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.
The visible and near-uv absorption and circular dichroic spectra were determined for spinach and poplar plastocyanin under a variety of conditions. The visible spectra showed that the copper center was invariant to changes in species, chemical modification with ethylenediamine, and addition of high concentrations of salt [2.7 M (NH4)2SO4]. In contrast, the near-uv spectra were sensitive to these conditions. Reduction of plastocyanin also altered its near-uv absorption and circular dichroic spectra. It is unlikely that these spectral changes were due to charge transfer bands since the near-uv CD spectrum of apo-plastocyanin was almost identical to that of reduced plastocyanin. There were no corresponding changes in the far-uv spectra which monitor protein secondary structure. The most likely explanation is that the protein has a flexible tertiary conformation. Conformational changes may be important in regulating electron transport. If plastocyanin is a mobile electron carrier, differential binding of the oxidized and reduced forms of plastocyanin to its reaction partners cytochrome f and P700 could facilitate electron transport.  相似文献   

2.
The thermal stability of plastocyanin (PC) was determined as a function of oxidation state of the copper center and the presence of oxidants, reductants, oxygen, and EDTA. It was found that the copper center and its ligands play a crucial role in maintaining the stability of PC. Thermal denaturation was monitored by using far-uv circular dichroism (CD) spectra to monitor changes in secondary structure, the near-uv CD ellipticity at 280 nm to monitor changes in tertiary structure, and the absorbance at 597 nm and the 255-nm CD transition to monitor changes in the copper center. Reduced PC (Tm = 71 degrees C) was found to be more stable than the oxidized form (Tm = 61 degrees C). The Tm was increased by addition of reductants, removal of oxygen, or addition of EDTA. Two distinct denatured forms (designated D1 and D2) were separated by anion exchange fast protein liquid chromatography. Neither form contained a native copper center. Form D2 retained the characteristic 280-nm CD band but showed an altered far-uv CD spectrum. Its formation was inhibited by the addition of reductants or the removal of oxygen. It could be refolded to form native, Cu-PC upon incubation with copper plus a reductant such as dithionite. These results suggest that its formation involves the reversible oxidation of a group on the PC molecule, possibly a ligand to the copper such as Cys 84 or Met 92. Form D1 occurred in the presence of ferricyanide or at high temperatures in the presence of oxygen. EDTA inhibited its formation. Form D1 lost the 280-nm CD transition and its far-uv CD spectrum was altered. No renaturation was observed suggesting that Form D1 is the product of an irreversible oxidation step possibly involving a histidine ligand to the copper. Forms D1 and D2 are not interconvertible and represent the endpoints of two different denaturation pathways.  相似文献   

3.
The Fourier-transform infrared spectroscopy (FTIR) technique with a diamond anvil cell has been applied for examination of the pressure-induced changes occurring in the secondary structure of the alpha-lactalbumin. This is the first high-pressure FTIR study of a calcium-binding protein which simultaneously takes into account spectral changes in both the calcium-ion-binding carboxyl groups' band and the amide I/I' vibrational band. Spectral behavior of three kinds of the protein: the undeuterated holoform, the fully deuterated holoform, and the undeuterated apoform was compared in the pressure range from 0.1 MPa up to 740 MPa. We found that the binding of calcium remarkably stabilizes the alpha-lactalbumin against pressure as it is followed approximately by a 200-MPa increase of the value of pressure at which denaturation occurs. A quantitative analysis of the band of antisymmetrical stretching vibrations of the calcium-binding carboxyl groups revealed that the pressure-induced changes in the calcium-binding loop occur in two stages. Binding of the calcium ion seemingly increases the pressure-stability of the calcium-binding loop to a higher degree than the pressure-stability of the secondary structure of the alpha-lactalbumin. We have also discussed in detail the complex pressure-enhanced H/D exchange in the alpha-lactalbumin. Finally, we have proposed a new assignment of major peaks in the helical region of the amide I/I' spectral band of the partially deuterated alpha-lactalbumin.  相似文献   

4.
The second derivative FTIR study of heat-induced and pressure-assisted cold-induced changes in the secondary structure of bovine alpha-lactalbumin was carried out for native holoprotein and calcium ion depleted apoprotein. The secondary structure and compactness of alpha-lactalbumin were examined in a temperature range from 20 to 80 degrees C during the heat treatment and 20 to -15 degrees C during the pressure-assisted cold treatment. This was the first FTIR study on the pressure-assisted cold denaturation of a protein. Because protein solutions had close to neutral pD and low ionic strength, the apoprotein remained in the molten globule state and the holoform maintained its native tertiary structure. In order to distinguish between unfolding-related and partially deuterated exchange-related spectral changes, we examined both the fully deuterated holoform and the partially deuterated holoform. The quantitative analysis of the spectral changes in the amide I/I' vibrational band revealed that the 3(10) helices were more prone to thermal unfolding than the alpha helices. We observed that the protein's compactness and secondary structure were both considerably stabilized against an increase and decrease in temperature by the presence of a calcium ion. Under the conditions of this study, only the apoprotein was susceptible to the cold denaturation. In contrast to this, an unexpected linear increase of the alpha-helical content was observed upon the cooling of the holoprotein under high pressure. The results were discussed in reference to the existing crystallographic data for crystals of human alpha-lactalbumin grown at two different temperatures.  相似文献   

5.
Infrared spectroscopy (IR) and differential scanning calorimetry (DSC) were used to study the biophysical properties of the PKCepsilon-C2 domain, a C2 domain that possess special characteristics as it binds to acidic phospholipids in a Ca2+-independent manner and no structural information about it is available to date. When the secondary structure was determined by IR spectroscopy in H2O and D2O buffers, beta sheet was seen to be the major structural component. Spectroscopic studies of the thermal denaturation in D2O showed a broadening in the amide I' band starting at 45 degrees C. Curve fitting analysis of the spectra demonstrated that two components appear upon thermal denaturation, one at 1623 cm(-1) which was assigned to aggregation and a second one at 1645 cm(-1), which was assigned to unordered or open loop structures. A lipid binding assay has demonstrated that PKCepsilon-C2 domain has preferential affinity for PIP2 although it exhibits maximal binding activity for phosphatidic acid when 100 mol% of this negatively charged phospholipid was used. Thus, phosphatidic acid containing vesicles were used to characterize the effect of lipid binding on the secondary structure and thermal stability. These experiments showed that the secondary structure did not change upon lipid binding and the thermal stability was very high with no significant changes occurring in the secondary structure after heating. DSC experiments demonstrated that when the C2-protein was scanned alone, it showed a Tm of 49 degrees C and a calorimetric denaturation enthalpy of 144.318 kJ x mol(-1). However, when phoshatidic acid vesicles were included in the mixture, the transition disappeared and further IR experiments demonstrated that the protein structure was not modified under these conditions.  相似文献   

6.
Among the various proposals that have been made in attempting to explain the ability of thermophiles to reproduce at high temperatures, there is no doubt that obligate and extreme thermophiles synthesize proteins (and other molecules) that have sufficient intrinsic molecular stability to withstand increased thermal stress. In contrast, the glyceraldehyde-3-phosphate dehydrogenase from the facultative thermophile Bacillus coagulans KU has been shown to be quite thermolabile in vitro. Thermal inactivation is not due to loss of bound NAD+. It has also been shown that the enzymatic activity can be thermostabilized in vitro by increased ionic strength. As previously reported [J. W. Crabb, A. L. Murdock, and R. E. Amelunxen (1975) Biochem. Biophys. Res. Commun. 62, 627; (1977) Biochemistry 16, 4840], the enzyme loses 94-97% of enzymatic activity after heat treatment at 55 degrees C for 5 min in 0.05 M sodium phosphate buffer (pH 7.1); however, by increasing the ionic strength to 1.8, complete protection was conferred at this temperature. Gel-filtration chromatography has been used to study the initial dissociation and subsequent aggregation of the glyceraldehyde-3-phosphate dehydrogenase after thermal inactivation. Aggregation occurs when the enzyme is heated at 50 degrees or 55 degrees C. Loss of enzymatic activity is correlated with changes in the tertiary structure as measured by the near-uv CD spectrum of the enzyme following heat inactivation, with essential disappearance of the peaks at 263 and 296 nm, and a blue shift of the far-uv spectrum, which is a measure of secondary structure. Estimation of secondary structure of the unheated protein from the far-uv CD data showed the enzyme contains approximately 26% alpha-helix, approximately 21% beta-structure, and approximately 53% disordered structure. Heat treatment at various temperatures resulted in only slight changes of the estimated secondary structure. Increased ionic strength prevents thermal alteration of the CD spectrum in both near- and far-uv regions. The data support the previous proposal that thermolabile enzymes such as the glyceraldehyde-3-phosphate dehydrogenase from the facultative thermophile B. coagulans are thermostabilized in vivo mainly by the intracellular charged macromolecular environment.  相似文献   

7.
Fourier transform infrared spectroscopy (FTIR) was used to investigate the secondary structure of 5'-nucleotidase from bull seminal plasma (BSP). Spectra of protein in both D2O and H2O were analyzed by deconvolution and second derivative methods in order to observe the overlapping components of the amide I band. The protein, which is made up of two apparently identical subunits and which contains two zinc atoms, was studied in its native form, in the presence of dithiotreitol (DTT) and after removal of the two zinc atoms by means of nitrilotriacetic acid (NTA). Deconvolved and second derivative spectra of amide I band showed that the native protein contains mostly beta-sheet structure with a minor content of alpha-helix. The quantitative analysis of the amide I components was performed by a curve-fitting procedure which revealed 54% beta-sheet, 18% alpha-helix, 22% beta-turns and 6% unordered structure. The second derivative and deconvolved spectra of amide I band showed that no remarkable changes in the secondary structure of 5'-nucleotidase were induced by either DTT or NTA. These results were confirmed by the curve-fitting analysis where little or no changes occurred in the relative content of amide I components when the protein was treated with DTT or with NTA. Major changes, however, were observed in the thermal denaturation behavior of the protein. The native protein showed denaturation at temperatures between 70 and 75 degrees C, while the maximum of denaturation was observed between 65 and 70 degrees C and between 55 and 60 degrees C in the presence of NTA and DTT, respectively. The results obtained indicate that the two separate subunits of the protein have essentially the same secondary structure as that of the native enzyme.  相似文献   

8.
The IR absorption frequencies as derived from second derivatives of the Fourier transform IR spectra of the amide I' bands of globular proteins in D2O are compared to those obtained from band fitting of the vibrational circular dichroism (VCD) spectra. The two sets of frequencies are in very good agreement, yielding consistent ranges where amide I' VCD and IR features occur. Use of VCD to complement the IR allows one to add sign information to the frequency information so that features occurring in the overlapping frequency ranges that might arise from different secondary structures can be better discriminated. From this comparison, it is clear that correlation just of the frequency of a given IR transition to secondary structure can lead to a nonunique solution. Different sign patterns were identified for correlated groups of globular proteins in restricted frequency ranges that have been previously assigned to defined secondary structural elements. Hence, different secondary structural elements must contribute band components to a given frequency range.  相似文献   

9.
Centrin is a low molecular mass (20 kDa) protein that belongs to the EF-hand superfamily of calcium-binding proteins. Local and overall changes were investigated for interactions between cations and Chlamydomonas centrin using Fourier transform infrared (FT-IR) and circular dichroic (CD) spectroscopies. FT-IR spectral features studied included the amide I' band and the side-chain absorbances for aspartate residues located almost exclusively at the calcium-binding sites in the spectral region of 1700-1500 cm(-1). The amide I' band is exquisitely sensitive to changes in protein secondary structure and is observed to shift from 1626.5 to 1642.7 cm(-1) in the presence and absence of calcium. These spectral bands are complex and were further studied using two-dimensional Fourier transform infrared (2D-FT-IR) correlation along with curve-fitting routines. Using these methods the secondary structure contributions were determined for holocentrin and apocentrin. The alpha-helical content in centrin was determined to be 60%-53% in the presence and absence of cations, respectively. Furthermore, the beta-strand content was determined to be 12%-36%, while the random coil component remained almost constant at 7%-13.5% in the presence and absence of cations, respectively. Changes in the side-chain band are mostly due to the monodentate coordination of aspartate to the cation. A shift of approximately 4 cm(-1) (for the COO- antisymmetric stretch in Asp) from 1565 to 1569 cm(-1) is observed for apocentrin and holocentrin, respectively. Thermal dependence revealed reversible conformational transition temperatures for apocentrin at 37 degrees C and holocentrin at 45 degrees C, suggesting greater stability for holocentrin.  相似文献   

10.
Despite highly conserved active-site structures, members of the plant peroxidase superfamily exhibit a wide range of pH optima. Horseradish peroxidase isozyme C (HRPC) is an ideal peroxidase to investigate the structural determinants of pH stability and activity in superfamily members. Conflicting reports exist on the low-pH stability of HRPC and consequently the pKa of the catalytic distal histidine, which is neutral in active peroxidases. Towards resolving such discrepancies, acid-induced changes in HRPC from two popular commercial suppliers were systematically analyzed. Specifically, FTIR v(CO) and Soret-CD spectra of HRPC-CO and Soret absorption of ferric HRPC were recorded to probe time-dependent heme-pocket changes at pH 3.0 in phosphate, citrate and formate buffers, while the FTIR amide I' and far-UV CD spectra were examined to probe changes in secondary structure. Both HRPC-CO samples exhibited identical pH 7.0 v(CO) bands at 1934 and 1905 cm-1. In the pH 3.0 spectrum of sample A, the 1934 cm-1 band was dominant while a broad 1969 cm-1 band appeared in sample B. The intensity of this band, which is assigned to solvent-exposed heme, was greater in citrate than phosphate buffer, but in formate the 1934 cm-1 band remained dominant. Other spectral changes mirrored the v(CO) trends. No time- or buffer-anion-dependent conformation changes were detected in 1 mM CaCl2, revealing that buffer-anion-dependent leaching of stabilizing Ca2+ from HRPC occurs at pH 3.0. Since the N-glycans present in HRPC are of the flexible protein-surface-shielding type, the variation in low-pH conformational stability of the HRPC samples could be attributed to heterogeneous glycosylation, which was detected by SDS-PAGE. It is further proposed that glycosylation patterns may affect the low-pH stability of class II and III plant peroxidases.  相似文献   

11.
Detailed differential scanning calorimetry (DSC), steady-state tryptophan fluorescence and far-UV and visible CD studies, together with enzymatic assays, were carried out to monitor the thermal denaturation of horseradish peroxidase isoenzyme c (HRPc) at pH 3.0. The spectral parameters were complementary to the highly sensitive but integral method of DSC. Thus, changes in far-UV CD corresponded to changes in the overall secondary structure of the enzyme, while that in the Soret region, as well as changes in intrinsic tryptophan fluorescence emission, corresponded to changes in the tertiary structure of the enzyme. The results, supported by data about changes in enzymatic activity with temperature, show that thermally induced transitions for peroxidase are irreversible and strongly dependent upon the scan rate, suggesting that denaturation is under kinetic control. It is shown that the process of HRPc denaturation can be interpreted with sufficient accuracy in terms of the simple kinetic scheme N -->k D where k is a first-order kinetic constant that changes with temperature, as given by the Arrhenius equation; N is the native state, and D is the denatured state. On the basis of this model, the parameters of the Arrhenius equation were calculated.  相似文献   

12.
The uncoupling protein of brown-adipose-tissue mitochondria has been purified in the form of mixed micelles with lipid and reduced Triton X-100. This surfactant has the advantage over conventional Triton X-100, that it does not interfere with amide bands in infrared spectra. The structure of the uncoupling protein in micellar form has been examined by Fourier-transform infrared spectroscopy (FTIR). In order to decompose the amide I contour into its components, band-narrowing (Fourier derivation and deconvolution) and band-decomposition techniques have been used. Combining data from spectra taken in H2O and 2H2O media, the following percentage distribution of secondary structure patterns has been obtained: 50% alpha-helix, 28-30% beta-structure; 13-15% beta-turns and 7% unordered. Thermal denaturation of the uncoupling protein has also been monitored by FTIR. In accordance with previous observations of different proteins, thermal denaturation is marked by a shift in the amide I maximum and the appearance of two new peaks in 2H2O, at around 1620 cm-1 and 1685 cm-1. Denaturation occurs in the 40-50 degrees C temperature range, in agreement with studies of GDP-binding capacity. Cooling down the thermally denatured protein produces a new change in its secondary structure; however, the original conformation is not restored. The uncoupling protein possesses a nucleotide-binding site. On addition of GDP, small changes in protein conformation occur, attributable to changes in tertiary structure. However, no detectable effects are seen in the presence or absence of the other physiological regulators, the free fatty acids. The uncoupling protein shares important similarities in its primary structure with other anion carriers of the mitochondrial membrane; one of these, the adenine-nucleotide translocator, has been used in a comparative study, applying the same FTIR techniques described above for the uncoupling protein. Both proteins have a similar proportion of alpha-helix, probably corresponding to the segments spanning the membrane, but the conformation of the polar domains appears to differ.  相似文献   

13.
Infrared spectra of 2.5 mM solutions of beta-lactoglobulin B were recorded as a function of pH (from pH 2 to pH 13) and as a function of temperature (from -100 degrees C to +90 degrees C). An analysis of the pH- and temperature-induced changes in the secondary structure was performed based on changes in the conformation-sensitive amide I bands of beta-lactoglobulin. Whereas the total amount of beta-structure remains constant (56-59%) between pH 2 and pH 10, the proportions of the various beta-components do change. In particular, the dimerization of the monomeric protein, induced by raising the pH from 2 to 3 , leads to an increase in the intensity of the 1636 cm-1 band (associated with antiparallel beta-sheet), at the expense of the 1626 cm-1 band (associated with exposed beta-strands). Both the thermal and alkaline denaturation of beta-lactoglobulin occur in two distinct stages. Although the spectra (i.e., the structures) after complete thermal or alkaline denaturation are clearly different, the spectrum of the protein after the first stage of thermal denaturation (at about 60 degrees C) is the same as that after the first stage of alkaline denaturation (at pH 11), suggesting a common denaturation intermediate, which probably represents a crossover point in a complex potential hypersurface.  相似文献   

14.
Panick G  Winter R 《Biochemistry》2000,39(7):1862-1869
In this paper, we illustrate the use of high-pressure Fourier transform infrared (FT-IR) spectroscopy to study the reversible presssure-induced unfolding and refolding of ribonuclease A (RNase A) and compare it with the results obtained for the temperature-induced transition. FT-IR spectroscopy monitors changes in the secondary structural properties (amide I' band) or tertiary contacts (tyrosine band) of the protein upon pressurization or depressurization. Analysis of the amide I' spectral components reveals that the pressure-induced denaturation process sets in at 5. 5 kbar at 20 degrees C and pH 2.5. It is accompanied by an increase in disordered structures while the content of beta-sheets and alpha-helices drastically decreases. The denatured state above 7 kbar retains nonetheless some degree of beta-like secondary structure and the molecule cannot be described as an extended random coil. Increase of pH from 2.5 to 5.5 has no influence on the structure of the pressure-denatured state; it slightly changes the stability of the protein only. All experimental evidence indicates that the pressure-denatured states of monomeric proteins have more secondary structure than the temperature-denatured states. Different modes of denaturation, including pressure, may correlate differently with the roughness of the energy scale and slope of the folding funnel. For these reasons we have also carried out pressure-jump kinetic studies of the secondary structural evolution in the unfolding/refolding reaction of RNase A. In agreement with the theoretical model presented by Hummer et al. [(1998) Proc. Natl. Acad. Sci. U.S.A. 95, 1552-1555], the experimental data show that pressure slows down folding and unfolding kinetics (here 1-2 orders of magnitude), corresponding to an increasingly rough landscape. The kinetics remains non-two-state under pressure. Assuming a two-step folding scenario, the calculated relaxation times for unfolding of RNase A at 20 degrees C and pH 2.5 can be estimated to be tau(1) approximately 0.7 min and tau(2) approximately 17 min. The refolding process is considerably faster (tau(1) approximately 0.3 min, tau(2) approximately 4 min). Our data show that the pressure stability and pressure-induced unfolding/refolding kinetics of monomeric proteins, such as wild-type staphylococcal nuclease (WT SNase) and RNase A, may be significantly different. The differences are largely due to the four disulfide bonds in RNase A, which stabilize adjacent structures. They probably lead to the much higher denaturation pressure compared to SNase, and this might also explain why the volume change of WT SNase upon unfolding is about twice as large.  相似文献   

15.
Calbindin-D(28K) is a biologically important protein required for normal neural function and for the transport of calcium in epithelial cells of the intestine and kidney. We have used fluorescence and circular dichroism (CD) spectroscopy to characterize the effects of calcium binding on the structure and stability of calbindin. Ca(2+) titration monitored by fluorescence spectroscopy reveals the presence of two classes of calcium-binding sites with association constants approximately 10(7.5) and approximately 10(8.9)M(-1). CD spectra in the far-UV spectral range show minor changes upon Ca(2+) titration, implying that the secondary structure of calbindin-D(28K) is not greatly affected. On the basis of the CD spectra in the near-UV spectral range, we conclude that the tertiary structure is more sensitive to Ca(2+) addition. The most significant change occurs between pCa 7.0 and pCa 8.0. The variations in the protein thermostability are correlated with those in the near-UV CD spectra. The enthalpy changes upon heat denaturation of calbindin in the apo-state are characteristic of proteins containing several weakly interacting domains with similar thermodynamical properties. Thus, calcium binding by calbindin-D(28K) largely affects the local structure around the aromatic residues and the thermal stability of the protein; the changes in the secondary structure are insignificant.  相似文献   

16.
Fourier transform infrared (FTIR) spectroscopy was used to investigate the structural and thermal denaturation of the C2 domain of PKC alpha (PKC-C2) and its complexes with Ca(2+) and phosphatidic acid vesicles. The amide I regions in the original spectra of PKC-C2 in the Ca(2+)-free and Ca(2+)-bound states are both consistent with a predominantly beta-sheet secondary structure below the denaturation temperatures. Spectroscopic studies of the thermal denaturation revealed that for the PKC-C2 domain alone the secondary structure abruptly changed at 50 degrees C. While in the presence of 2 and 12.5 mM Ca(2+), the thermal stability of the protein increased to 60 and 70 degrees C, respectively. Further studies using a mutant lacking two important amino acids involved in Ca(2+) binding (PKC-C2D246/248N) demonstrated that these mutations were inherently more stable to thermal denaturation than the wild-type protein. Phosphatidic acid binding to the PKC-C2 domain was characterized, and the lipid-protein binding became Ca(2+)-independent when 100 mol% phosphatidic acid vesicles were used. The mutant lacking two Ca(2+) binding sites was also able to bind to phosphatidic acid vesicles. The effect of lipid binding on secondary structure and thermal stability was also studied. Beta-sheet was the predominant structure observed in the lipid-bound state, although the percentage represented by this structure in the total area of the amide I band significantly decreased from 60% in the lipid-free state to 47% in the lipid-bound state. This decrease in the beta-sheet component of the lipid-bound complex correlates well with the significant increase observed in the 1644 cm(-1) band which can be assigned to loops and disordered structure. Thermal stability after lipid binding was very high, and no sign of thermal denaturation was observed in the presence of lipids under the conditions that were studied.  相似文献   

17.
Ye M  Zhang QL  Li H  Weng YX  Wang WC  Qiu XG 《Biophysical journal》2007,93(8):2756-2766
The infrared (IR) absorption of the amide I band for the loop structure may overlap with that of the alpha-helices, which can lead to the misassignment of the protein secondary structures. A resolution-enhanced Fourier transform infrared (FTIR) spectroscopic method and temperature-jump (T-jump) time-resolved IR absorbance difference spectra were used to identify one specific loop absorption from the helical IR absorption bands of horse heart cytochrome c in D2O at a pD around 7.0. This small loop consists of residues 70-85 with Met-80 binding to the heme Fe(III). The FTIR spectra in amide I' region indicate that the loop and the helical absorption bands overlap at 1653 cm(-1) at room temperature. Thermal titration of the amide I' intensity at 1653 cm(-1) reveals that a transition in loop structural change occurs at lower temperature (Tm=45 degrees C), well before the global unfolding of the secondary structure (Tm approximately 82 degrees C). This loop structural change is assigned as being triggered by the Met-80 deligation from the heme Fe(III). T-jump time-resolved IR absorbance difference spectra reveal that a T-jump from 25 degrees C to 35 degrees C breaks the Fe-S bond between the Met-80 and the iron reversibly, which leads to a loop (1653 cm(-1), overlap with the helical absorption) to random coil (1645 cm(-1)) transition. The observed unfolding rate constant interpreted as the intrachain diffusion rate for this 16 residue loop was approximately 3.6x10(6) s(-1).  相似文献   

18.
CD spectra of bovine pancreatic ribonuclease A (RNase A) and its subtilisin-modified form (RNase S) have been calculated, based upon high-resolution structures from x-ray diffraction. All known transitions in the peptide and side-chain groups, especially the aromatic and disulfide groups, have been included. Calculations have been performed with both the matrix method and with first-order perturbation theory. A newly developed method for treating the electrostatic interactions among transition charge densities and between static charge distributions and transition charge densities is used. The effects of local electrostatic fields upon the group transition energies are included for all transitions. Rotational strengths generated by the matrix method were combined with Gaussian band shapes to generate theoretical CD spectra. The calculated spectra reproduce the signs and approximate magnitudes of the near-uv CD bands of both RNase A and S. Agreement is most satisfactory for the negative 275 nm band, dominated by tyrosine contributions. In agreement with two previous studies by other workers, coupling between Tyr 73 and Tyr 115 is the single most important factor in this band. The positive band observed near 240 nm is dominated by disulfide contributions, according to our results. The far-uv CD spectrum is poorly reproduced by the calculations. The observed 208 nm band, characteristic of α-helices, is absent from the calculated spectrum, probably because the helices in RNase are short. A strong positive couplet centered near 190 nm is predicted but not observed. Possible reasons for these incorrect predictions of the current theoretical model in the far-uv are discussed. © 1997 John Wiley & Sons, Inc.  相似文献   

19.
Decatur SM 《Biopolymers》2000,54(3):180-185
The effect of N-acetylation on the conformation of alanine-rich helical peptides is examined using isotope-edited Fourier transform infrared (FTIR) spectroscopy. A series of peptides with sequence AA(AAKAA)(3)AAY has been prepared; each peptide incorporates four (13)C-labeled alanines. These peptides have two amide I' bands in their FTIR spectra: one corresponding to the (12)C amino acids, and one assigned to the (13)C amino acids. The intensity and frequency of the (13)C amide I' band varies systematically with the position of the labels in the sequence and the presence or absence of an N-acetyl capping group. The intensity of the (13)C amide I' band correlates with helix stability at the labeled residues as predicted by thermodynamic models of the helix-coil transition. These results suggest that FTIR spectroscopy combined with specific isotope labeling can be used to dissect the conformation of helical peptides at the residue level.  相似文献   

20.
The interactions of brain-derived neurotrophic factor (BDNF) with the extracellular domain of its receptor (trkB) are investigated by employing isotope-edited Fourier transform IR (FTIR) spectroscopy. The protein secondary structures of individual BDNF and trkB in solutions are compared with those in their complex. The temperature dependence of the secondary structures of BDNF, trkB, and their complex is also investigated. Consistent with the crystal structure, we observe by FTIR spectroscopy that BDNF in solution contains predominantly beta strands (approximately 53%) and relatively low contents of other secondary structures including beta turns (approximately 16%), disordered structures (approximately 12%), and loops (approximately 18%) and is deficient in alpha helix. We also observe that trkB in solution contains mostly beta strands (52%) and little alpha helix. Conformational changes in both BDNF and trkB are observed upon complex formation. Specifically, upon binding of BDNF, the conformational changes in trkB appear to involve mostly beta turns and disordered structures while the majority of the beta-strand conformation remains unchanged. The IR data indicate that some of the disordered structures in the loop regions are likely converted to beta strands upon complex formation. The FTIR spectral data of BDNF, trkB, and their complex indicate that more amide NH groups of trkB undergo H-D exchange within the complex than those of the ligand-free receptor and that the thermal stability of trkB is decreased slightly upon binding of BDNF. The FT-Raman spectra of BDNF, trkB, and their complex show that the six intramolecular disulfide bonds of trkB undergo significant conformational changes upon binding of BDNF as a result of changes in the tertiary structure of trkB. Taken together, the FTIR and Raman data are consistent with the loosening of the tertiary structure of trkB upon binding of BDNF, which leads to more solvent exposure of the amide NH group and decreased thermal stability of trkB. This finding reveals an intriguing structural property of the neurotrophin ligand-receptor complex that is in contrast to other ligand-receptor complexes such as a cytokine-receptor complex that usually shows protection of the amide NH group and increased thermal stability upon complex formation.  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