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 765 毫秒
1.
C C Chen  Y Zhu  J A King  L B Evans 《Biopolymers》1992,32(10):1375-1392
Under physiological conditions, many polypeptide chains spontaneously fold into discrete and tightly packed three-dimensional structures. The folded polypeptide chain conformation is believed to represent a minimum Gibbs energy of the system, governed by the weak interactions that operate between the amino acid residues and between the residues and the solvent. A semiempirical molecular thermodynamic model is proposed to represent the Gibbs energy of folding of aqueous homopolypeptide systems. The model takes into consideration both the entropy contribution and the enthalpy contribution of folding homopolypeptide chains in aqueous solutions. The entropy contribution is derived from the Flory-Huggins expression for the entropy of mixing. It accounts for the entropy loss in folding a random-coiled polypeptide chain into a specific polypeptide conformation. The enthalpy contribution is derived from a molecular segment-based Non-Random Two Liquid (NRTL) local composition model [H. Renon and J. M. Prausnitz (1968) AIChE J., Vol. 14, pp. 135-142; C.-C. Chen and L. B. Evans (1986) AIChE J., Vol. 32, pp. 444-454], which takes into consideration of the residue-residue, residue-solvent, and solvent-solvent binary physical interactions along with the local compositions of amino acid residues in aqueous homopolypeptides. The UNIFAC group contribution method [A. Fredenslund, R. L. Jones, and J. M. Prausnitz (1975) AIChE J., 21, 1086-1099; A. Fredenslund, J. Gmehling, and P. Rasmussen (1977) Vapor-Liquid Equilibrium Using UNIFAC, Elsevier Scientific Publishing Company, Amsterdam], developed originally to estimate the excess Gibbs energy of solutions of small molecules, was used to estimate the NRTL binary interaction parameters. The model yields a hydrophobicity scale for the 20 amino acid side chains, which compares favorably with established scales [Y. Nozaki and C. Tanford (1971) Journal of Biological Chemistry, Vol. 46, pp. 2211-2217; E. B. Leodidis and T. A. Hatton (1990) Journal of Physical Chemistry, Vol. 94, pp. 6411-6420]. In addition, the model generates qualitatively correct thermodynamic constants and it accurately predicts thermodynamically favorable folding of a number of aqueous homopolypeptides from random-coiled states into alpha-helices. The model further facilitates estimation of the Zimm-Bragg helix growth parameter s and the nucleation parameter sigma for amino acid residues [B. H. Zimm and J. K. Bragg (1959) Journal of Chemical Physics, Vol. 31, pp. 526-535]. The calculated values of the two parameters fall into the ranges suggested by Zimm and Bragg.  相似文献   

2.
J A Zitzewitz  C R Matthews 《Biochemistry》1999,38(31):10205-10214
The alpha subunit of tryptophan synthase (alphaTS) from Escherichia coli is a 268-residue 8-stranded beta/alpha barrel protein. Two autonomous folding units, comprising the first six strands (residues 1-188) and the last two strands (residues 189-268), have been previously identified in this single structural domain protein by tryptic digestion [Higgins, W., Fairwell, T., and Miles, E. W. (1979) Biochemistry 18, 4827-4835]. The larger, amino-terminal fragment, alphaTS(1-188), was overexpressed and independently purified, and its equilibrium and kinetic folding properties were studied by absorbance, fluorescence, and near- and far-UV circular dichroism spectroscopies. The native state of the fragment unfolds cooperatively in an apparent two-state transition with a stability of 3.98 +/- 0.19 kcal mol(-1) in the absence of denaturant and a corresponding m value of 1.07 +/- 0.05 kcal mol(-1) M(-1). Similar to the full-length protein, the unfolding of the fragment shows two kinetic phases which arise from the presence of two discrete native state populations. Additionally, the fragment exhibits a significant burst phase in unfolding, indicating that a fraction of the folded state ensemble under native conditions has properties similar to those of the equilibrium intermediate populated at 3 M urea in full-length alphaTS. Refolding of alphaTS(1-188) is also complex, exhibiting two detectable kinetic phases and a burst phase that is complete within 5 ms. The two slowest isomerization phases observed in the refolding of the full-length protein are absent in the fragment, suggesting that these phases reflect contributions from the carboxy-terminal segment. The folding mechanism of alphaTS(1-188) appears to be a simplified version of the mechanism for the full-length protein [Bilsel, O., Zitzewitz, J. A., Bowers, K.E, and Matthews, C. R.(1999) Biochemistry 38, 1018-1029]. Four parallel channels in the full-length protein are reduced to a pair of channels that most likely reflect a cis/trans proline isomerization reaction in the amino-terminal fragment. The off- and on-pathway intermediates that exist for both full-length alphaTS and alphaTS(1-188) may reflect the preponderance of local interactions in the beta/alpha barrel motif.  相似文献   

3.
X-ray absorption spectroscopic (XAS) studies on cytochrome C1 from beef heart mitochondria were conducted to identify the effect of the hinge protein [Kim, C.H., & King, T.E. (1983) J. Biol. Chem. 258, 13543-13551] on the structure of the heme site in cytochrome c1. A comparison of XAS data of highly purified "one-band" and "two-band" cytochrome c1 [Kim, C.H., & King, T.E. (1987) Biochemistry 26, 1955-1961] demonstrates that the hinge protein exerts a rather pronounced effect on the heme environment of the cytochrome c1: a conformational change occurs within a radius of approximately 5 A from the heme iron in cytochrome c1 when the hinge protein is bound to cytochrome c1. This result may be correlated with the previous observations that the structure and reactivity of cytochrome c1 are affected by the hinge protein [Kim, C.H., & King, T.E. (1987) Biochemistry 26, 1955-1961; Kim, C.H., Balny, C., & King, T.E. (1987) J. Biol. Chem. 262, 8103-8108].  相似文献   

4.
M J Rooman  J P Kocher  S J Wodak 《Biochemistry》1992,31(42):10226-10238
A recently developed procedure to predict backbone structure from the amino acid sequence [Rooman, M., Kocher, J. P., & Wodak, S. (1991) J. Mol. Biol, 221, 961-979] is fine tuned to identify protein segments, of length 5-15 residues, that adopt well-defined conformations in the absence of tertiary interactions. These segments are obtained by requiring that their predicted lowest energy structures have a sizable energy gap relative to other computed conformations. Applying this procedure to 69 proteins of known structure, we find that regions with largest energy gaps--those having highly preferred conformations--are also the most accurately predicted ones. On the basis of previous findings that such regions correlate well with sites that become structured early during folding, our approach provides the means of identifying such sites in proteins without prior knowledge of the tertiary structure. Furthermore, when predictions are performed so as to ignore the influence of residues flanking each segment along the sequence, a situation akin to excising the considered peptide from the rest of the chain, they offer the possibility of identifying protein segments liable to adopt well-defined conformations on their own. The described approach should have useful applications in experimental and theoretical investigations of protein folding and stability, and aid in designing peptide drugs and vaccines.  相似文献   

5.
R Schwyzer 《Biochemistry》1986,25(15):4281-4286
Equilibrium thermodynamic and kinetic estimations were used to confirm the rather unusual conformation, orientation, and accumulation of dynorphin A-(1-13)-tridecapeptide (dynorphin1-13) on the surface of neutral lipid membranes, as observed by Erne et al. [Erne, D., Sargent, D. F., & Schwyzer, R. (1985) Biochemistry 24, 4261-4263]. I started from the premise that the most stable conformation of molecularly disperse peptides in contact with the hydrophobic phase of a membrane is helical [Henderson, R. (1979) Soc. Gen. Physiol. Ser. 33, 3-15]. Calculation of the Gibbs free energy difference for the transfer of increasing numbers m of N-terminal residues of dynorphin1-13 from their random-coil conformation in water to their alpha-helical conformation in a hydrophobic phase, with the values provided by Von Heijne and Blomberg [Von Heijne, G., & Blomberg, C. (1979) Eur. J. Biochem. 97, 175-181], showed an energy minimum at m = 9 that corresponded to the observed apparent association constant of 9 X 10(4) L/mol. This confirmed our experimental observations. The orientation of dynorphin1-13 in the interphase was estimated by calculation of the molecular amphiphilic moment A. This force vector was defined in analogy to the "helical" and "structural" hydrophobic moments of Eisenberg et al. [Eisenberg, D., Weiss, R. M., & Terwilliger, T. C. (1982) Nature (London) 299, 371-374]. It takes into account the segregation of hydrophobic and hydrophilic residues with respect to the center of the alpha-helix. A peptide located in a hydrophobic-hydrophilic gradient experiences a torque that tends to orient A in a direction perpendicular to the surfaces of equal hydrophobicity.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

6.
T A Holak  J H Prestegard 《Biochemistry》1986,25(19):5766-5774
Sequence-specific assignments of 1H NMR resonances were obtained for the backbone protons in acyl carrier protein (ACP) from Escherichia coli, a protein of 77 residues. The observations, in the NOESY spectra, of 1H-1H sequential and medium-range connectivities indicate the presence of three or four alpha-helical segments joined by short sequences of mixed conformations. The observations are used to refine a secondary structure model previously proposed on the basis of a Chou-Fasman algorithm [Rock, C. O., & Cronan, J. E., Jr. (1979) J. Biol. Chem. 254, 9778-9785].  相似文献   

7.
The equilibrium and kinetic properties for the urea-induced unfolding of the alpha subunit of tryptophan synthase from Escherichia coli, Salmonella typhimurium, and five interspecies hybrids were compared to determine the role of protein folding in evolution. The parent proteins differ at 40 positions in the sequence of 268 amino acids, and the hybrids differ by up to 15 amino acids from the Escherichia coli alpha subunit. The results show that all the proteins follow the same folding mechanism and are consistent with a previously proposed hypothesis [Hollecker, M., & Creighton, T. E. (1983) J. Mol. Biol. 168, 409; Krebs, H., Schmid, F. X., & Jaenicke, R. (1983) J. Mol. Biol. 169, 619] that the folding mechanisms are conserved in homologous proteins. Analysis of the kinetic data suggests that the 15 positions at which the parent proteins differ in the amino folding unit, residues 1-188, do not play a role in a rate-limiting step in folding that has been previously identified as the association of the amino and carboxyl folding units [Beasty, A. M., Hurle, M. R., Manz, J. T., Stackhouse, T. S., Onuffer, J. J., & Matthews, C. R. (1986) Biochemistry 25, 2965]. One or more of the 25 positions at which the parent proteins differ in the carboxyl folding unit, residues 189-268, do appear to play a role in this same rate-limiting step.  相似文献   

8.
Overman SA  Thomas GJ 《Biochemistry》1999,38(13):4018-4027
The study of filamentous virus structure by Raman spectroscopy requires accurate band assignments. In previous work, site- and residue-specific isotope substitutions were implemented to elucidate definitive assignments for Raman bands arising from vibrational modes of the alpha-helical coat protein main chain and aromatic side chains in the class I filamentous phage, fd [Overman, S. A., and Thomas, G. J., Jr. (1995) Biochemistry 34, 5440-5451; Overman, S. A., and Thomas, G. J., Jr. (1998) Biochemistry 37, 5654-5665]. Here, we extend the previous methods and expand the assignment scheme to identify Raman markers of nonaromatic side chains of the coat protein in the native fd assembly. This has been accomplished by Raman analysis of 11 different fd isotopomers selectively incorporating deuterium at specific sites in either alanine, aspartic acid, glutamic acid, glycine, isoleucine, leucine, lysine, serine, or valine residues of the coat protein. Raman markers are also identified for the corresponding deuterated side chains. In combination with previous assignments, the results provide a comprehensive understanding of coat protein contributions to the Raman signature of the fd virion and validate Raman markers assigned to the packaged single-stranded DNA genome. The findings described here show that nonaromatic side chains contribute prolifically to the fd Raman signature, that marker bands for specific nonaromatics differ in general from those observed in corresponding polypeptides and amino acids, and that the frequencies and intensities of many nonaromatic markers are sensitive to secondary and higher-order structures. Nonaromatic markers within the 1200-1400 cm-1 interval also interfere seriously with the diagnostic Raman amide III band that is normally exploited in secondary structure analysis. Implications of these findings for the assessment of protein conformation by Raman spectroscopy are considered.  相似文献   

9.
A N Stroup  L M Gierasch 《Biochemistry》1990,29(42):9765-9771
A family of mutants of the P22 bacteriophage tailspike protein has been characterized as temperature sensitive for folding (tsf) by King and co-workers [King, J. (1986) Bio/Technology 4, 297-303]. There is substantial evidence that the tsf mutations alter the folding pathway but not the stability of the final folded protein. Several point mutations are known to cause the tsf phenotype; most of these occur in regions of the tailspike sequence likely to take up reverse turns. Hence, it has been hypothesized that the correct folding of the P22 tailspike protein requires formation of turns and that the mutations causing tsf phenotypes interfere at this critical stage. We have tested this hypothesis by study of isolated peptides corresponding to a region of the P22 tailspike harboring a tsf mutation. Comparison of the tendencies of wild-type and tsf sequences to adopt turn conformations was achieved by the synthesis of peptides with flanking cysteine residues and the use of a thiol-disulfide exchange assay. We find that the wild-type sequence, either as a decapeptide (Ac-CVKFPGIETC-CONH2) or as a dodecapeptide (Ac-CYVKFPGIETLC-CONH2), has a 3-5-fold greater tendency for its termini to approach closely enough to form the intramolecular disulfide than do the peptide sequences corresponding to the tsf mutant sequences, which have a Gly----Arg substitution (Ac-CVKFPRIETC-CONH2 or Ac-CYVKFPRIETLC-CONH2). A peptide with a D-Arg substituted for the Gly has a slightly higher turn propensity than does the wild type. Together with data from nuclear magnetic resonance analysis of the oxidized peptides, this suggests that a type II beta turn is favored by the wild-type sequence. Our results on isolated peptides from the P22 tailspike protein support the model for its folding that includes reverse turn formation as a critical step.  相似文献   

10.
Equilibrium and kinetic studies on the folding of a series of amino acid replacements at position 211 in the alpha subunit of tryptophan synthase from Escherichia coli were performed in order to determine the role of this position in the rate-limiting step in folding. Previous studies [Beasty, A. M., Hurle, M. R., Manz, J. T., Stackhouse, T., Onuffer, J. J., & Matthews, C. R. (1986) Biochemistry 25, 2965-2974] have shown that the rate-limiting step corresponds to the association/dissociation of the amino (residues 1-188) and carboxy (residues 189-268) folding units. In terms of the secondary structure, the amino folding unit consists of the first six strands and five alpha helices of this alpha/beta barrel protein. The carboxy folding unit comprises the remaining two strands and three alpha helices; position 211 is in strand 7. Replacement of the wild-type glycine at position 211 with serine, valine, and tryptophan at most alters the rate of dissociation of the folding units; association is not changed significantly. In contrast, glutamic acid and arginine dramatically decelerate and accelerate, respectively, both association and dissociation. The difference in effects is attributed to long-range electrostatic interactions for these charged side chains; steric effects and/or hydrogen bonding play lesser roles. When considered with previous data on replacements at other positions in the alpha subunit [Hurle, M. R., Tweedy, N. B., & Matthews, C. R. (1986) Biochemistry 25, 6356-6360], it is clear that beta strands 6 (in the amino folding unit) and 7 (in the carboxy folding unit and containing position 211) dock late in the folding process.  相似文献   

11.
Time-resolved fluorescence anisotropy of a bound extrinsic probe was studied in an effort to characterize dynamic properties of the transient partially folded forms that appear during the folding of the alpha-subunit of tryptophan synthase (alphaTS) from Escherichia coli. Previous studies have shown that alphaTS, a single structural domain, can be cleaved into autonomously folding amino- and carboxy-folding units comprising residues 1-188 and 189-268, respectively [Higgins, W., Fairwell, T., and Miles, E. W. (1979) Biochemistry 18, 4827-4835]. By use of a double-kinetic approach [Jones, B. E., Beechem, J. M., and Matthews, C. R. (1995) Biochemistry 34, 1867-1877], the rotational correlation time of 1-anilino-8-naphthalene sulfonate bound to nonpolar surfaces of folding intermediates was measured by time-correlated single photon counting at varying time delays following initiation of folding from the urea-denatured form by stopped-flow techniques. Comparison of the rotational correlation times for the full-length alphaTS and the amino-terminal fragment suggests that folding of the amino-terminal fragment and carboxy-terminal fragment is coordinated, not autonomous, on the milliseconds to seconds time scale. If a spherical shape is assumed, the apparent hydrodynamic radius of alphaTS after 5 ms is 26.8 A. The radius increases to 28.5 A by 1 s before decreasing to the radius for native alphaTS, 24.7 A, on a longer time scale (>25 s). Viewed within the context of the kinetic folding model of alphaTS [Bilsel, O., Zitzewitz, J. A., Bowers, K. E. , and Matthews, C. R. (1999) Biochemistry 38, 1018-1029], the initial collapse reflects the formation of an off-pathway burst-phase intermediate in which at least part of the carboxy folding unit interacts with the amino folding unit. The subsequent increase in rotational correlation time corresponds to the formation of an on-pathway intermediate that leads to the native conformation. The apparent increase in the radius for the on-pathway intermediate may reflect a change in the interaction of the two-folding units, thereby forming a direct precursor for the alpha/beta barrel structure.  相似文献   

12.
In this study it is established by calculation which regular conformations single-stranded DNA and RNA can adopt in the complex with the single-stranded DNA binding protein GP32 of bacteriophage T4. In order to do so, information from previous experiments about base orientations and the length and diameter of the complexes is used together with knowledge about bond lengths and valence angles between chemical bonds. It turns out that there is only a limited set of similar conformations which are in agreement with experimental data. The arrangement of neighboring bases is such that there is ample space for aromatic residues of the protein to partly intercalate between the bases, which is in agreement with a previously proposed model for the binding domain of the protein [Prigodich, R. V., Shamoo, Y., Williams, K. R., Chase, J. W., Konigsberg, W. H., & Coleman, J. E. (1986) Biochemistry 25, 3666-3671]. Both C2'endo and C3'endo sugar conformations lead to calculated DNA conformations that are consistent with experimental data. The orientation of the O2' atoms of the sugars in RNA can explain why the binding affinity of GP32 for polyribonucleotides is lower than for polydeoxyribonucleotides.  相似文献   

13.
Simulation analysis of the stability mutant R96H of T4 lysozyme   总被引:14,自引:0,他引:14  
B Tidor  M Karplus 《Biochemistry》1991,30(13):3217-3228
Free energy simulation methods are used to analyze the effects of the mutation Arg 96----His on the stability of T4 lysozyme. The calculated stability change and the lack of significant structural rearrangement in the folded state due to the mutation are in agreement with experimental studies [Kitamura, S., & Sturtevant, J. M. (1989) Biochemistry 28, 3788-3792; Weaver, L. H., et al. (1989) Biochemistry 28, 3793-3797]. By use of thermodynamic integration, the contributions of specific interactions to the free energy change are evaluated. It is shown that a number of contributions that stabilize the wild type or the mutant partially cancel in the overall free energy difference; some of these involve the unfolded state. Comparison of the results with conclusions based on structural and thermodynamic data leads to new insights into the origin of the stability difference between wild-type and mutant proteins. Of particular interest is the importance of the contributions of more distant residues, solvent water, and the covalent linkage of the mutated amino acid. Also, the analysis of the interactions of Arg/His 96 with the C-terminal end of a helix (residues 82-90) makes it clear that the nearby carbonyl groups (Tyr 88 and Asp 89) make the dominant contribution, that the amide groups do not contribute significantly, and that the helix-dipole model is inappropriate for this case.  相似文献   

14.
Joseph E  Wei W  Tittmann K  Jordan F 《Biochemistry》2006,45(45):13517-13527
The X-ray crystal structure of pyruvamide-activated yeast pyruvate decarboxylase (YPDC) revealed a flexible loop spanning residues 290 to 304 on the beta-domain of the enzyme, not seen in the absence of pyruvamide, a substrate activator surrogate. Site-directed mutagenesis studies revealed that residues on the loop affect the activity, with some residues reducing k(cat)/K(m) by at least 1000-fold. In the pyruvamide-activated form, the loop located on the beta domain can transfer information to the active center thiamin diphosphate (ThDP) located at the interface of the alpha and gamma domains. The sigmoidal v(0)-[S] curve with wild-type YPDC attributed to substrate activation is modulated for most variants, but is not abolished. Pre-steady-state stopped-flow studies for product formation on these loop variants provided evidence for three enzyme conformations connected by two transitions, as already noted for the wild-type YPDC at pH 5.0 [Sergienko, E. A., and Jordan, F. (2002) Biochemistry 41, 3952-3967]. (1)H NMR analysis of the intermediate distribution resulting from acid quench [Tittmann et al. (2003) Biochemistry 42, 7885-7891] with all YPDC variants indicated that product release is rate limiting in the steady state. Apparently, the loop is not solely responsible for the substrate activation behavior, rather it may affect the behavior of residue C221 identified as the trigger for substrate activation. The most important function of the loop is to control the conformational equilibrium between the "open" and "closed" conformations of the enzyme identified in the pyruvamide-activated structure [Lu et al. (2000) Eur. J. Biochem. 267, 861-868].  相似文献   

15.
16.
Bacteriophage M13 major coat protein has been isolated with cholate and reconstituted in dimyristoyl- and dioleoylphosphatidylcholine (DMPC and DOPC, respectively) bilayers by dialysis. Fourier transform infrared spectra of DMPC/coat protein recombinants confirmed that, whereas the protein isolated by phenol extraction was predominantly in a beta-sheet conformation, the cholate-isolated coat protein contained a higher proportion of the alpha-helical conformation [cf. Spruijt, R. B., Wolfs, C. J. A. M., & Hemminga, M. A. (1989) Biochemistry 28, 9158-9165]. The cholate-isolated coat protein/lipid recombinants gave different electron spin resonance (ESR) spectral line shapes of incorporated lipid spin labels, as compared with those from recombinants with the phenol-extracted protein that were studied previously [Wolfs, C. J. A. M., Horváth, L. I., Marsh, D., Watts, A., & Hemminga, M. A. (1989) Biochemistry 28, 9995-10001]. Plots of the ratio of the fluid/motionally restricted components in the ESR spectra of spin-labeled phosphatidylglycerol were linear with respect to the lipid/protein ratio in the recombinants up to 20 mol/mol. The corresponding values of the relative association constants, Kr, and number of association sites, N1, on the protein were Kr approximately 1 and N1 approximately 4 for DMPC recombinants and Kr approximately 1 and N1 approximately 5 for DOPC recombinants. Simulation of the two-component lipid spin label ESR spectra with the exchange-coupled Bloch equations gave values for the off-rate of the lipids leaving the protein surface of 2.0 x 10(7) s-1 at 27 degrees C in DMPC recombinants and 3.0 x 10(7) s-1 at 24 degrees C in DOPC recombinants.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

17.
To investigate the pH dependence of the conformational stability of barnase, urea denaturation curves were determined over the pH range 2-10. The maximum conformational stability of barnase is 9 kcal mol-1 and occurs between pH 5 and 6. The dependence of delta G on urea concentration increases from 1850 cal mol-1 M-1 at high pH to about 3000 cal mol-1 M-1 near pH 3. This suggests that the unfolded conformations of barnase become more accessible to urea as the net charge on the molecule increases. Previous studies suggested that in 8 M urea barnase unfolds more completely than ribonuclease T1, even with the disulfide bonds broken [Pace, C.N., Laurents, D. V., & Thomson, J.A. (1990) Biochemistry 29, 2564-2572]. In support of this, solvent perturbation difference spectroscopy showed that in 8 M urea the Trp and Tyr residues in barnase are more accessible to perturbation by dimethyl sulfoxide than in ribonuclease T1 with the disulfide bonds broken.  相似文献   

18.
Y Gao  J Boyd  R J Williams  G J Pielak 《Biochemistry》1990,29(30):6994-7003
Resonance assignments for the main-chain, side-chain, exchangeable side chain, and heme protons of the C102T variant of Saccharomyces cerevisiae iso-1-cytochrome c in both oxidation states (with the exception of Gly-83) are reported. (We have also independently assigned horse cytochrome c.) Some additional assignments for the horse protein extend those of Wand and co-workers [Wand, A. J., Di Stefano, D. L., Feng, Y., Roder, H., & Englander, S. W. (1989) Biochemistry 28, 186-194; Feng, Y., Roder, H., Englander, S. W., Wand, A. J., & Di Stefano, D. L. (1989) Biochemistry 28, 195-203]. Qualitative interpretation of nuclear Overhauser enhancement data allows the secondary structure of these two proteins to be described relative to crystal structures. Comparison of the chemical shift of the backbone protons of the C102T variant and horse protein reveals significant differences resulting from amino acid substitution at positions 56 and 57 and further substitutions between residue 60 and residue 69. Although the overall folding of yeast iso-1-cytochrome c and horse cytochrome c is very similar, there can be large differences in chemical shift for structurally equivalent residues. Chemical shift differences of amide protons (and to a lesser extent alpha protons) represent minute changes in hydrogen bonding. Therefore, great care must be taken in the use of differences in chemical shift as evidence for structural changes even between highly homologous proteins.  相似文献   

19.
Thermodynamics of apocytochrome b5 unfolding.   总被引:4,自引:4,他引:0       下载免费PDF全文
Apocytochrome b5 from rabbit liver was studied by scanning calorimetry, limited proteolysis, circular dichroism, second derivative spectroscopy, and size exclusion chromatography. The protein is able to undergo a reversible two-state thermal transition. However, transition temperature, denaturational enthalpy, and heat capacity change are reduced compared with the holoprotein. Apocytochrome b5 stability in terms of Gibbs energy change at protein unfolding (delta G) amounts to delta G = 7 +/- 1 kJ/mol at 25 degrees C (pH 7.4) compared with delta G = 25 kJ/mol for the holoprotein. Apocytochrome b5 is a compact, native-like protein. According to the spectral data, the cooperative structure is mainly based in the core region formed by residues 1-35 and 79-90. This finding is in full agreement with NMR data (Moore, C.D. & Lecomte, J.T.J., 1993, Biochemistry 32, 199-207).  相似文献   

20.
Scott KA  Daggett V 《Biochemistry》2007,46(6):1545-1556
The problem of how a protein folds from a linear chain of amino acids to the three-dimensional structure necessary for function is often investigated using proteins with a low degree of sequence identity that adopt different folds. The design of pairs of proteins with a high degree of sequence identity but different folds offers the opportunity for a complementary study; in two highly similar sequences, which residues are the most important in directing folding to a particular structure? Here we use molecular dynamics simulations to characterize the folding-unfolding pathways of a pair of proteins designed by Bryan and co-workers [Alexander, P. A., et al. (2005) Biochemistry 44, 14045-14054; He, Y. N., et al. (2005) Biochemistry 44, 14055-14061]. Despite being 59% identical, the two protein sequences fold to two different structures. The first sequence folds to the alpha+beta protein G structure and the second to the all-alpha-helical protein A structure. We show that the final protein structure is determined early along the folding pathway. In folding to the protein G structure, the single alpha-helix (alpha1) and the beta3-beta4 turn fold early. Formation of the hairpin turn essentially prevents folding to helical structure in this region of the protein. This early structure is then consolidated by formation of long-range hydrophobic interactions between alpha1 and the beta3-beta4 turn. The protein A sequence differs both in the residues that form the beta3-beta4 turn and also in many of the residues that form the early hydrophobic interactions in the protein G structure. Instead, in the protein A sequence, a more hierarchical mechanism is observed, with helices folding before many of the tertiary interactions are formed. We find that small, but critical, sequence differences determine the topology of the protein early along the folding pathway, which help to explain the process by which one fold can evolve into another.  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