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 31 毫秒
1.
In the first paper heat capacities Cp, of polyglycine, poly(L -alanine), and poly (L -valine) were analyzed using approximate group vibrations and fitting the Cp contributions of the skeletal vibrations to a two-parameter Tarasov function. In this second paper all other poly (amino acid) s are similarly analyzed. Heat capacities were measured by differential scanning calorimetry in the temperature range of 230–390 K for poly(L -leucine), poly(L -serine), poly (sodium-L -aspartate), poly(sodium-L -glutamate), poly(L -asparagine), poly(L -phenylalanine), poly(L -tyrosine), poly(L -methionine), poly (L -tryptophane), poly(L -proline), poly(L -lysine · HBr), poly(L -histidine), poly(L -histidine- HCl), and poly(L -arginine · HCl). Good agreement exists between experiment and calculation. Predictions of heat capacities were made for all not-measured poly (amino acid) s. Enthalpies, entropies, and Gibbs functions for the solid state have been derived. © 1993 John Wiley & Sons, Inc.  相似文献   

2.
M Daurel  P Delhaes  E Dupart 《Biopolymers》1975,14(4):801-823
The heat capacities of L -Alanine, tri(L -alanine), and poly (L -alanine) (α helicoidal form and β pleated sheet structure) have been measured between 1.5 and 300°K with a standard adiabatic calorimeter. In the solid state, the heat capacity is in general dut to three parts which are additive in first-order approximation. (1) The lattice vibrations or “acoustical modes” which are the largest at low temperatures. The low-temperature lattice specific heat is proportional to T, T2, or T3 for an ideal one-, two- or three-dimensional solid, respectively. (2) The so-called group vibrations or “optical modes” which, due to their high frequencies, usually take effect only at higher temperatures. (3) The defects and unharmonic effects. The α-amino acid and its trimer present a specific heat thermal variation characteristic of molecular solids which is correctly fitted with an empirical law proposed by Kitaigorodskii. This author assumes that, for such solids, the molecular lattice point has six degrees of freedom (three of translation and three of rotation). Thus the lattice contribution of the specific heat satisfies the Debye approximation in agreement with the doubling of the number of degrees of freedom per molecule (compared with atomic crystals). The polypeptide behavior is, however, different. The specific heat for each form exhibits a thermal dependence connected with a strong vibrational anisotropy. The model proposed earlier by Tarasov accounts well for these results. In the case of the β form, we have observed the predicted three- and two-dimensional behavior due to the intermolecular H bondings responsible for the sheet structure. For the α form we observed a one-dimensional pattern at higher temperature, since each peptidic chain vibrates separately. The comparison with other spectroscopic and theoretical investigations shows a large discrepancy. However, we have attemped to account for the “optical contribution” to the specific heat of poly(L -alanine) by using a continuum of mean frequencies as suggested by Wunderlich. Vibrational frequency spectra are proposed to explain our results, but the overlapping of acoustical and optical branches in the case of the α form outlines the limits of macroscopic models. It is quite likely that the acoustical spectrum is greatly affected by the intramolecular H bonding. At low temperature the specific heat is a physical property sensitive to the long-range order of the macromolecule, and therefore further spectroscopic and theoretical investigations are necessary to explain correctly these experimental results.  相似文献   

3.
U Hahn  H Hanssum  H Rüterjans 《Biopolymers》1985,24(7):1147-1156
The anisotropic rotational motion of the backbone and the side chains of poly(L -glutamic acid) in the α-helical structure was investigated using the 13C-T1 and T2 relaxation times of all carbon atoms with directly attached protons, obtained at a 13C-Larmor frequency of 67.89 MHz. The evaluation of the nmr data was carried out according to the previously derived anisotropic diffusion model, in which the macromolecule is considered a rigid rod. The rotation of the backbone is characterized by two diffusion constants, D1 and D3, describing the rotation perpendicular to and around the symmetry axis. The additional internal motion of the Cβ-methylene group is described as a jump process with a jump rate, k1, between two allowed rotametric states. Steric considerations indicate that the occupation of the third rotameric position is forbidden. The rotation of the Cγ-methylene group is decribed as a one-dimensional diffusion process around the Cβ–Cγ bond. Investigation of the temperature dependence of the relaxation parameters led to the temperature dependence of the dynamic parameters. Activation energies were determined from these data. The dynamic parameters obtained for poly(L -glutamic acid) at 291 K are compared with the corresponding results of a previous study of poly(L -lysine). The development of an anisotropic diffusion model for the motions of the rod-shaped poly(L -lysine) α-helix and its application to the interpretation of the 13C-relaxation data of this molecule have already been published previously. In this model, both the overall molecular tumbling and the various internal motions have been characterized by diffusion constants or jump rates typical for each process. These dynamic parameters can be calculated from the spin–lattice relaxation times, the spin–spin relaxation times and the NOE factors of the Cα, Cβ, and Cγ nuclei of the polypetide. In the present paper, we describe the application of the above-mentioned dynamic model to the interpretation of 13C-relaxation studies of a further homopolypeptide, poly(L -glutamic acid), in the α-helical structure. Furthermore, we studied the temperature dependence of the relaxation times of this polymer and determined the anisotropic diffusion parameters at each temperature. From their temperature dependence and from comparison of our present results with the data of our previous study of poly(L -lysine), we were able to derive new insights into the intramolecular diffusion processes and the excitation of various motions.  相似文献   

4.
In this paper two points are considered: the methods of evaluating the helical content θ and the calculation of the parameters of the transition from experimental data and its interpretation. The parameter ΔH obtained is in good agreement with the calorimetric one and v is found to be independent of temperature and solvent and in agreement with the ordinarily accepted value for poly(γ-benzyl-L -glutamate). The different methods of estimating θ are discussed for both polypeptides.  相似文献   

5.
Iwao Satake  Jen Tsi Yang 《Biopolymers》1976,15(11):2263-2275
The binding isotherms of sodium decyl sulfate to poly(L -ornithine), poly(D ,L -ornithine), and poly(L -lysine) at neutral pH were determined potentiometrically. The nature of a highly cooperative binding in all three cases suggests a micelle-like clustering of the surfactant ions onto the polypeptide side groups. The hydrophobic interaction between the nonpolar groups overshadows the coulombic interaction between the charged groups. The titration curves can be interpreted well by the Zimm–Bragg theory. The average cluster size of bound surfactant ions is sufficiently large to promote the β-structure of (L -Lys)n even at a very low binding ratio of surfactant to polypeptide residue, whereas the onset of the helical structure for (L -Orn)n begins after about 7 surfactant ions are bound to two turns of the helix. The CD results are consistent with this explanation.  相似文献   

6.
The helix-coil transition temperature Tc of poly(γ-benzyl L -glutamate) in binary solvent mixtures of dichloroacetic acid and 1,4-dichlorobutane, 1-chlorooctane, or 1-chlorododecane have been measured. A treatment is presented with which the transition enthalpy can be calculated from the observed dependence of Tc on solvent composition. Results are compared with previously obtained calorimetric data. The underlying assumptions of the calculation are discussed.  相似文献   

7.
31P- and 1H-nmr and laser Raman spectra have been obtained for poly[d(G-T)]·[d(C-A)] and poly[d(A-T)] as a function of both temperature and salt. The 31P spectrum of poly[d(G-T)]·[d(C-A)] appears as a quadruplet whose resonances undergo separation upon addition of CsCl to 5.5M. 1H-nmr measurements are assigned and reported as a function of temperature and CsCl concentration. One dimensional nuclear Overhauser effect (NOE) difference spectra are also reported for poly[d(G-T)]·[d(C-A)] at low salt. NOE enhancements between the H8 protons of the purines and the C5 protons of the pyrimidines, (H and CH3) and between the base and H-2′,2″ protons indicate a right-handed B-DNA conformation for this polymer. The NOE patterns for the TH3 and GH1 protons in H2O indicate a Watson–Crick hydrogen-bonding scheme. At high CsCl concentrations there are upfield shifts for selected sugar protons and the AH2 proton. In addition, laser Raman spectra for poly[d(A-T)] and poly[d(G-T)]·[d(C-A)] indicate B-type conformations in low and high CsCl, with predominantly C2′-endo sugar conformations for both polymers. Also, changes in base-ring vibrations indicate that Cs+ binds to O2 of thymine and possibly N3 of adenine in poly[d(G-T)]·[d(C-A)] but not in poly[d(A-T)]. Further, 1H measurements are reported for poly[d(A-T)] as a function of temperature in high CsCl concentrations. On going to high CsCl there are selective upfield shifts, with the most dramatic being observed for TH1′. At high temperature some of the protons undergo severe changes in linewidths. Those protons that undergo the largest upfield shifts also undergo the most dramatic changes in linewidths. In particular TH1′, TCH3, AH1′, AH2, and TH6 all undergo large changes in linewidths, whereas AH8 and all the H-2′,2″ protons remain essentially constant. The maximum linewidth occurs at the same temperature for all protons (65°C). This transition does not occur for d(G-T)·d(C-A) at 65°C or at any other temperature studied. These changes are cooperative in nature and can be rationalized as a temperature-induced equilibrium between bound and unbound Cs+, with duplex and single-stranded DNA. NOE measurements for poly[d(A-T)] indicate that at high Cs+ the polymer is in a right-handed B-conformation. Assignments and NOE effects for the low-salt 1H spectra of poly[d(A-T)] agree with those of Assa-Munt and Kearns [(1984) Biochemistry 23 , 791–796] and provide a basis for analysis of the high Cs+ spectra. These results indicate that both polymers adopt a B-type conformation in both low and high salt. However, a significant variation is the ability of the phosphate backbone to adopt a repeat dependent upon the base sequence. This feature is common to poly[d(G-T)]·[d(C-A)], poly[d(A-T)], and some other pyr–pur polymers [J. S. Cohen, J. B. Wouten & C. L Chatterjee (1981) Biochemistry 20 , 3049–3055] but not poly[d(G-C)].  相似文献   

8.
Poly-β-benzyl-L -aspartate (poly[Asp(OBzl)]) forms either a lefthanded α-helix, β-sheet, ω-helix, or random coil under appropriate conditions. In this paper the Raman spectra of the above poly[Asp(OBzl)] conformations are compared. The Raman active amide I line shifts from 1663 cm?1 to 1679 cm?1 upon thermal conversion of poly[Asp(OBzl)] from the α-helical to β-sheet conformation while an intense line appearing at 890 cm?1 in the spectrum of the α-helix decreases in intensity. The 890 cm?1 line also displays weak intensity when the polymer is dissolved in chloroform–dichloroacetic acid solution and therefore is converted to the random coil. This line probably arises from a skeletal vibration and is expected to be conformationally sensitive. Similar behavior in the intensity of skeletal vibrations is discussed for other polypeptides undergoing conformational transitions. The Raman spectra of two cross-β-sheet copolypeptides, poly(Ala-Gly) and poly(Ser-Gly), are examined. These sequential polypeptides are model compounds for the crystalline regions of Bombyx mori silk fibroin which forms an extensive β-sheet structure. The amide I, III, and skeletal vibrations appeared in the Raman spectra of these polypeptides at the frequencies and intensities associated with β-sheet homopolypeptides. Since the sequential copolypeptides are intermediate in complexity between the homopolypeptides and the proteins, these results indicate that Raman structure–frequency correlations obtained from homopolypeptide studies can now be applied to protein spectra with greater confidence. The perturbation scheme developed by Krimm and Abe for explaining the frequency splitting of the amide I vibrations in β-sheet polyglycine is applied to poly(L -valine), poly-(Ala-Gly), poly(Ser-Gly), and poly[Asp(OBzl)]. The value of the “unperturbed” frequency, V0, for poly[Asp(OBzl)] was significantly greater than the corresponding values for the other polypeptides. A structural origin for this difference may be displacement of adjacent hydrogen-bonded chains relative to the standard β-sheet conformation.  相似文献   

9.
Poly(L -arginine) assumes the α-helix in the presence of the tetrahedral-type anions or some polyanions by forming the “ringed-structure bridge” between guanidinium groups and anions which is stabilized by a pair of hydrogen bonds and electrostatic interaction [Ichimura, S., Mita, K. & Zama, M. (1978) Biopolymers 17 , 2769–2782; Mita, K., Ichimura, S. & Zama, M. (1978) Biopolymers 17 , 2783–2798]. This paper describes the parallel CD studies on the conformational effects on poly (L -homoarginine) of various mono-, di-, polyvalent anions and some polyanions, as well as alcohol and sodium dodecylsulfate. The random coil to α-helix transition of poly(L -homoarginine) occurred only in NaClO4 solution or in the presence of high content of ethanol or methanol. The divalent and polyvalent anions of the tetrahedral type (SO, HPO, and P2O), which are strong α-helix-forming agents for poly(L -arginine), failed to induce the α-helical conformation of poly(L -homoarginine). By complexing with poly(L -glutamic acid) or with polyacrylate, which is also a strong α-helix-forming agent for poly(L -arginine), poly(L -homoarginine) only partially formed the α-helical conformation. Monovalent anions (OH?, Cl?, F?, and H2PO) did not change poly(L -homoarginine) to the α-helix, and in the range of pH 2–11, the polypeptide remained in an unordered conformation. In sodium dodecylsulfate, poly(L -homoarginine) exhibited the remarkably enlarged CD spectrum of an extended conformation, while poly(L -arginine) forms the α-helix by interacting with the agent. Thus poly(L -homoarginine), compared with poly(L -arginine), has a much lower ability to form the α-helical conformation by interacting with anions. The stronger hydrophobicity of homoarginine residue in comparison with the arginine residue would provide unfavorable conditions to maintain the α-helical conformation.  相似文献   

10.
By use of a polycondensation procedure free of racemization, stereoregular polymethionines have been synthesized from C-activated D -methionyl-L -methionine and L -methionyl-D -methionyl-L -methionine. The poly(D -methionyl-L -methionine) and poly(L -methionyl-D -methionyl-L -methionine) so prepared are soluble in chloroform and can be purified through dissolution in this solvent and precipitation by ligroin. Poly(D -Met-L -Met)which is obtained in a 25% yield, is about 5000 in average molecular weight. It has no discernible optical activity when examined between 400 and 600 nm in a trifluoroacetic acid solution. Poly(L -Met-D -Met-L -Met) (40% yield, M. W. = 10,000) is an optically active polymer. [α]43624 ≈ + 170° for a chloroformic solution (c = 0.2 CHCl3).  相似文献   

11.
K Itoh  B M Foxman  G D Fasman 《Biopolymers》1976,15(3):419-455
β-Helical poly(L -glutamic acid) in a gel state was found to be easily converted to the antiparallel β form by heating. Two β forms were obtained, depending on the temperature of heating. Temperatures between 40° and 85°C produced a β form with a spacing between pleated sheets (d001) of 9.03 Å, termed β1. If the heating was carried out at temperatures higher than 85°C, the β1 form underwent another conformational transition reducing the d001 value from 9.03 to 7.83 Å (termed β2) without any prominent change in the fiber repeat distance (i.e., the polypeptide backbone conformation). The time course of these two transitions was followed by measuring the infrared spectra of the samples, and it was concluded that the α → β1 transition in its initial stage obeys a pseudo-first order rate process with activation enthalpy and entropy of 54 kcal/mol and 92 eu, respectively. On the other hand, the typical sigmoidal conversion curves observed for the transition between the two types of β forms (β1 → β2) indicate that this transition proceeds via a socalled “nucleation and growth” process. The kinetic theory of phase transitions developed by Avrami can be applied with success to explain this transition. The infrared spectra, in the region from 1800 to 200 cm?1, were measured for these two β forms and the results showed that the conformation of the side chains and the mode of the hydrogen bonding between the side-chain carboxyl groups undergo appreciable change during the transition. The heat-induced conformational transition of poly(L -Glu78 L -Val22) was also studied. The copolymer was transformed from the α-helical conformation directly to the β2 form. The reason for this was thought to be due to the fact that the L -valine residues and the L -glutamyl residues near the L -valine residues have a strong tendency to take the more compact β2 form.  相似文献   

12.
Molecular motion in solid poly(L -alanine), Poly(L -leucine), poly(L -valine), and polyglycine has been investigated through measurement of the portion spin-lattice relaxation time at 30 and 60 MHz between 110 and 350°K. Rapid random reoriention of sied-chain methyl groups provides the dominent source of relaxation in the first three; activation energies are 10.5 ± 1 1, 8.5 ± 1 kJ/mol, respectively, significantly lower than in the monomeric crystals. Relaxation times in poyglycine are two orders of magnitude longer than in the monomeric crystals. Relaxation times in polyglycine, significantly lower than in the monomeric crystals. Relaxation times in polyglycine are two orders of magnitude longer and are attributed mainly to segmental motions of the polymer chains. Evidence of nonexponential recovery of nuclear magnetization was encountered in the first three homopolyamino acids but not in polylycine, and was attributed to the correlated time to characterize these motions gave quite good agreement with the data; some improvement was obtained for two polymers using a Cole-Davidson distribution of correlation times. For biopolymers using a Cole-Davidson distribution of correlation times. For biopolymers generally it is concluded that rapid methyl group reorientation is a common dynamical feature and an important source of nuclear magnetic relaxation.  相似文献   

13.
The fixation of dien-Pt on poly(I)·poly(C) leads to only minor changes in the uv and CD spectra at ambient temperature, showing that there is little perturbation of the secondary structure in the rb range studied (up to 0.30). However, the melting profiles show two steps. The Tm for strand separation increases linearly from 61°C (rb = 0) to 80°C (rb = 0.18), after which it declines on further increasing the rb. The second melting step is not complete at 100°C, and the magnitude of the absorbance change in this second step also appears to be at a maximum at rb = 0.18. Although dien-Pt can only coordinate to one base, the nmr spectra at 80°C also show a second type of interaction with the adjacent bases, which is only destroyed in the presence of a strong denaturing agent, 5M guanidinium hydrochloride. From these results and the spectrophotometric data, we observe that dien-Pt forms a triple sandwich by hydrogen bonding of the platinum amino groups to the adjacent hypoxanthine bases (N7). The presence of these hydrogen bonds accounts for the increased stability (maximal at one Pt to three hypoxanthine bases) and their rupture is seen in the second melting step. No interaction has been observed with poly(C) strand. Reaction of dien-Pt with poly(I) shows the formation of the same triple sandwich structure in the nmr spectra.  相似文献   

14.
Sequential polypeptides (L -Arg-X-Gly)n were prepared as synthetic models of arginine-rich histones to study their structure and their stereospecific interactions with DNA. In our previous work the conformational characteristics of poly(L -Arg-L -Ala-Gly), poly(L -Arg-L -Val-Gly), and poly(L -Arg-L -Leu-Gly) have already been analyzed. To obtain further insight into the influence of the X residue side chain on the conformation of the (L -Arg-X-Gly)n polytripeptides, we now report their synthesis and cd properties when X represents the amino acid residues Ile, Nva, and Nle. The pentachlorophenyl active esters of the appropriate tripeptides were used to perform the polymerization, and the toluene-4-sulfonyl group was used to protect the arginine guanido group. CD spectroscopy showed that, in 100% trifluoroethanol, the degree of helical conformation increased in the order Ile → Nle → Nva. An equilibrium between β-turn, α-helix, and random-coil conformers occurred in 100% hexafluoroisopropyl alcohol, while a rise in the temperature or the addition of water favored the α-helix, the highest percentage of which was observed in a mixture of hexafluoroisopropyl alcohol: water (20 : 80) and in the order Ile → Nle → Nva. In aqueous solutions (at pH 7 and 12) the polymers behaved as a random coil, but they were forced to a less aperiodic structure, over a range of ionic strengths (0–0.5M NaF). A rise in temperature of up to 70°C in 100% trifluoroethanol resulted in a decrease of the α-helix percentage of the polymers, while in aqueous solutions the aperiodic structure decreased with increasing temperature. This study proved the importance of the nature of the X residue (length, Cβ branching) in relation to the structural order of the sequential polypeptides. We concluded that the polymers prepared are suitable models for arginine-rich histones.  相似文献   

15.
Infrared spectroscopy, X-ray diffraction, and nuclear magnetic resonance spectroscopy have been used in investigating the conformation of two stereoregular polymethionines, poly(D -methionyl-L -methionine) and poly(L -methionyl-D -methionyl-L -methionine). When dissolved in a helicogenic solvent, such as chloroform and hexafluoroisopropanol, the polytripeptide is in an α-helical conformation. A helix-to-coil transition can then be induced by addition of trifluoroacetic acid. On the other hand, it appears that the most stable conformation of poly(D -Met-L -Met) is a β antiparallel folded structure in which the linear polypeptide segments are near to the planar extension. This structure has been evidenced through X-ray examination of oriented films, casted from solutions in chloroform. It has also been identified in solution in the same solvent, by use of infrared spectroscopy and by measuring the δHα chemical shift which characterizes the Hα proton in the peptide units. This δHα value is found equal to 5.4 ppm and differs significantly from those which are usually attributed to the α-helical conformation (δHα = 4.2 ppm) and to the random coil (δHα = 4.6 ppm). The β folded conformation of the poly(D -Met-L -Met) appears to be comparatively less stable than the α-helical one for the poly(L -Met) macromolecular stereoisomer since hexafluoroisopropanol is a helicogenic solvent for this last solute and a destabilizing one for the poly(D -Met-L -Met) β folded conformer. X-ray examinations carried out with stretched films, casted from a solution of poly(D -Met-L -Met) in chloroform, result in several data concering the cross β structure of this stereoregular polypeptide in the solid state.  相似文献   

16.
Normal-mode calculation for methylated Z-DNA poly(dG-m5dC).(dG-m5dC)   总被引:1,自引:0,他引:1  
X M Hua  E W Prohofsky 《Biopolymers》1988,27(4):645-655
Normal modes of methylated Z-DNA poly(dG-m5dC) · (dG-m5dC) are computed by helix-lattice dynamics. Good agreement with Raman spectral data is obtained. We discuss improvements in the formulation of the problem that allow us to greatly reduce the size of the matrix used. This leads to greatly reduced calculation times. The improvements come from using knowledge of the C2 and time-reversal symmetries.  相似文献   

17.
The far infrared spectra of poly(L -proline) I (190–35 cm?1) and II (400–35 cm?1) were obtained in the solid state at both 300° and 110°K. A significant difference in the region below 100 cm?1 was observed. A very intense band located at 60 cm?1 in the infrared spectrum of form II has no counterpart in form I. This indicates the sensitivity of low-frequency vibrations to the difference in conformation assumed by both forms in the solid state. Additional bands observed in this study are correlated with ir and Raman data previously reported and tentative assignments are made using the results of normal mode calculations (in the single-chain approximation) which have been reported.  相似文献   

18.
C. P. Beetz  G. Ascarelli 《Biopolymers》1982,21(8):1569-1586
We have measured the ir absorption of 5′CMP, 5′IMP, and poly(I)·poly(C) from ~25 to ~500 cm?1. From a comparison of the data with the previously measured absorption of the corresponding nucleosides and bases we can identify several “lines” associated with the deformation of the ribose ring. Out-of-plane deformation of the bases contributes strongly to vibrations near 200 cm?1. The same ribose vibrations observed in the nucleotides are found in poly(I)·poly(C). They sharpen with increasing water absorption. A study of the spectra of poly(I)·poly(C) as a function of the adsorbed water indicates that water does not contribute in a purely additive fashion to the polynucleotide spectrum but depends on the conformation of the helix. However, the only spectral feature that shifts drastically with conformation is near 45 cm?1. Measurements at cryogenic temperatures indicate some sharpening of the spectrum of poly(I)·poly(C). Instead, no sharpening is observed in the spectrum of the nucleotides. Shear degradation of poly(I)·poly(C) produces significant spectral changes in the 200-cm?1 region and sharpening of the features assigned to the low-frequency ribose-ring vibrations.  相似文献   

19.
The circular dichroism of Ac-(Ala)x-OMe and H-Lys-(Lys)x-OH with x = 1, 2, 3, and 4 has been measured in aqueous solutions. The oligomers with x = 4 show similar circular dichroism spectra in water when the lysyl amino groups are protonated, and they respond in similar fashion to heating and to sodium perchlorate. Both oligomers at 15°C exhibit a positive circular dichroism band at 217–218 nm, which is eliminated by the isothermal addition of 4 M sodium perchlorate or by heating. The positive circular dichroism of the lysine oligomer is also eliminated when the pH is elevated to deprotonate the amino groups. Positive circular dichroism is still observed for Ac-(Ala)4-OMe at elevated pH. Circular dichroism spectra have been estimated for poly(L -alanine) and poly(L -lysine) as statistical coils under the above conditions, based on the trends established with the oligomers. Poly(L -lysine) and poly(L -alanine) are predicted to exhibit similar circular dichroism behavior in aqueous solution so long as the lysyl amino groups are protonated. The circular dichroism of the statistical coil of poly(L -lysine), but not poly(L -alanine), is predicted to change when the pH is elevated sufficiently to deprotonate the lysyl amino groups. These results suggest that the unionized lysyl side chains participate in interactions that are not available to poly(L -alanine). Hydrophobic interactions may occur between the unionized lysyl side chains. Protonation of the lysyl amino groups is proposed to disrupt these interactions, causing poly(L -alanine) and protonated poly(L -lysine) to have similar circular dichroism properties.  相似文献   

20.
l-Histidine as pseudobiospecific ligand was immobilized onto poly(ethylene vinyl alcohol) hollow-fiber membranes to obtain an affinity support for immunoglobulin G (IgG) purification. The interaction of human IgG with the affinity membranes was studied by chromatography and equilibrium binding analysis. Adsorption was possible over a broad pH range and was found to depend strongly on the nature of the buffer ions rather than on ionic strength. With zwitterionic buffers like morpholinopropanesulfonic acid (Mops) and hydroxyethylpiperazineethanesulfonic acid (Hepes), much higher adsorption capacities were obtained than with other buffers like Tris-HCl and phosphate buffers. An inhibition analysis revealed that non-zwitterionic buffers competitively inhibit IgG binding, whereas Mops and Hepes in their zwitterionic form do not. By choosing the appropriate buffer system, it was possible to adsorb specifically different IgG subsets. The IgG molecules were found to adsorb on membrane immobilized histidine via their Fab part. Determination of dissociation constants at different temperatures allowed calculation of thermodynamic adsorption parameters. Decrease in KD with increasing temperature and a positive entropy value between 20 and 35°C (in Mops buffer) indicated that adsorption is partially governed by hydrophobic forces in that temperature range, whereas at lower temperatures, electrostatic forces are more important for adsorption.  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