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 62 毫秒
1.
The secondary structure of photosystem II reaction centers isolated from pea has been deduced from quantitative analysis of the component bands of the infrared amide I spectral region, determined by FTIR spectroscopy. The analysis shows the isolated complex to consist of 40% alpha-helix, 10% beta-sheet, 14% beta-strands (or extended chains), 17% turns, 15% loops, and 3% nonordered segments. These structural protein elements were determined for samples in H2O, in D2O, and in dried films. The isolated reaction center, composed of proteins D1,D2,cytochrome b559, and PsbI, has been predicted to contain a total of 13 transmembrane alpha-helices, which conveys a percentage of this type of structure congruent with the structural determination deduced from FTIR spectra. The process of thermal destabilization of the reaction centers has also been studied by FTIR spectroscopy, showing a clear main conformational transition at 42 degrees C, which indicates a high thermal sensitivity of the secondary structure of this protein complex. Such thermal instability may correlate with the well-described high sensitivity of photosystem II to damage and may relate to the process of rapid protein degradation that photosystem II suffers during photoinhibition of plants.  相似文献   

2.
N-terminally truncated recombinant 68-kDa human topoisomerase (topo) I exhibits the same DNA-driving activities as the wild-type protein. In the present study, Raman and circular dichroism techniques were employed for detailed structural characterization of the 68-kDa human topo I and its transformations induced by the suicide sequence-specific oligonucleotide (solig) binding and cleavage. Spectroscopic data combined with statistical prediction techniques were employed to construct a model of the secondary structure distribution along the primary protein structure in solution. The 68-kDa topo I was found to consist of ca. 59% alpha-helix, 24% beta-strand and/or sheets, and 17% other structures. A secondary structure transition of the 68-kDa topo I was found to accompany solig binding and cleavage. Nearly 15% of the alpha-helix of 68-kDa topo I is transferred within the other structures when in the complex with its DNA substrate. Raman spectroscopy analysis also shows redistribution of the structural rotamers of the 68-kDa topo I disulfide bonds and significant changes in the H-bonding of the Tyr residues and in the microenvironment/conformation of the Trp side chains. No structural modifications of the DNA substrate were detected by spectroscopic techniques. The data presented provide the first direct experimental evidence of the human topo I conformational transition after the cleavage step in the reaction of binding and cleavage of DNA substrate by the enzyme. This evidence supports the model of the enzyme function requiring the protein conformational transition. The most probable location of the enzyme transformations was the core and the C-terminal conservative 68-kDa topo I structural domains. By contrast, the linker domain was found to have an extremely low potential for solig-induced structural transformations. The pattern of redistribution of protein secondary structures induced by solig binding and covalent suicide complex formation supports the model of an intramolecular bipartite mode of topo I/DNA interaction in the substrate binding and cleavage reaction.  相似文献   

3.
1. The effect of guanidine hydrochloride (GuHCl) on pig heart lipoamide dehydrogenase [NADH: lipoamide oxidoreductase, EC 1.6.4.3.] was investigated by means of enzymatic activity and optical measurements (CD, absorption, and fluorescence spectra). The activity of the enzyme decreased on increasing the concentration of GuHCl and the enzyme was completely inactivated in 2.0 M GuHCl. 2. The contents of alpha-helix, beta, and unordered forms in lipoamide dehydrogenase were estimated to be 34, 14, and 52%, respectively. On increasing the concentration of GuHCl, the content of alpha-helix in lipoamide dehydrogenase decreased, whereas the content of the beta form hardly changed. 3. The native lipoamide dehydrogenase showed absorption, CD, and fluorescence spectra characteristic of bound FAD in the visible region, suggesting hydrophobic interaction between the protein moiety and FAD chromophore. The absorption, CD, and fluorescence spectra of the enzyme in 2.0 M GuHCl were similar to those of free FAD in the buffer, suggesting the release of FAD from the protein moiety. 4. The protein fluorescence spectrum of lipoamide dehydrogenase had a maximum at 350 nm blue-shifted by 8 nm from that of tryptophan in aqueous solution. The maximum of the enzyme in 2.0 M GuHCl was red-shifted to 357 nm. This suggests exposure of tryptophan residues to a polar environment. The maximum, 352nm, of the apoenzyme shifted to 350 nm on addition of FAD. These results show that the conformation in the microenvironment of some tryptophan residues in lipoamide dehydrogenase is affected by the dissociation-association of FAD. 5. The contents of alpha-helix, beta, and unordered forms in the apoenzyme were estimated to be 35, 8, and 57%, respectively. These values are similar to those of the native holoenzyme. The alpha-helical structure in the apoenzyme molecule was more sensitive to GuHCl than that in the holoenzyme. FAD and two hydrophobic probes, 8-anilinonaphthalene-1-sulfonate (ANS) and 4 benzolamido-4'-aminostilbene-2,2'-disulfonate (MBAS), which can bind to the apoenzyme, stabilized the alpha-helical structure in the apoenzyme molecule.  相似文献   

4.
The conformation of lens crystallins in vivo or in a highly concentrated solution is not well established. Most studies were carried out in dilute solutions in which protein-protein interaction is minimal. In order to see whether there is conformational change (tertiary and secondary structures) when crystallin solutions are brought to high concentrations, we have performed the following molecular spectroscopic measurements: circular dichroism (CD) and Fourier transform infrared (FTIR). Near-UV CD measurements showed a more than two-fold increase in CD intensity (molar ellipticity) for the total water-soluble (WS) protein from young calf lens nucleus in a highly concentrated solution (> 300 mg/ml in a 0.01-mm cell), when compared with a dilute solution (1000-fold dilution in a 10-mm cell). The individual crystallins in concentrated solutions also showed an increase in CD intensity, but of different magnitude: alpha-crystallin > beta-crystallin > gamma-crystallin. The increased CD indicates that lens crystallins are in a more compact structure in highly concentrated solutions; they likely undergo a transition from a mobile to an immobile state. Change in near-UV CD usually is caused by restricted mobility of aromatic side groups, particularly Trp. The transition involves not only a change in protein tertiary and/or quaternary structure, but also in protein backbone structure. The change of protein backbone structure was drawn from FTIR measurements. FTIR spectra, sensitive to the secondary structure in the amide I region, could be measured for a highly concentrated solution for which far-UV CD measurement is not feasible. The secondary structure that showed prominent change for alpha-crystallin in a highly concentrated solution was beta-conformation: increase in beta-turn with a concomitant decrease of alpha-helix structure.  相似文献   

5.
Conformational changes in the prion protein (PrP) seem to be responsible for prion diseases. We have used conformation-dependent chemical-shift measurements and rotational-resonance distance measurements to analyze the conformation of solid-state peptides lacking long-range order, corresponding to a region of PrP designated H1. This region is predicted to undergo a transformation of secondary structure in generating the infectious form of the protein. Solid-state NMR spectra of specifically 13C-enriched samples of H1, residues 109-122 (MKHMAGAAAAGAVV) of Syrian hamster PrP, have been acquired under cross-polarization and magic-angle spinning conditions. Samples lyophilized from 50% acetonitrile/50% water show chemical shifts characteristic of a beta-sheet conformation in the region corresponding to residues 112-121, whereas samples lyophilized from hexafluoroisopropanol display shifts indicative of alpha-helical secondary structure in the region corresponding to residues 113-117. Complete conversion to the helical conformation was not observed and conversion from alpha-helix back to beta-sheet, as inferred from the solid-state NMR spectra, occurred when samples were exposed to water. Rotational-resonance experiments were performed on seven doubly 13C-labeled H1 samples dried from water. Measured distances suggest that the peptide is in an extended, possibly beta-strand, conformation. These results are consistent with the experimental observation that PrP can exist in different conformational states and with structural predictions based on biological data and theoretical modeling that suggest that H1 may play a key role in the conformational transition involved in the development of prion diseases.  相似文献   

6.
The native tropomyosin molecule is a parallel, registered, alpha-helical coiled coil made from two 284-residue chains. Long excised subsequences (> or = 95 residues) form the same structure with comparable thermal stability. Here, we investigate local stability using shorter subsequences (20-50 residues) that are chemically synthesized or excised from various regions along the protein chain. Thermal unfolding studies of such shorter peptides by CD in the same solvent medium used in extant studies of the parent protein indicate very low helix content, almost no coiled-coil formation, and high thermal lability of such secondary structure as does form. This behavior is in stark contrast to extant data on leucine-zipper peptides and short "designed" synthetic peptides, many of which have high alpha-helix content and form highly stable coiled coils. The existence of short coiled coils calls into question the older idea that short subsequences of a protein have little structure. The present study supports the older view, at least in its application to tropomyosin. The intrinsic local alpha-helical propensity and helix-helix interaction in this prototypical alpha-helical protein is sufficiently weak as to require not only dimerization, but macro-molecular amplification in order to attain its native conformation in common benign media near neutral pH.  相似文献   

7.
The effect of ethanol on the secondary conformational structure of proteins of the human gastric mucosa was investigated by attenuated total reflection/Fourier transform infrared (ATR/FT-IR) spectroscopy. The IR peak intensity and position of each structural component of gastric mucosa was found to change significantly with the ethanol concentration and length of exposure. The peak intensity due to the beta-sheet and/or beta-turn conformational structure in amide I and II bands of gastric mucosa clearly increased after treatment with ethanol. Moreover, the peak at 1635 cm-1 shifted to 1630 cm-1 after treatment with 40% ethanol for 3 h, or 80% ethanol for 1 h, and a distinct shoulder also appeared at 1643 cm-1. This shift occurred more rapidly and was more pronounced after exposure of mucosa to 80% ethanol, compared with the effect of 40% ethanol, but the alpha-helical structure at the amide I and II bands was not influenced by either concentration of ethanol. Ethanol treatment might also transform the secondary structure of amide III in gastric mucosa from an alpha-helix to a mainly random coil with extensive unfolding. The absorption between 1180 and 980 cm-1, which is assigned to glycoprotein structure, was also reduced after treatment with ethanol. This strongly indicates that ethanol influences the conformation of the lipids and proteins of human gastric mucosa, leading to their deformation.  相似文献   

8.
Dimethyl sulfate modification was used to probe for tertiary structural elements in the group II intron PI.LSU/2 from the mitochondrial pre-ribosomal RNA of the brown alga Pylaiella littoralis. Modification of the lariat form of the intron under conditions that allow both native folding and conformational homogeneity is found to be generally consistent with secondary and tertiary structural features identified previously for group II ribozymes. A comparison of chemical probing at temperatures just below and above the first melting transition illustrates the cooperative unfolding of tertiary structure and identifies novel candidates for tertiary interactions in addition to defining elements of secondary structure. Substitution of the GAAA terminal loop of domain V is shown to be compatible with retention of conformational homogeneity (despite the loss of an important tertiary interaction), but produces a concise methylation footprint in domain I at the site previously shown to harbor the receptor for that loop. The analysis also identified two nucleotide positions in domain V with novel secondary and potential tertiary structural roles. The proposed refinement of domain V secondary structure is supported by an expanded comparative analysis of group II sequences and bears increased resemblance to U2:U6 snRNA pairing in the spliceosome.  相似文献   

9.
N-Terminal fragments of the rat liver elongation factor EF-2 containing 162 (17 kDa) and 244 (27 kDa) amino acid residues of 857 (95 kDa) residues of the native protein were synthesized in E. coli cells and in a wheat germ cell-free translation system, and their conformations were studied. Both fragments were synthesized as inclusion bodies (nonspecific molecular aggregates). The conformations of the fragments in a solution were studied at neutral pH values by CD, fluorescence spectroscopy, scanning microcalorimetry, viscosimetry, gel-filtration, limited proteolysis, and interaction with monospecific anti-EF-2 antibodies and GroEL/ES molecular chaperone. Under nondenaturing conditions, both fragments existed in a solution as associates within a broad range of molecular masses, contained a considerable amount of elements of the intramolecular secondary structure, and represented globules without rigid tertiary structure (molten globules). A rigid tertiary structure was not formed even after the interaction of the fragments with the GroEL/ES molecular chaperone, thus indicating that the C-terminal fragment is essential for the formation of the rigid tertiary structure. Both fragments contained conformational antigenic determinants similar to those in the whole protein; i.e., despite the absence of the rigid tertiary structure, the fragments contained elements whose structure was similar to that of the corresponding regions in the whole protein.  相似文献   

10.
In aqueous solution, while cytochrome c is a stably folded protein with a tightly packed structure at the secondary and tertiary levels, its heme-free precursor, apocytochrome c, shows all features of a structureless random coil. However, upon interaction with phospholipid vesicles or lysophospholipid micelles, apocytochrome c undergoes a conformational transition from its random coil in solution to an alpha-helical structure on association with lipid. The driving forces of this lipid-induced folding process of apocytochrome c were investigated for the interaction with various phospholipids and lysophospholipids. Binding of apocytochrome c to negatively charged phospholipid vesicles induced a partially folded state with approximately 85% of the alpha-helical structure of cytochrome c in solution. In contrast, in the presence of zwitterionic phospholipid vesicles, apocytochrome c remains a random coil, suggesting that negatively charged phospholipid headgroups play an important role in the mechanism of lipid-induced folding of apocytochrome c. However, negatively charged lysophospholipid micelles induce a higher content of alpha-helical structure than equivalent negatively charged diacylphospholipids in bilayers, reaching 100% of the alpha-helix content of cytochrome c in solution. Furthermore, micelles of lysolipids with the same zwitterionic headgroup of phospholipid bilayer vesicles induce approximately 60% of the alpha-helix content of cytochrome c in solution. On the basis of these results, we propose a mechanism for the folding of apocytochrome c induced by the interaction with lipid, which accounts for both electrostatic and hydrophobic contributions. Electrostatic lipid-protein interactions appear to direct the polypeptide to the micelle or vesicle surface and to induce an early partially folded state on the membrane surface. Hydrophobic interactions between nonpolar residues in the protein and the hydrophobic core of the lipid bilayer stabilize and extend the secondary structure upon membrane insertion.  相似文献   

11.
Human tyrosine hydroxylase exists as four isoforms (hTH1-4), generated by alternative splicing of pre-mRNA, with tissue-specific distribution. Unphosphorylated hTH3 and hTH1 were produced in large amounts in Escherichia coli and purified to homogeneity. The phosphorylation sites were determined after labeling with [32P]phosphate in the presence of cAMP-dependent protein kinase (PKA) and calmodulin-dependent protein kinase II (CaM-PKII). Ser40 was phosphorylated by PKA, and both Ser19 and Ser40 were phosphorylated by CaM-PKII. The enzyme kinetics of hTH3 were determined in the presence of various concentrations of the natural co-substrate (6R)-tetrahydrobiopterin and compared with those of recombinant hTH1 (similar to rat TH). We show that, under initial velocity conditions, excess (6R)-tetrahydrobiopterin inhibits hTH3 and hTH1. The TH catalytic constants (kcat) were determined for each of the two isoenzymes: hTH3 is about five times more active than hTH1. Phosphorylation by CaM-PKII did not affect the kinetic parameters of hTH3. The classical activation of TH by PKA phosphorylation, demonstrated for hTH1, was not observed with hTH3. Furthermore, hTH3 escapes activity regulation by phosphorylation and is always more active than phosphorylated hTH1. The properties of the hTH3 enzyme may be relevant to diseases affecting dopaminergic cells.  相似文献   

12.
An amino-terminal deletion mutant (residues 1-43) and a carboxy-terminal deletion mutant (residues 187-243) of human apoliprotein A-I (apo hA-I) have been produced from a bacterial expression system to explore the importance of the missing residues for the conformation of apo hA-I. Our focus has been to study the lipid-free structure of apo hA-I to understand how discrete domains influence the conformational plasticity of the protein and, by inference, the mechanism of lipid binding. All spectral and physical measurements indicate that both apo delta(1-43)A-I and apo delta(187-243)A-I have folded, tertiary structures. These structures differ in the specific arrangement of helical domains based, in part, on their relative thermodynamic stability, near- and far-UV CD, limited proteolysis, and the accessibility of tryptophans to fluorescence quenchers. In addition, all data indicate that the folded domains of apo hA-I and apo delta(187-243)A-I are very similar. Results from analytical ultracentrifugation suggest that lipid-free apo hA-I and the deletion mutants each exist in a dynamic equilibrium between a loosely folded, helical bundle and an elongated monomeric helical hairpin. The conformational heterogeneity is consistent with significant ANS binding exhibited by all three proteins and could help to explain the facile lipid binding properties of apo hA-I.  相似文献   

13.
The conformational properties of the molecular chaperone GroEL in the presence of ATP, its non-hydrolyzable analog 5'-adenylimidodiphosphate (AMP-PNP), and ADP have been analyzed by differential scanning calorimetry (DSC), Fourier-transform infra-red (FT-IR) and fluorescence spectroscopy. Nucleotide binding to one ring promotes a decrease in the Tm value of the GroEL thermal transition that is reversed when both rings are filled with nucleotide, indicating that the sequential occupation of the two protein rings by these nucleotides has different effects on the GroEL thermal denaturation process. In addition, ATP induces a conformational change in GroEL characterized by (a) the appearance of a reversible low temperature endotherm in the DSC profiles of the protein, and (b) an enhanced binding of the hydrophobic probe 8-anilino-naphthalene-1-sulfonate (ANS), which strictly depends on ATP hydrolysis. The similar sensitivity to K+ of the temperature range where activation of the GroEL ATPase activity, the low temperature endotherm, and the increase of the ANS fluorescence are abserved strongly indicates the existence of a conformational state of GroEL during ATP hydrolysis, different from that generated on ADP or AMP-PNP binding. To achieve this intermediate conformation, GroEL mainly modifies its tertiary and quaternary structures, leading to an increased exposure of hydrophobic surfaces, with minor rearrangements of its secondary structure.  相似文献   

14.
Conformational studies on an isolated integral membrane protein are reported. Lipoprotein of Escherichia coli outer membrane was released from murein by treatment with either lysozyme or trypsin. The isolated lysozyme-released lipoprotein (lipoprotein I) contained 2 or 3 muropeptides covalently linked at the C-terminal end, while the trypsin-released lipoprotein (lipoprotein II) was free of muropeptides and lacked the C-terminal peptide Tyr-Arg-Lys. Circular dichroism spectra of the two preparations were essentially identical, and they show an alpha-helix content of about 80%. According to calculations based on the Chou-Fasman rules for proteins of known sequence, lipoprotein is 64% alpha-helix and 15% beta-structure. Infrared spectroscopy qualitatively supports these values. The conformation was stable in the pH range of 5 - 12. Danaturation of lipoprotein by heat, 8 M urea, or sodium dodecylsulphate was a fully reversible, cooperative process. The thermal denaturation of lipoprotein occurs in two steps with transition points at 79.4 degrees C for lipoprotein I and at 85.1 degrees C for lipoprotein II. Lioprotein markedly changes conformation at dodecylsulphate concentrations where micelle formation sets in. The unusual behaviour of the lipoprotein convormation in sodium dodecylsulphate is discussed in relation to the lipoprotein conformation and aggregation within the membrane.  相似文献   

15.
Parathyroid hormone (PTH) regulates mineral metabolism and bone turnover by activating specific receptors located on osteoblastic and renal tubular cells and is fully functional as the N-terminal 1-34 fragment, PTH-(1-34). Previously, a "U-shaped" conformation with N- and C-terminal helices brought in close proximity by a turn has been postulated. The general acceptance of this hypothesis, despite limited experimental evidence, has altered the direction of the design of PTH-analogs. Examining the structure of human PTH-(1-34) under conditions that encompass the different environments the hormone may experience in the approach to and interaction with the G-protein-coupled receptor (including benign aqueous and saline solutions and in the presence of dodecylphosphocholine), we observe no evidence for a U-shape conformation or any tertiary structure. Instead, the N- and C-terminal helical domains, which vary in length and stability depending on the conditions, are separated by a highly flexible region of undefined conformation. These observations are in complete accord with recent conformational studies of PTH-related protein analogs containing lactams (Mierke, D. F., Maretto, S., Schievano, E. , DeLuca, D., Bisello, A., Mammi, S., Rosenblatt, M., Peggion, E., and Chorev, M. (1997) Biochemistry 36, 10372-10383) or a model amphiphilic alpha-helix (Pellegrini, M., Bisello, A., Rosenblatt, M., Chorev, M., and Mierke, D. F. (1997) J. Med. Chem. 40, 3025-3031). Reliable structural data from different environmental conditions are absolutely requisite for the next step in the design of non-peptide PTH analogs.  相似文献   

16.
17.
Acid-induced unfolding of proteins often results in an intermediate structure, called the molten globule structure or "A" state, which retains at least partial secondary structure but lacks a rigid tertiary structure. Acid-induced unfolding has been studied extensively for alpha-helical proteins, while few studies have been done on proteins containing only beta-strands. Tumor necrosis factor-alpha (TNF-alpha) is a trimer in which the individual subunits consist of antiparallel beta-sheet, organized into a jellyroll beta-sandwich. We have found previously [Narhi et al. (1996) Biochemistry 35, 11447-11453] that thermal denaturation of TNF-alpha results in an aggregate which contains a substantial amount of alpha-helix and that the addition of trifluoroethanol induces alpha-helix in both murine and human TNF-alpha. Here we show that acid also can induce alpha-helix in these proteins. At acidic pH (below 4), both human and murine TNF-alpha convert to a monomeric form, as determined by sedimentation and diffusion constants obtained from sedimentation velocity experiments. The sedimentation coefficient indicated that this monomer was only slightly expanded relative to the native state. Near-UV circular dichroic (CD) analysis showed a loss of tertiary structure. These structural features coincide with the notion that the acid-induced structure of TNF-alpha is a molten globule. What is unique in this protein is that TNF-alpha acquires alpha-helical structure, which is not present in the native structure as determined by both CD and Fourier transform infrared spectroscopy. Even more surprising is that TNF-alpha at pH 3.3 undergoes a very gradual noncooperative change in secondary structure upon heating, which results in an increase in alpha-helical content. At pH 2.2 in the absence of salt, TNF-alpha shows considerable alpha-helix, although heating does not change the spectrum. At pH 2.2, physiological salt decreases the amount of alpha-helix at ambient temperature, and upon heating, we see the noncooperative increase in alpha-helix as observed at pH 3.3 with low salt. The addition of salt at low pH induces reassociation but to a range of oligomers rather than a unique trimer structure. This acid-induced formation of an alpha-helical monomer of TNF-alpha may be related to its known interaction with lipid bilayers.  相似文献   

18.
The effect of sodium dodecyl sulfate (SDS) on enzyme activity, electrophoretic behavior, and conformation of Escherichia coli beta-galactosidase is presented. Fourier-transform infrared spectroscopy (FT-IR), previously used to study the structure of native beta-galactosidase has been applied to examine the detergent effects on the enzyme. At 20 degrees C, the presence of 1% SDS does not cause appreciable changes in the secondary structure, and enzyme activity is preserved; however, 10% SDS produces complete enzyme inactivation and FT-IR spectroscopy indicates a concomitant change in conformation. Thermal denaturation of beta-galactosidase starts at approximately 53 degrees C in the absence and at approximately 46 degrees C in the presence of 1% SDS, indicating tertiary structure changes; also, a good correlation between structural (FT-IR) and functional (Arrhenius plots) data is observed. The secondary structure of thermally denatured beta-galactosidase contains mainly extended structures, and intermolecular interactions produce protein aggregation. In the presence of 10% SDS, however, the hydrophobic segments of the protein are stabilized by SDS into helical structures without protein aggregation. At 30 degrees C, in the presence of 1% SDS, two protein bands are resolved by gel electrophoresis, only one of them being active. A model for SDS-galactosidase interaction is proposed, according to which, at low surfactant concentrations, SDS molecules bind the outer surface of the protein, without affecting the protein core. Higher detergent concentrations produce a larger conformational change involving enzyme inactivation and increased accessibility of the solvent to the protein core. Increasing temperature in the presence of 10% SDS leads to a facilitated access of surfactant molecules to the inner protein regions and to an increase of the beta-galactosidase alpha-helical content.  相似文献   

19.
We are examining possible roles of native and non-native interactions in early events in protein folding by a systematic analysis of the structures of fragments of proteins whose folding pathways are well characterised. Seven fragments of the 110-residue protein barnase, corresponding to the progressive elongation from its N terminus, have been characterised by a battery of biophysical and spectroscopic methods. Barnase is a multi-modular protein that folds via an intermediate in which the C-terminal region of its major alpha-helix (alpha-helix1, residues Thr6-His18) is substantially formed as is also its anti-parallel beta-sheet, centred around a beta-hairpin (residues Ser92-Leu95). Fragments up to, and including, residues 1-95 (fragment B95), appeared to be mainly disordered, although a small amount of helical secondary structure in each was inferred from far-UV CD experiments, and fluorescence studies indicated some native-like tertiary interactions in B95. The largest fragment (residues 1-105, B105) is compactly folded. The secondary structure in alpha-helix1 in the seven fragments was found by NMR to increase with increasing chain length faster than the build-up of tertiary interactions, indicating that alpha-helix1 is being stabilised by non-native interactions. This behaviour contrasts with that in fragments of the 64-residue chymotrypsin inhibitor 2 (CI2), in which tertiary and secondary structures build up in parallel with increasing length. CI2 consists of a single module of structure that folds without a detectable intermediate. The largest fragment of barnase, B105, has interactions that resemble its folding intermediate, whereas one of the largest fragments of CI2 (residues 1-60) resembles the folding transition state. The folding pathways of both proteins are consistent with a scheme in which there are low levels of native-like secondary structure in the denatured state that become stabilised by long-range interactions as folding proceeds. Neither protein forms a stable fold when lacking the last ten residues at the C terminus. Since at least 20 amino acid residues are bound to the ribosome during protein biosynthesis, these small proteins do not fold until they have left the ribosome, and so the studies of the folding of such proteins in vitro may be relevant to their folding in vivo, especially as the molecular chaperone GroEL binds only weakly to denatured CI2 and does not discernibly alter the folding mechanism of barnase.  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