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.
Abstract

Acid-induced structural changes of a mouse IgG2a monoclonal antibody (MN12) as indicated by Jiskoot et al. (Eur. J. Biochem. 201, 223–232 (1991)) were studied by measuring the transient electric birefringence of MN12 in aqueous solution and in glycerol-water mixtures at different pH conditions. A multi-exponential analysis program, DISCRETE (Provencher, S.W., Biophys. J. 16, 27–41 (1976)), and a constrained inverse Laplace transform program, CONTIN (Provencher, S.W., Comp. Phys. Comm. 27, 213–227 (1982)) have been used to determine the number of exponentials needed to represent the data and their decay times.

Measurement of the time-resolved electric field induced birefringence makes it possible to study rotational processes on a timescale from several tens of nanoseconds to microseconds. This enabled us to monitor the segmental flexibility and the rotational motion of single antibody molecules as well as the occurrence of aggregates.

The results show an increase in hydrodynamic dimensions of MN12 upon lowering the pH from 6.6 to 2.7. Additionally, the original segmental flexibility, which could be monitored for the samples in glycerol-water mixtures, is altered at low pH. The results have been interpreted as swelling of MN12 followed by dimerization.

A short discussion on the use of DISCRETE and CONTIN is added.  相似文献   

2.
The pH dependence of the conformation of a mouse IgG2a, kappa monoclonal antibody (MN12) was investigated by several physical techniques, including fluorescence spectroscopy, near-ultraviolet and far-ultraviolet CD, and electric-field-induced transient birefringence measurements. The intensity of the intrinsic tryptophan fluorescence remained constant in the pH range from 3.5 to 10.0. A conformational alteration in the MN12 molecule was observed in the pH region between pH 3.5 and 2.5, as reflected by a substantial enhancement of the fluorescence quantum yield. This effect was more pronounced at high ionic strengths. The fluorescence emission was unaltered, indicating that the acid-induced conformational state is different from a completely unfolded state. This was confirmed by CD and fluorescence polarisation measurements. Iodide and acrylamide fluorescence quenching studies indicated a gradually increasing accessibility of MN12 tryptophan residues with decreasing pH. At low pH precipitation was observed in the presence of iodide. One rotational relaxation time (0.16-0.18 microseconds) was observed for MN12 by electric-field-induced transient birefringence measurements at low ionic strength. After exposure of MN12 to low pH for 1 h, the relaxation time was increased to 0.23 microseconds; a further increase to 0.30 microseconds was observed after 24 h. The combined results suggest an acid-induced expansion and enhanced flexibility of MN12, which eventually leads to irreversible aggregation.  相似文献   

3.
The dynamics of single tryptophan (W) side chain of protease subtilisin Carlsberg (SC) and myelin basic protein (MBP) were used for probing the surface of these proteins. The W side chains are exposed to the solvent, as shown by the extent of quenching of their fluorescence by KI. Time-resolved fluorescence anisotropy measurements showed that the rotational motion of W is completely unhindered in the case of SC and partially hindered in the case of MBP. The rotational correlation time (phi) associated with the fast local motion of W did not scale linearly with the bulk solvent viscosity (eta) in glycerol-water mixtures. In contrast, phi values of either W side chains in the denatured proteins or the free W scaled almost linearly with eta, as expected by the Stokes-Einstein relationship. These results were interpreted as indicating specific partitioning of water at the surface of the proteins in glycerol-water mixtures.  相似文献   

4.
A new procedure based on the statistical method of "variable selection" is used to predict the secondary structure of proteins from circular dichroism spectra. Variable selection adds the flexibility found in the Provencher and Gl?ckner method (S. W. Provencher and J. Gl?ckner, 1981, Biochemistry 20, 33-37) to the method of Hennessey and Johnson (J. P. Hennessey and W. C. Johnson, 1981, Biochemistry 20, 1085-1094). Two analytical methods are presented for choosing a solution from the series generated by the Provencher and Gl?ckner method, and this improves the technique. All three methods are compared and it is shown that both the variable selection method and the improved Provencher and Gl?ckner methods have equivalent reliability superior to the original Hennessey and Johnson method. For the new variable selection method, correlation coefficients calculated between X-ray structure and predicted secondary structures for data measured to 178 nm are: 0.97 for alpha-helix, 0.75 for beta-sheet, 0.50 for beta-turn, and 0.89 for other structures. Although the variable selection method improves the analysis of circular dichroism data truncated at 190 nm, data measured to 178 nm gives superior results. It is shown that improving the fit to the measured CD beyond the accuracy of the data can result in poorer analyses.  相似文献   

5.
The conformation of porcine serum ferric transferrin (Tf) and its stability against denaturation were studied by circular dichroism. Tf was estimated to have 19-24% alpha-helix and 50-55% beta-sheet based on the methods of Chang et al. (Chang, C.T., Wu, C.-S.C., & Yang, J.T., 1978, Anal. Biochem. 91, 13-31) and Provencher and Glöckner (Provencher, S.W. & Glöckner, J., 1981, Biochemistry 20, 33-37). Removal of the bound ferric ions (apo-Tf) did not alter the overall conformation, but there were subtle changes in local conformation based on its near-UV CD spectrum. The Tfs were stable between pH 3.5 and 11. Denaturation by guanidine hydrochloride (Gu-HCl) showed two transitions at 1.6 and 3.4 M denaturant. The process of denaturation by acid and base was reversible, whereas that by Gu-HCl was partially reversible. The irreversible thermal unfolding of Tfs began at temperatures above 60 degrees C and was not complete even at 80 degrees C. The bound irons (based on absorbance at 460 nm) were completely released at pH < 4 or in Gu-HCl solution above 1.7 M, when the protein began to unfold, but they remained intact in neutral solution even at 85 degrees C. The NH2- and COOH-terminal halves of the Tf molecule obtained by limited trypsin digestion had CD spectra similar to the spectrum of native Tf, and the COOH-terminal fragment had more stable secondary structure than the NH2-terminal fragment.  相似文献   

6.
The quantitative analysis of circular dichroic spectra of native human plasma fibronectin according to the method of Provencher and Gl?ckner [Provencher, S. W., & Gl?ckner, J. (1981) Biochemistry 20, 33-37] indicated the presence of beta-sheet (79%), beta-turn (21%), but no alpha-helix or random coil in the secondary structure. The calf alveolar heparan sulfates induced a change in the conformation of fibronectin: the magnitude of the change depended on the molecular properties of the particular heparan sulfate preparations.  相似文献   

7.
G Deléage  B Roux  C Marion 《Biochemistry》1986,25(10):2854-2858
The electric birefringence technique was used to investigate the steady-state birefringence, the orientational relaxation time, and the orientation mechanism of pig heart mitochondrial F1 adenosine-5'-triphosphatase (F1-ATPase). The electrooptical properties of this enzyme in solution were studied as functions of pH, protein concentration, and applied electric field. The F1-ATPase exhibits a surprising negative electric birefringence with a specific Kerr constant of -1.5 X 10(-3) esu cgs. The field-independent relaxation time was found to be 0.65 +/- 0.05 microseconds, corresponding to a rotational diffusion constant of 2.55 X 10(5) s-1. The overall size and shape of F1-ATPase have been calculated from both translational and rotational diffusion constants. The enzyme may be assumed to be an oblate ellipsoid of revolution with dimensions of about 170 X 170 X 70 A. The orientation mechanism of F1-ATPase was analyzed by fitting experimental birefringence rising curves with theoretical rising functions. The ratio of the permanent to induced dipole moment is found to be very high; therefore, the birefringence of F1-ATPase is due to a strong permanent dipole moment in a direction perpendicular to the long axes of the particle. These particular electric properties can be explained by the oligomeric structure of the protein and seem likely to play a role in its mechanism of functioning.  相似文献   

8.
The technique of transient electrical birefringence was used to compare some of the electric and structural dynamic properties of myosin subfragment 1 (S1(elc, rlc)), which has both the essential and regulatory light chains bound, to S1(elc), which has only an essential light chain. The rates of rotational Brownian motion indicate that S1(elc, rlc) is larger, as expected. The permanent electric dipole moment of S1(elc, rlc) is also larger, indicating that the regulatory light chain portion of S1(elc, rlc) has a dipole moment and that it is aligned head-to-tail with the dipole moment of the S1(elc) portion. The permanent electric dipoles decrease with increasing ionic strength, apparently because of ion binding to surface charges. Both S1(elc, rlc) and S1(elc) have intrinsic segmental flexibility, as detected by the ability to selectively align segments with a brief weak electric field. However, unlike S1(elc), which can be structurally distorted by the action of a brief strong electric field, S1(elc, rlc) is stiffer and cannot be distorted by fields as high as 7800 V/cm applied to its approximately 8000 D permanent electric dipole moment. The S1 . MgADP . Pi analog S1 . MgADP . Vi is smaller than S1 . MgADP, for both S1(elc, rlc) and S1(elc). Interestingly, the smaller, stiffer S1(elc, rlc) . MgADP . Vi complex retains intrinsic segmental flexibility. These results are discussed within a framework of current hypotheses of force-producing mechanisms that involve S1 segmental motion and/or the loss of cross-bridge flexibility during force production.  相似文献   

9.
Myosin subfragment 1 has tertiary structural domains   总被引:4,自引:0,他引:4  
S Highsmith  D Eden 《Biochemistry》1986,25(8):2237-2242
Transient electrical birefringence measurements were made on skeletal muscle myosin subfragment 1 (S1) at 3.7 degrees C in 10 mM tris(hydroxymethyl)aminomethane-acetate and 0.10 mM MgCl2, pH 7.0. The specific birefringence for 4.5 microM S1 was determined from steady-state measurements to be (8.1 +/- 0.3) X 10(-7) (cm/statvolt)2. For electric fields in the range of 2.47-24.7 statvolts/cm, the alignment was due to a large permanent dipole moment for S1, estimated to be 8500 +/- 2000 D. The duration and the strength of the transient electric field was varied, and the temporal response of the decay of the birefringence signal was analyzed. The rate of rotational motion after the field was removed increased with increasing field strength for short (0.35-microseconds) pulses and decreased with increasing pulse lengths for all field strengths. The rate of decay from a steady-state birefringence signal was independent of field strength. A model of S1 structure is proposed, which is consistent with these data and most other data on S1 structure. In this model, S1 is composed of two tertiary structural domains that are connected by a flexible linkage with a substantial restoring force. The electric dipole moments on the two domains are arranged head to tail. The segmental movement of the domains is restricted to certain directions. The average conformation of the molecule is elongated, but it can be made more compact by the torque exerted by an electric field. The structural changes depend on the strength and duration of the pulse.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

10.
Epidemiological evidence suggests that moderate consumption of red wine reduces the incidence of Alzheimer disease (AD). To study the protective effects of red wine, experiments recently were executed in the Tg2576 mouse model of AD. These studies showed that a commercially available grape seed polyphenolic extract, MegaNatural-AZ (MN), significantly attenuated AD-type cognitive deterioration and reduced cerebral amyloid deposition (Wang, J., Ho, L., Zhao, W., Ono, K., Rosensweig, C., Chen, L., Humala, N., Teplow, D. B., and Pasinetti, G. M. (2008) J. Neurosci. 28, 6388-6392). To elucidate the mechanistic bases for these observations, here we used CD spectroscopy, photo-induced cross-linking of unmodified proteins, thioflavin T fluorescence, size exclusion chromatography, and electron microscopy to examine the effects of MN on the assembly of the two predominant disease-related amyloid beta-protein alloforms, Abeta40 and Abeta42. We also examined the effects of MN on Abeta-induced cytotoxicity by assaying 3-[4,5-dimethylthiazol-2-yl]-2,5-diphenyltetrazolium bromide metabolism and lactate dehydrogenase activity in Abeta-treated, differentiated pheochromocytoma (PC12) cells. Initial studies revealed that MN blocked Abeta fibril formation. Subsequent evaluation of the assembly stage specificity of the effect showed that MN was able to inhibit protofibril formation, pre-protofibrillar oligomerization, and initial coil --> alpha-helix/beta-sheet secondary structure transitions. Importantly, MN had protective effects in assays of cytotoxicity in which MN was mixed with Abeta prior to peptide assembly or following assembly and just prior to peptide addition to cells. These data suggest that MN is worthy of consideration as a therapeutic agent for AD.  相似文献   

11.
The intrinsic birefringences of TMV, tropocollagen, and paramyosin were calculated from flow birefringence measurements using the theory of Peterlin and Stuart. The values are -0.029, -0.029, and -0.030, respectively. The intrinsic birefringences of TMV and tropocollagen were measured as a function of the refractive index of the solvent in glycerol-water mixtures. In both cases the values were not constant and became less negative as the refractive index increased. Theoretical calculations showed that the large solvent effect could not be caused by a hydration shell of index different from that of the bulk solvent. It is concluded that either (a) the intrinsic birefringence calculated from the Peterlin-Stuart theory is incorrect or (b) the intrinsic birefringence depends markedly on the solvent. These results are of importance to the problem of quantitative polarized light microscopy since the separation of form and intrinsic birefringence contributions is based on the assumption that intrinsic birefringence is independent of solvent.  相似文献   

12.
Lu Y  Weers B  Stellwagen NC 《Biopolymers》2001,61(4):261-275
DNA restriction fragments ranging from 79 to 789 base pairs in length have been characterized by transient electric birefringence (TEB) measurements at various temperatures between 4 and 43 degrees C. The DNA fragments do not contain runs of four or more adenine residues in a row and migrate with normal electrophoretic mobilities in polyacrylamide gels, indicating that they are not intrinsically curved or bent. The low ionic strength buffers used for the measurements contained 1 mM Tris Cl, pH 8.0, EDTA, and variable concentrations of Na(+) or Mg(2+) ions. The rotational relaxation times were obtained by fitting the TEB field-free decay signals with a nonlinear least-squared fitting program; the decay of the birefringence was monoexponential for fragments < or = 241 base pair (bp) in length and multiexponential for larger fragments. The terminal relaxation times, characteristic of the end-over-end rotation of the DNA molecules, were then used to determine the persistence length (p) and hydrodynamic radius (r) of DNA as a function of temperature and ionic strength, using several different hydrodynamic models. The specific values obtained for p and r are model dependent. The wormlike chain model of P. J. Hagerman and B. H. Zimm (Biopolymers 1981, Vol. 20, pp. 1481-1502) combined with the revised Broersma equation (J. Newman et al., Journal of Mol Biol 1997, Vol. 116, pp. 593-606) appears to be the most suitable for describing the flexibility of DNA in low ionic strength solutions. The values of p and r obtained from the global least squares fitting of this equation are independent of DNA length, and the deviations of the individual values from the average are reasonably small. The consensus r value calculated for DNA in various low ionic strength solutions containing 1 mM Tris buffer is 14.7 +/- 0.4 A at 20 degrees C. The consensus p values decrease from 814 approximately 564 A in solutions containing 1 mM Tris buffer plus 0.2-1 mM NaCl and decrease still further to 440 A in solutions containing 0.2 mM Mg(2+) ions. The persistence length exhibits a shallow maximum at 20 degrees C and decreases slowly upon either increasing or decreasing the temperature, regardless of the model used to fit the data. By contrast, the consensus values of the hydrodynamic radius are independent of temperature. The calculated persistence lengths and hydrodynamic radii are compared with other data in the literature.  相似文献   

13.
Goldfarb NE  Lam MT  Bose AK  Patel AM  Duckworth AJ  Dunn BM 《Biochemistry》2005,44(48):15725-15733
Human cathepsin D (hCatD) is an aspartic peptidase with a low pH optimum. X-ray crystal structures have been solved for an active, low pH (pH 5.1) form (CatD(lo)) [Baldwin, E. T., Bhat, T. N., Gulnik, S., Hosur, M. V., Sowder, R. C., Cachau, R. E., Collins, J., Silva, A. M., and Erickson, J. W. (1993) Proc. Natl. Acad. Sci. U.S.A. 90, 6796-6800] and an inactive, high pH (pH 7.5) form (CatD(hi)) [Lee, A. Y., Gulnik, S. V., and Erickson, J. W. (1998) Nat. Struct. Biol. 5, 866-871]. It has been suggested that ionizable switches involving the carboxylate side chains of E5, E180, and D187 may mediate the reversible interconversion between CatD(hi) and CatD(lo) and that Y10 stabilizes CatD(hi) [Lee, A. Y., Gulnik, S. V., and Erickson, J. W. (1998) Nat. Struct. Biol. 5, 866-871]. To test these hypotheses, we generated single point mutants in "short" recombinant human pseudocathepsin D (srCatD), a model kinetically similar to hCatD [Beyer, B. M., and Dunn, B. M. (1996) J. Biol. Chem. 271, 15590-15596]. E180Q, Y10F, and D187N exhibit significantly higher kcat/Km values (2-, 3-, and 6-fold, respectively) at pH 3.7 and 4.75 compared to srCatD, indicating that these residues are important in stabilizing the CatD(hi). E5Q exhibits a 2-fold lower kcat/Km compared to srCatD at both pH values, indicating the importance of E5 in stabilizing the CatD(lo). Accordingly, full time-course "pH-jump" (pH 5.5-4.75) studies of substrate hydrolysis indicate that E180Q, D187N, and Y10F have shorter kinetic lag phases that represent the change from CatD(hi) to CatD(lo) compared to srCatD and E5Q. Intrinsic tryptophan fluorescence reveals that the variants have a native-like structure over the pH range of our assays. The results indicate that E180 and D187 participate as an electrostatic switch that initiates the conformational change of CatD(lo) to CatD(hi) and Y10 stabilizes CatD(hi) by hydrogen bonding to the catalytic Asp 33. E5 appears to play a less significant role as an ionic switch that stabilizes CatD(lo).  相似文献   

14.
15.
We have developed a straightforward method to separate linear-dichroism and birefringence contributions to electric-field induced signals in a conventional birefringence setup. The method requires the measurement of electric birefringence for three different angular positions of the analyzer. It is demonstrated that the presence of linear dichroism can significantly influence the measured signals and lead to completely erroneous calculations of the birefringence signal and field-free decay times if its contribution is not taken into account. The new method is used to determine electric birefringence and linear dichroism of trimeric Photosystem 1 complexes from the cyanobacterium Synechocystis PCC 6803 in the detergents n-dodecyl-beta-D-maltoside and n-octyl-beta-D-glucoside. It is concluded that the orientation of the particles in the field is predominantly caused by a permanent electric dipole moment that is directed parallel to the symmetry axis of the particles. Comparison of the decay times obtained with dodecylmaltoside and octylglucoside supports a model in which the thickness of the disc-like complexes remains similar (7-8 nm) upon replacing dodecylmaltoside by octylglucoside, whereas the diameter increases from 14.4 +/- 0.2 to 16.6 +/- 0.2 nm because of an increased thickness of the detergent layer. This change in diameter is in good agreement with electron-microscopy results on Photosystem 2 complexes in dodecylmaltoside and octylglucoside (Dekker, J. P., E. J. Boekema, H. T. Witt, and M. Rögner. 1988. Biochim. Biophys. Acta 936:307-318). The value of approximately 16.6 nm for the diameter of Photosystem 1 trimers in dodecylmaltoside is in good agreement with recent results obtained from electron microscopy in combination with extensive image analysis (Kruip, J., E. J. Boekema, D. Bald, A. F. Boonstra, and M. Rögner. 1993. J. Biol. Chem. 268:23353-23360).  相似文献   

16.
Transient electrical birefringence characterization of heavy meromyosin   总被引:3,自引:0,他引:3  
S Highsmith  D Eden 《Biochemistry》1985,24(18):4917-4924
Heavy meromyosin (HMM) and myosin subfragment 1 (S1) were prepared from myosin by using low concentrations of alpha-chymotrypsin. The light chain distribution in HMM was identical with that of myosin, within experimental error, when analyzed on 12% polyacrylamide gels after electrophoresis. Specific birefringences and birefringence decay times were measured by transient electrical birefringence in 5 mM KCl, 5 mM tris(hydroxymethyl)aminomethane (pH 7), and 1 mM MgCl2 at 4 degrees C under gentle conditions that reduced the CaATPase activity by less than 10%. For solutions of HMM, by use of electric field pulses shorter than 0.5 microseconds, the birefringence decay signal from the S1 portions of HMM could be resolved and the rotational motions of the S1 moieties observed directly. The rotation relaxation time, adjusted to 20 degrees C, was 0.34 microseconds; this is in quantitative agreement with previous hydrodynamic results obtained by using covalently attached probes. The assignment of the fast decay time obtained with HMM to the S1 portions was confirmed by birefringence decay measurements on free S1, for which the relaxation time was 0.13 microseconds, corrected to 20 degrees C. The specific birefringences for S1 and HMM, respectively, were 0.37 X 10(-6) and 12.8 X 10(-6) (cm/statvolt)2. Thus, for much longer electric field pulses, the signal from HMM is due almost entirely to its subfragment 2 (S2) portion, and its rotational dynamics can also be monitored directly by using electrical birefringence. The decay of the signal from the S2 portion could be adequately fit without evoking bending of the S2 portion of HMM other than at its junction with S1.  相似文献   

17.
The conformational dynamics of domain III in annexin V bound to negatively charged phospholipid vesicles of 1-palmitoyl-2-oleoyl-sn-glycerophosphocholine and 1-palmitoyl-2-oleoyl-sn-glycerophosphoserine or incorporated into reverse micelles of water/sodium bis(2-ethylhexyl) sulfosuccinate in isooctane, used to mimic the phospholipid/water interface, was studied by steady-state and time-resolved fluorescence of its single tryptophan residue (W187). Upon interaction with sonicated phospholipid vesicles in the presence of calcium, or upon incorporation into reverse micelles without calcium, a progressive 12-14 nm red shift of the fluorescence emission spectrum of W187 is observed. The indole environment becomes therefore more polar than in the unbound protein. Three major lifetime populations describe the fluorescence intensity decays of W187 in both systems. A long-lived excited-state population characterizes the membrane-bound state of the protein. The existence of local conformers with different subnanosecond mobility is suggested by specific association between lifetimes and correlation times both for the protein in buffer and in interaction with the membrane surface. The interaction of the protein with the membrane surface preserves the existence of a rapid unhindered rotational motion, which is coupled with all three lifetimes. The longest lifetime is coupled to restricted motions in subnanosecond and nanosecond time scales. The overall amplitude of rotation of the indole ring is increased in the membrane-bound conformation of the protein. In reverse micelles, the local dynamics reported by W187 is also considerably increased whereas the overall folding of the protein remains unaffected. The same conformational change of domain III can therefore be provoked by different conditions: calcium binding at high concentration, mild acidic pH [Sopkova, J., Vincent, M., Takahashi, M., Lewit-Bentley, A. , and Gallay, J. (1998) Biochemistry 37, 11962-11970] and the interaction of the protein with the membrane surface. The high flexibility of domain III in the membrane-bound protein suggests that this domain may not be crucial for the interaction of the protein with the membrane, in contrast with previous models. Our data are compatible with atomic force microscopy results which suggest that domain III of annexin V does not interact strongly with the membrane surface [Reviakine, I., Bergma-Schutter, W., and Brisson, A. (1998) J. Struct. Biol. 121, 356-361].  相似文献   

18.
S Highsmith  D Eden 《Biochemistry》1987,26(10):2747-2750
The effects of limited trypsinolysis of myosin subfragment 1 (S1) on its structural dynamics were investigated by using the method of transient electric birefringence. Conversion of S1 by trypsin to produce S1 (T) did not change the specific Kerr constant [(8.1 +/- 0.3) X 10(-7) and (8.0 +/- 0.3) X 10(-7) cm2/statvolt2 for S1(T) and S1, respectively] or the degree of alignment in a weak electric field, suggesting that the size of S1 and its permanent electric dipole moment are not modified by trypsin. On the other hand, the relaxation time for the field-free rotation, after achieving a steady-state birefringence signal, was reduced from 316 ns for S1 to 269 ns for S1(T), at 3.7 degrees C, suggesting that trypsinolysis increases the flexibility of the connections between S1 segments or introduces additional segmental motions. For both S1 and S1(T), the rate of decay for a steady-state signal was independent of the field strength, between 3.34 and 20.3 statvolt/cm. Shortening the duration of the weak electric field pulses to 0.35 microseconds, so that steady-state signals were not achieved, decreased the relaxation times for S1 and S1(T) to 240 and 210 ns, respectively, which is consistent with the segmented flexible S1 structure proposed earlier [Highsmith, S., & Eden, D. (1986) Biochemistry 25, 2237]. When the strength of the electric field was increased to above 10 statvolt/cm, in order to make the interaction energy for the S1(T) electric dipole moment in the electric field greater than the thermal energy, the relaxation time after a 0.35-microseconds pulse decreased from 210 to 170 ns as the field was increased from 7 to 20 statvolt/cm. (ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

19.
The estimation of protein secondary structure from circular dichroism spectra is described by a multivariate linear model with noise (Gauss-Markoff model). With this formalism the adequacy of the linear model is investigated, paying special attention to the estimation of the error in the secondary structure estimates. It is shown that the linear model is only adequate for the alpha-helix class. Since the failure of the linear model is most likely due to nonlinear effects, a locally linearized model is introduced. This model is combined with the selection of the estimate whose fractions of secondary structure summate to approximately one. Comparing the estimation from the CD spectra with the X-ray data (by using the data set of W.C. Johnson Jr., 1988, Annu. Rev. Biophys. Chem. 17, 145-166) the root mean square residuals are 0.09 (alpha-helix), 0.12 (anti-parallel beta-sheet), 0.08 (parallel beta-sheet), 0.07 (beta-turn), and 0.09 (other). These residuals are somewhat larger than the errors estimated from the locally linearized model. In addition to alpha-helix, in this model the beta-turn and "other" class are estimated adequately. But the estimation of the antiparallel and parallel beta-sheet class remains unsatisfactory. We compared the linear model and the locally linearized model with two other methods (S. W. Provencher and J. Gl?ckner, 1981, Biochemistry 20, 1085-1094; P. Manavalan and W. C. Johnson Jr., 1988, Anal. Biochem. 167, 76-85). The locally linearized model and the Provencher and Gl?ckner method provided the smallest residuals. However, an advantage of the locally linearized model is the estimation of the error in the secondary structure estimates.  相似文献   

20.
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