 °C selected +9 to +13 °C. For A. antarcticus, thermopreference was not influenced by its cold hardened state. The distribution of field specimens was further assessed within two combined temperature and humidity gradient systems: (i) 0–3 °C/12% RH, 3–6 °C/33% RH, 6–9 °C/75% RH and 9–12 °C/100% RH and (ii) 0–3 °C/100% RH, 3–6 °C/75% RH, 6–9 °C/33% RH and 9–12 °C/12% RH. In gradient (i), C. antarcticus distributed homogeneously, but, in gradient (ii), C. antarcticus preferred 0–3 °C/100% RH. Alaskozetes antarcticus selected temperatures between 0 and +6 °C regardless of RH conditions. Cryptopygus antarcticus appears better able than A. antarcticus to opportunistically utilize developmentally favourable thermal microclimates, when moisture availability is not restricted. The distribution of A. antarcticus appears more influenced by temperature, especially during regular freeze‐thaw transitions, when this species may select low temperature microhabitats to maintain a cold‐hardened state.  相似文献   

18.
朱晓艳  韩苗  韩天富  韩峥  王钢 《微生物学报》2022,62(8):3124-3136
【目的】环境中无处不在的气-液界面能够影响细菌的运动和养分的传输扩散,进而调控微生物的种群互作和群落结构。因此,系统地认知微生物在微观界面的运动特征对于理解和解析微生物多样性的产生、维持机制以及生态功能至关重要。【方法】本文基于微流体显微系统(超高速荧光显微镜和数字全息显微镜),以具备主动运动能力的模式菌株铜绿假单胞菌(Pseudomonas aeruginosaPAO1)为研究对象,观测并解析了细菌细胞在气-液界面的二维运动特征和气-液-固界面的二维与三维运动特征。【结果】PAO1既能在气-液界面处执行近似直线轨迹的运动,也能在气-液界面下方执行顺时针或逆时针旋转的圆周运动(最小曲率半径Rmin=3μm)。在气-液-固界面处,6.45%的不运动细胞聚集于气-液-固界面边缘处,并在该处完成不可逆附着;同时,游动细胞由于受到液滴内部毛细管流和马兰戈尼(Marangoni)涡流运动的综合作用,直线游动至距界面约40μm内的区域后,其运动轨迹转变为垂直界面方向返回或以近似界面平行方向运动并附着,这些行为显著调节了PAO1的空间分布,促使了其朝向气-液-固界面的迁移,表明个体PAO1的鞭毛在此...  相似文献   

19.
Unformulated conidia of Beauveria bassiana were stored at five different temperatures (0°, 10°, 20°, 30° and 40°C) at six different relative humidities (RH) (0, 33, 53, 75, 85 and 98%). Conidial viabilities and virulence against third instar larvae of Helicoverpa armigera were determined over a 24‐month period. Conidia survived longest at lower temperatures (0–20°C) and lower RH levels (0–53% RH). At higher temperatures (30–40°C) conidia did not survive. When the temperature was decreased from 30°C to 0°C, at nearly all RH levels the longevity of conidia increased. Conidia remained virulent for third instar larvae of H. armigera under favourable storage conditions for 24 months.  相似文献   

20.
The collagen triple helix has a larger accessible surface area per molecular mass than globular proteins, and therefore potentially more water interaction sites. The effect of deuterium oxide on the stability of collagen model peptides and Type I collagen molecules was analyzed by circular dichroism and differential scanning calorimetry. The transition temperatures (Tm) of the protonated peptide (Pro‐Pro‐Gly)10 were 25.4 and 28.7°C in H2O and D2O, respectively. The increase of the Tm of (Pro‐Pro‐Gly)10 measured calorimetrically at 1.0°C min?1 in a low pH solution from the protonated to the deuterated solvent was 5.1°C. The increases of the Tm for (Gly‐Pro‐4(R)Hyp)9 and pepsin‐extracted Type I collagen were measured as 4.2 and 2.2°C, respectively. These results indicated that the increase in the Tm in the presence of D2O is comparable to that of globular proteins, and much less than reported previously for collagen model peptides [Gough and Bhatnagar, J Biomol Struct Dyn 1999, 17, 481–491]. These experimental results suggest that the interaction of water molecules with collagen is similar to the interaction of water with globular proteins, when the ratio of collagen to water is very small and collagen is monomerically dispersed in the solvent. © 2009 Wiley Periodicals, Inc. Biopolymers 93: 93–101, 2010. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  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