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 343 毫秒
1.
Carboxymethyl cellulose is widely used in many industrial aspects and also in laboratory due to its good biocompatibility. However, special researches on infrared especially aiming at the hydrogen bonds structure of carboxymethyl cellulose were relatively poor. We demonstrate here a full view of infrared spectroscopy in the temperature range of 40–220 °C, mainly aiming at the hydrogen bonds in the NaCMC film. The two important transition points was defined with DSC and together with Infrared analysis, that is, 100 °C corresponding to the complete loss of water molecules and 170 °C to the starting temperature point the O6H6 being oxidized. The series of IR spectra during heating from 40 to 220 °C was analyzed by the two-dimensional correlation method. We found that the water molecules bound with CO groups and OH groups. With the evaporating of water molecules, the hydrated CO groups gradually transited into non-hydrated CO groups. As the temperature continued to increase, the intrachain hydrogen bonds were weakened and transited into weak hydrogen bonds. When the temperature was higher than 170 °C, the O6H6 groups were gradually oxidized and thus the interchain hydrogen bonds formed between CH2COONa groups and O6H6 were weakened. In summary, we defined the main sorts of hydrogen bonds in carboxymethyl cellulose and pictured the changes of the hydrogen bonds structure during heating process, which may provide for the further application in both industry aspects and laboratory use.  相似文献   

2.
The crystal structure of beta-D-glucopyranosyl-(1-->4)-alpha-D-glucopyranose (alpha-cellobiose) in a complex with water and NaI was determined with Mo K(alpha) radiation at 150 K to R=0.027. The space group is P2(1) and unit cell dimensions are a=9.0188, b=12.2536, c=10.9016 A, beta=97.162 degrees. There are no direct hydrogen bonds among cellobiose molecules, and the usual intramolecular hydrogen bond between O-3 and O-5' is replaced by a bridge involving Na+, O-3, O-5', and O-6'. Both Na+ have sixfold coordination. One I(-) accepts six donor hydroxyl groups and three C-H***I(-) hydrogen bonds. The other accepts three hydroxyls, one Na+, and five C-H***I(-) hydrogen bonds. Linkage torsion angles phi(O-5) and psi(C-5) are -73.6 and -105.3 degrees, respectively (phi(H)=47.1 degrees and psi(H)=14.6 degrees ), probably induced by the Na+ bridge. This conformation is in a separate cluster in phi,psi space from most similar linkages. Both C-6-O-H and C-6'-O-H are gg, while the C-6'-O-H groups from molecules not in the cluster have gt conformations. Hybrid molecular mechanics/quantum mechanics calculations show <1.2 kcal/mol strain for any of the small-molecule structures. Extrapolation of the NaI cellobiose geometry to a cellulose molecule gives a left-handed helix with 2.9 residues per turn. The energy map and small-molecule crystal structures imply that cellulose helices having 2.5 and 3.0 residues per turn are left-handed.  相似文献   

3.
Native cellulose in higher plants forms crystalline fibrils a few nm across, with a substantial fraction of their glucan chains at the surface. The accepted crystal structures feature a flat-ribbon 21 helical chain conformation with every glucose residue locked to the next by hydrogen bonds from O-3' to O-5 and from O-2 to O-6'. Using solid-state NMR spectroscopy we show that the surface chains have a different C-6 conformation so that O-6 is not in the correct position for the hydrogen bond from O-2. We also present evidence consistent with a model in which alternate glucosyl residues are transiently or permanently twisted away from the flat-ribbon conformation of the chain, weakening the O-3' - 0-5 hydrogen bond. Previous molecular modelling and the modelling studies reported here indicate that this 'translational' chain conformation is energetically feasible and does not preclude binding of the surface chains to the interior chains, because the surface chains share the axial repeat distance of the 21 helix. Reduced intramolecular hydrogen bonding allows the surface chains to form more hydrogen bonds to external molecules in textiles, wood, paper and the living plant.  相似文献   

4.
The FTIR spectra were measured for raw Uplands Sicala-V2 cotton fibers over a temperature range of 40-325 degrees C to explore the temperature-dependent changes in the hydrogen bonds of cellulose. These cotton-cellulose spectra exhibited complicated patterns in the 3800-2800 cm(-1) region and thus were analyzed by both the exploratory principal component analysis (PCA) and two-dimensional (2-D) correlation spectroscopy methods. The exploratory PCA showed that the spectra separate into two groups on the basis of thermal degradation of the cotton-cellulose and the consequent breakage of intersheet H-bonds present in its structure. Frequency variables, which strongly contributed to each principal component highlighted in its loadings plot, were linked to the frequencies assigned to vibrations of the OH groups involved in different kinds of H-bonds, as well as to vibrations of the CH groups. Deeper insights into reorganization of the temperature-dependent hydrogen bonding were obtained by 2-D correlation spectroscopy. Synchronous and asynchronous spectra were analyzed in the temperature ranges of 40 to 150 and 250 to 320 degrees C, the ranges indicated by PCA. Detailed band assignments of the OH stretching region and changes in the patterns of the hydrogen bonding network of the cotton-cellulose were proposed with the aid of the 2-D correlation spectroscopy analysis. Below 150 degrees C, distinctly different bands assigned to the less stable Ialpha and the more stable Ibeta interchain H-bonds O-6-H-6...O-3' were observed at about 3230 and 3270 cm(-1), respectively. Evaporation of water entrapped in the cellulose network was examined by means of the band at about 3610 cm(-1). The cooperativity of hydrogen bonds, which play a key role in the cellulose conformation, was monitored by frequencies assigned to intrachain H-bonds. It was possible to separate the frequencies assigned to the O-2-H-2...O-6 and O-3-H-3...O-5 intrachain H-bonds into two separate ranges, the spread of which was controlled by the cooperativity effect. The temperature dependence of the asynchronous spectra indicated that the less stable O-3-H-3...O-5 bonds gave rise to an absorption extending from 3300 to 3384 cm(-1), while the more stable O-2-H-2...O-6 bonds were characterized by the absorption between 3400 and 3470 cm(-1). The final breaking of the inter- and intrachain H-bonds, which occurs at the higher temperatures, was monitored by the asynchronous peaks at 3533 and 3590 cm(-1), respectively. On the basis of both the exploratory PCA and 2-D correlation spectroscopy investigations, it was possible to extract well-defined wavenumber ranges assigned to different kinds of intra- and interchain hydrogen bonds, as well as to the free OH groups of the cotton-cellulose.  相似文献   

5.
Cotton fiber cellulose is highly crystalline and oriented; when native cellulose (cellulose I) is treated with certain alkali concentrations, intermolecular hydrogen bonds are broken and Na-cellulose I is formed. At higher alkali concentrations Na-cellulose II forms, wherein intermolecular and intramolecular hydrogen bonds are broken, ultimately resulting in cellulose II polymers. Crystallinity changes in cotton fibers were observed and assigned using attenuated total reflectance Fourier transform infrared (ATR FT-IR) spectroscopy and X-ray diffraction (XRD) subsequent to sodium hydroxide treatment and compared with an in situ protein-binding methodology using cellulose-directed carbohydrate-binding modules (CBMs). Crystallinity changes observed using CBM probes for crystalline cellulose (CBM2a, CBM3a) and amorphous cellulose (CBM4-1, CBM17) displayed close agreement with changes in crystallinity observed with ATR-FTIR techniques, but it is notable that crystallinity changes observed with CBMs are observed at lower NaOH concentrations (2.0 mol dm(-3)), indicating these probes may be more sensitive in detecting crystallinity changes than those calculated using FTIR indices. It was observed that the concentration of NaOH at which crystallinity changes occur as analyzed using the CBM labeling techniques are also lower than those observed using X-ray diffraction techniques. Analysis of crystallinity changes in cellulose using CBMs offers a new and advantageous method of qualitative and quantitative assessment of changes to the structure of cellulose that occur with sodium hydroxide treatment.  相似文献   

6.
Vibrational spectroscopy using polarized incident radiation can be used to determine the orientation of X-H bonds with respect to coordinates such as crystallographic axes. The adaptation of this approach to polymer fibers is described here. It requires spectral intensity to be quantified around a 180 degrees range of polarization angles and not just recorded transversely and longitudinally as is normal in fiber spectroscopy. Mercerized cellulose II is used as an example. The unit cell of the cellulose II lattice contains six distinct hydroxyl groups engaged in a complex network of hydrogen bonds that hold the cellulose chains laterally together. A formalism is described to relate the variation in intensity of each O-H stretching mode to the angle between its transition moment and the chain axis as the polarization axis is rotated with respect to the fiber axis. It was necessary to include the effect of dispersion in chain orientation around the mean and the averaging of all rotational positions of the chains round their axis. The two crystallographically distinct O(2)-H groups, which are each hydrogen-bonded to only one acceptor oxygen, show a close match in orientation between the transition moments of their stretching bands and the O-H bond axis. The two O(3)-H groups each have a three-centered hydrogen bond to O-5 and O-6 of the next residue in the same chain. The transition moments of their stretching modes lay between the acceptor oxygens. Hydrogen bonding from the O(6)-H groups is still more complex but again the transition moment of each O-H bond lay within the cone of orientations described by the acceptor oxygens, provided that one additional acceptor oxygen excluded from the published crystal structure was considered. The transition moments for the O-H stretching modes were approximately aligned with the O-H bond axes, but the alignment was not necessarily exact. This approach is not restricted to hydroxyl groups, but it is particularly useful for the elucidation of hydrogen bonding in fibrous polymers for which crystallographic data on proton positions are not available.  相似文献   

7.
For the first time, 2-O-methyl- (2MC) and 3,6-di-O-methyl-cellulose (36MC) were synthesized via 3-O-allyl- and 3-O-methyl-cellulose, respectively. Position 6 of 3-O-allyl- and 3-O-methyl-cellulose was protected with the 4-methoxytrityl groups. The reaction time and temperature were optimized to achieve a high regioselectivity at C-6 and to prevent the introduction of the 4-methoxytrityl group at C-2 of the polymer. It was found that the substituent at C-3 of 3-O-functionalized celluloses influenced the reactivity of the hydroxyl group at C-6. The structure was characterized by means of (1)H and (13)C NMR spectroscopy of the acetates of 2MC and 36MC. 2MC and 36MC were soluble in water and did not show thermoreversible gelation.  相似文献   

8.
The binding of galactose-specific lectins from Erythrina indica (EIL), Erythrina arborescens (EAL), Ricinus communis (agglutinin; RCA-I), Abrus precatorius (agglutinin; APA), and Bandeiraea simplicifolia (lectin I; BSL-I) to fluoro-, deoxy-, and thiogalactoses were studied in order to determine the strength of hydrogen bonds between the hydroxyl groups of galactose and the binding sites of the proteins. The results have allowed insight into the nature of the donor/acceptor groups in the lectins that are involved in hydrogen bonding with the sugar. The data indicate that the C-2 hydroxyl group of galactose is involved in weak interactions as a hydrogen-bond acceptor with uncharged groups of EIL and EAL. With RCA-I, the C-2 hydroxyl group forms two weak hydrogen bonds in the capacity of a hydrogen-bond acceptor and a donor. On the other hand, there is a strong hydrogen bond between the C-2 hydroxyl group of galactose, which acts as a donor, and a charged group on BSL-I. The C-2 hydroxyl group of the sugar is also a hydrogen-bond donor to APA. The lectins are involved in strong hydrogen bonds through charged groups with the C-3 and C-4 hydroxyl groups of galactose, with the latter serving as hydrogen-bond donors. The C-6 hydroxyl group of the sugar is weakly hydrogen bonded with neutral groups of EIL, EAL, and APA. With BSL-I, however, a strong hydrogen bond is formed at this position with a charged group of the lectin. The C-6 hydroxyl groups is a hydrogen-bond acceptor for EIL and EAL, a hydrogen-bond donor for APA and BSL-I, and appears not to be involved in binding to RCA-I. The data with the thiosugars indicate the involvement of the C-1 hydroxyl group of galactose in binding to EIL, EAL, and BSL-I, but not to RCA-I and APA. We have also performed a similar analysis of the binding data of fluoro- and deoxysugars to concanavalin A [Poretz, R. D. and Goldstein, I. J. (1970) Biochemistry 9, 2890-2896]. This has allowed comparison of the donor/acceptor properties and free energies of hydrogen bonding of the hydroxyl groups of methyl alpha-D-mannopyranoside to concanavalin A with the results in the present study. On the basis of this analysis, new assignments are suggested for amino acid residues of concanavalin A [corrected] that may be involved in hydrogen bonding to the sugar.  相似文献   

9.
The X-ray crystal structures of three monosaccharide derivatives prepared by the reaction of sulfanilamide with D-ribose, D-arabinose, and D-mannose have been determined. The derivatives are N-(p-sulfamoylphenyl)-alpha-D-ribopyranosylamine (1), N-(p-sulfamoylphenyl)-alpha-D-arabinopyranosylamine (2), and N-(p-sulfamoylphenyl)-beta-D-mannopyranosylamine monohydrate (3). The monosaccharide ring of 1 and 2 has the 1C4 conformation, stabilized in 1 by an intramolecular hydrogen bond from 0-2 to 0-4. Compound 3 has the 4C1 conformation at the monosaccharide ring and the gt conformation at the C-6-O-6 side chain. Occupancy of the water molecule in the crystal of 3 actually examined was 22%. The degree of interaction between sulfamoyl groups and monosaccharide moieties varies from structure to structure. The packing arrangement of 2 involves hydrogen bonding between sulfamoyl groups and monosaccharide hydroxyl groups, but interactions of this type are fewer in 1, and in 3 the hydrogen bonds are either strictly between monosaccharide hydroxyl groups or strictly between sulfamoyl groups. Pairs of hydrogen bonds (two-point contacts) link neighboring molecules in all three structures, between screw-axially related molecules in 1 and 2 and between translationally related molecules in 3. The contact in 3 defined by the O-3-H...O-5 and O-6-H...O-4 hydrogen bonds is found in several other N-aryl-beta-D-mannopyranosylamine crystal structures and is apparently an especially favorable mode of intermolecular interaction in these compounds.  相似文献   

10.
The hydrogen bonding arrangement in anhydrous β-chitin, a homopolymer of N-acetylglucosamine, was directly determined by neutron fiber diffraction. Data were collected from a sample prepared from the bathophilous tubeworm Lamellibrachia satsuma in which all labile hydrogen atoms had been replaced by deuterium. Initial positions of deuterium atoms on hydroxyl and acetamide groups were directly located in Fourier maps synthesized using phases calculated from the X-ray structure and amplitudes measured from the neutron data. The hydrogen bond arrangement in the refined structure is in general agreement with predictions based on the X-ray structure: O3 donates a hydrogen bond to the O5 ring oxygen atom of a neighboring residue in the same chain; N2 and O6 donate hydrogen bonds to the same carbonyl oxygen O7 of an adjacent chain. The intramolecular O3···O5 hydrogen bond has the most energetically favorable geometry with a hydrogen to acceptor distance of 1.77 ? and a hydrogen bond angle of 171°.  相似文献   

11.
IR spectra (1600-1800 and 3000-3650 cm-1) of lincomycin base solutions in inert (CCl4 and C2Cl4), proton acceptor (dioxane, dimethylsulfoxide and triethyl amine) and proton donor (CHCl3, CD3OD and D2O) solvents were studied. Analysis of the concentration and temperature changes in the spectra revealed that association in lincomycin in the inert solvents was due to intramolecular hydrogen linkage involving amide and hydroxyl groups. Disintegration of the associates after the solution dilution and temperature rise was accompanied by formation of intramolecular bonds stabilizing the stable conformation structure of the lincomycin molecule. The following hydrogen linkage in the conformation was realized: NH...N (band v NH...N at 3340 cm-1), OH...O involving the hydroxyl at C-7 and O atoms in the D-galactose ring (band v OH...O at 3548 cm-1), a chain of the hydrogen bonds OH...OH...OH in the lincomycin carbohydrate moiety (band v OH...O at 3593 cm-1 and v OH of the end hydroxyl group at 3625 cm-1). Bonds NH and C-O of the amide group were located in transconformation. Group C-O did not participate in the intramolecular hydrogen linkage.  相似文献   

12.
Josefin E. Utas  Dick Sandström 《BBA》2006,1757(12):1592-1596
The X-ray crystal structure of the mono-hydrate of 2,2-bis(imidazol-1-ylmethyl)-4-methylphenol has been determined. Three hydrogen bonds hold water very tightly in the crystal, as determined by deuterium solid-state NMR. The hydrogen bond between the phenolic hydroxyl and water appears to have about the same strength as the direct hydrogen bond to imidazole, suggesting that the structure can be a good model for hydrogen bonds that are mediated by a water molecule in enzymes.  相似文献   

13.
The distribution of carboxymethyl substituents in the alpha-(1 --> 6)-linked maltotriosyl repeating units of a carboxymethylpullulan (CMP) series was investigated by high resolution NMR spectroscopy on very short oligomers (DPn = 1.2-1.5) obtained by acid hydrolysis. A series of 2D NMR experiments on parent pullulan, hydrolysed pullulan and CMP was used to assign the proton and carbon chemical shifts of CMP acid hydrolysates. The degree of substitution (DS) and the relative distribution of -CH2COONa groups at OH-2, OH-3, OH-4 and OH-6 of glucose residues (DSi) were determined from 1H NMR measurements. From a set of CMP samples, widely different in degree of substitution, it was observed that the substitution at C-2 is predominant and decreases according to the order C-2 > C-3 > C-6 > C-4. Taking into account the availability of each OH group in the parent pullulan, an order of relative reactivity of hydroxyl groups is defined according to the relation: Ri = DSi/ni, where ni is the number of free OH groups in a maltotriose unit (MTU) for a given site C-i, the reactivity order was found to be OH-2 > OH-4 > OH-6 > OH-3.  相似文献   

14.
The thermotropic phase behaviour of the ceramide N-octadecanoylphytosphingosine (CER3) was investigated using differential scanning calorimetry, X-ray powder diffraction and FT-IR spectroscopy. CER3 was shown to be a polymorphic substance depending on the crystallisation conditions. Three different solid states were found. The FT-IR results elucidate changes in the hydrogen bonding interactions of the ceramide head group. It was shown that the amide I and the amide II vibration bands are quite sensitive to the phase transitions of CER. There are clear shifts in the band positions of those bands passing the phase transitions. Furthermore, changes were observed in the NH- and OH- stretching region. The study shows that there are strong inter- and intramolecular hydrogen bonds between hydroxy groups in the ceramide head group. There are also strong hydrogen bonds to the amide oxygen as shown by the band positions of the amide vibrations. The H-bonding network and conformation of the head group of CER3 alters due to the phase transitions.  相似文献   

15.
The structure of bacterial cellulose is affected by the bacterial strain used, culture media and cultivation conditions. In this study, acid-treated multi-walled carbon nanotubes (MWNTs) were added into a static culture medium and their effect on bacterial cellulose structure was studied by scanning electron microscopy (SEM), atomic force microscopy (AFM), Fourier transform infrared spectroscopy (FT-IR), CP/MAS (13)C NMR and X-ray diffractometry. The bacterial cellulose ribbons and the MWNTs interwound and formed a three-dimensional network architecture. Band-like assemblies with sharp bends and rigidity were also produced in the presence of MWNTs. The intermolecular hydrogen bonds in bacterial cellulose produced in the presence of MWNTs were weakened. The crystal structure, cellulose I(alpha) content, crystallinity index (CrI) and crystallite size all changed. The results may suggest that the acid-treated MWNTs containing hydroxyl groups interact with the sub-elementary bacterial cellulose fibrils, subsequently interfering with the aggregation and crystallization.  相似文献   

16.
We determined the crystal structure of anhydrous chitosan at atomic resolution, using X‐ray fiber diffraction data extending to 1.17 Å resolution. The unit cell [a = 8.129(7) Å, b = 8.347(6) Å, c = 10.311(7) Å, space group P212121] of anhydrous chitosan contains two chains having one glucosamine residue in the asymmetric unit with the primary hydroxyl group in the gt conformation, that could be directly located in the Fourier omit map. The molecular arrangement of chitosan is very similar to the corner chains of cellulose II implying similar intermolecular hydrogen bonding between O6 and the amine nitrogen atom, and an intramolecular bifurcated hydrogen bond from O3 to O5 and O6. In addition to the classical hydrogen bonds, all the aliphatic hydrogens were involved in one or two weak hydrogen bonds, mostly helping to stabilize cohesion between antiparallel chains. © 2016 Wiley Periodicals, Inc. Biopolymers 105: 361–368, 2016.  相似文献   

17.
The X-ray diffraction analysis, (13)C CP MAS NMR spectra and powder X-ray diffraction patterns were obtained for selected methyl glycosides: alpha- and beta-d-lyxopyranosides (1, 2), alpha- and beta-l-arabinopyranosides (3, 4), alpha- and beta-d-xylopyranosides (5, 6) and beta-d-ribopyranoside (7) and the results were confirmed by GIAO DFT calculations of shielding constants. In X-ray diffraction analysis of 1 and 2, a characteristic shortening and lengthening of selected bonds was observed in molecules of 1 due to anomeric effect and, in crystal lattice of 1 and 2, hydrogen bonds of different patterns were present. Also, an additional intramolecular hydrogen bond with the participation of ring oxygen atom was observed in 1. The observed differences in chemical shifts between solid state and solution come from conformational effects and formation of various intermolecular hydrogen bonds. The changes in chemical shifts originating from intermolecular hydrogen bonds were smaller in magnitude than conformational effects. Furthermore, the powder X-ray diffraction (PXRD) performed for 4, 5 and 7 revealed that 7 existed as a mixture of two polymorphs, and one of them probably consisted of two non-equivalent molecules.  相似文献   

18.
The properties of cellulose materials are dependent on interactions between and within the cellulose chains. To investigate the deformation behavior of cellulose and its relation to molecular straining, sheets with fibers oriented preferably in one direction were studied by dynamic FT-IR spectroscopy. Celluloses with different origins (spruce pulp, Cladophora cellulose, cotton linters) were used. The sheets were stretched sinusoidally at low strains and small amplitudes while being irradiated with polarized infrared radiation. The cellulose fibers showed mainly an elastic response. The cellulose fibers showed mainly an elastic response. The glucose rings and the C-O-C bridges connecting adjacent rings, as well as the O(3)H.O(5) intramolecular hydrogen bonds are the components mainly deformed under stress, whereas the O(2)H.O(6) intramolecular hydrogen bonds play a minor role. The load distribution was also found to be different in the different allomorphic forms of cellulose I, namely, I(alpha) and I(beta).  相似文献   

19.
The relationship between the physicochemical properties of lignocellulosic substrates and enzyme digestion is still not well known. After different pretreatments, cellulase hydrolysis and measurements of physicochemical characteristics by column solute exclusion, particle size analysis, X‐ray diffraction, Fourier transform infrared spectroscopy and solid state 13C nuclear magnetic resonance were performed in this study. Partial least squares was then applied to seek the key factors limiting the rate and extent of cellulose digestion. According to the PLS results, the most important factor for cellulose digestion was accessible interior surface area, followed by delignification and the destruction of the hydrogen bonds. The cellulose digestion at 2 and 24 hr were improved with the increased accessibility of interior surface area to the reporter molecules of 5.1‐nm diameter. Removal of lignin and breaking of hydrogen bonds were also found to significantly promote cellulose conversion. Other properties, including the breakdown of intramolecular hydrogen bonds, cellulose crystallinity, and hemicellulose content, had less effect on the efficiency of enzymatic hydrolysis. © 2009 American Institute of Chemical Engineers Biotechnol. Prog., 2010  相似文献   

20.
Gibberellins (GAs) are phytohormones regulating various developmental processes in plants. In rice, the initial GA-signaling events involve the binding of a GA to the soluble GA receptor protein, GID1. Although X-ray structures for certain GID1/GA complexes have recently been determined, an examination of the complexes does not fully clarify how GID1s discriminate among different GAs. Herein, we present a study aimed at defining the types of forces important to binding via a combination of isothermal titration calorimetry (ITC) and computational docking studies that employed rice GID1 (OsGID1), OsGID1 mutants, which were designed to have a decreased possible number of hydrogen bonds with bound GA, and GA variants. We find that, in general, GA binding is enthalpically driven and that a hydrogen bond between the phenolic hydroxyl of OsGID1 Tyr134 and the C-3 hydroxyl of a GA is a defining structural element. A hydrogen-bond network that involves the C-6 carboxyl of a GA that directly hydrogen bonds the hydroxyl of Ser198 and indirectly, via a two-water-molecule network, the phenolic hydroxyl of Tyr329 and the NH of the amide side-chain of Asn255 is also important for GA binding. The binding of OsGID1 by GA(1) is the most enthalpically driven association found for the biologically active GAs evaluated in this study. This observation might be a consequence of a hydrogen bond formed between the hydroxyl at the C-13 position of GA(1) and the main chain carbonyl of OsGID1 Phe245. Our results demonstrate that by combining ITC experiments and computational methods much can be learned about the thermodynamics of ligand/protein binding.  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