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.
  总被引:3,自引:5,他引:3  
A chemically synthesized gene for ribonuclease A has been expressed in Escherichia coli using a T7 expression system (Studier, F.W., Rosenberg, A.H., Dunn, J.J., & Dubendorff, J.W., 1990, Methods Enzymol. 185, 60-89). The expressed protein, which contains an additional N-terminal methionine residue, has physical and catalytic properties close to those of bovine ribonuclease A. The expressed protein accumulates in inclusion bodies and has scrambled disulfide bonds; the native disulfide bonds are regenerated during purification. Site-directed mutations have been made at each of the two cis proline residues, 93 and 114, and a double mutant has been made. In contrast to results reported for replacement of trans proline residues, replacement of either cis proline is strongly destabilizing. Thermal unfolding experiments on four single mutants give delta Tm approximately equal to 10 degrees C and delta delta G0 (apparent) = 2-3 kcal/mol. The reason is that either the substituted amino acid goes in cis, and cis<==>trans isomerization after unfolding pulls the unfolding equilibrium toward the unfolded state, or else there is a conformational change, which by itself is destabilizing relative to the wild-type conformation, that allows the substituted amino acid to form a trans peptide bond.  相似文献   

2.
    
Under physiological conditions, peptidyl-prolyl cis/trans isomerases catalyze powerfully the cis/trans isomerization of the -Xaa-Pro- bond (Xaa: natural amino acids) in oligopeptides and proteins (PPIases; EC 5.2.1.8). However, incorporation of proline containing tetrapeptide-4-nitroanilides in micelles and phospholipid vesicles also leads to increased rates of this unimolecular conformational interconversion. The isomerization rate was dependent on the detergent and vesicle concentration, respectively. The observed rate constants fit the pseudophase model of micellar catalysis allowing the calculation of micellar turnover numbers (kcismic) and dissociation constants (KCmic). Comparing kcismic values to the rate constants of the uncatalyzed cis to trans isomerization, an acceleration factor of about 20-fold was obtained for Suc-Ala-Phe-Pro-Phe-4-nitroanilide (Suc: succinyl) bound to zwitterionic SB12 (N-dodecyl-N,N-dimethylammonium-3-propanesulfonate) micelles. In addition, a marked increase in the population of the trans conformer relative to cis was noted for all investigated combinations of peptides and detergents. In a series of tetrapeptides, Suc-Ala-Xaa-Pro-Phe-4-nitroanilide kcismic/KCmic as well as kcismic values are linearly correlated with the high performance liquid chromatography capacity factor R′ describing the hydrophobicity of the amino acid Xaa. The same correlation can describe quantitatively the dependency of kcar/Km on substrate hydrophobicity for the FKBP12-catalyzed isomerization. Despite the great differences in catalytic power, these results confirm the suspicion that micelles and FKBP12 may share a common component in the catalytic mechanism of peptidyl-prolyl bond isomerization. © 1997 John Wiley & Sons, Inc. Biopoly 42: 49–60, 1997  相似文献   

3.
    
Inspection of high resolution three-dimensional (3D) structures from the protein database reveals an increasing number of cis-Xaa-Pro and cis-Xaa-Yaa peptide bonds. However, we are still far from being able to predict whether these bonds will remain cis upon single-site substitution of Pro or Yaa and/or cleavage of a peptide bond close to it in the sequence. We have chosen oxidized Escherichia coli thioredoxin (Trx), a member of the Trx superfamily with a single alpha/beta domain and cis P76 to determine the effect of single-site substitution and/or cleavage on this isomer. Standard two-dimensional (2D) NMR analysis were performed on cleaved Trx (1-73/74-108) and its P76A variant. Analysis of the NOE connectivities indicates remarkable similarity between the secondary and supersecondary structure of the noncovalent complexes and Trx. Analysis of the 2D version of the HCCH-TOCSY and HMQC-NOESY-HMQC and 13C-filtered HMQC-NOESY spectra of cleaved Trx with uniformly 13C-labeled 175 and P76 shows surprising conservation of both cis P76 and packing of 175 against W31. A similar NMR analysis of its P76A variant provides no evidence for cis A76 and shows only subtle local changes in both the packing of 175 and the interstrand connectivities between its most protected hydrophobic strands (beta2 and beta4). Indeed, a molecular simulation model for the trans P76A variant of Trx shows only subtle local changes around the substitution site. In conclusion, cleavage of R73 is insufficient to provoke cis/trans isomerization of P76, but cleavage and single-site substitution (P76A) favors the trans isomer.  相似文献   

4.
    
  相似文献   

5.
    
The protein folding process is often in vitro rate‐limited by slow cis‐trans proline isomerization steps. Importantly, the rate of this process in vivo is accelerated by prolyl isomerases (PPIases). The archetypal PPIase is the human cyclophilin 18 (Cyp18 or CypA), and Arg 55 has been demonstrated to play a crucial role when studying short peptide substrates in the catalytic action of Cyp18 by stabilizing the transition state of isomerization. However, in this study we show that a R55A mutant of Cyp18 is as efficient as the wild type to accelerate the refolding reaction of human carbonic anhydrase II (HCA II). Thus, it is evident that the active‐site located Arg 55 is not required for catalysis of the rate‐limiting prolyl cis‐trans isomerization steps during the folding of a protein substrate as HCA II. Nevertheless, catalysis of cis‐trans proline isomerization in HCA II occurs in the active‐site of Cyp18, since binding of the inhibitor cyclosporin A abolishes rate acceleration of the refolding reaction. Obviously, the catalytic mechanisms of Cyp18 can differ when acting upon a simple model peptide, four residues long, with easily accessible Pro residues compared with a large protein molecule undergoing folding with partly or completely buried Pro residues. In the latter case, the isomerization kinetics are significantly slower and simpler mechanistic factors such as desolvation and/or strain might operate during folding‐assisted catalysis, since binding to the hydrophobic active site is still a prerequisite for catalysis.  相似文献   

6.
    
The cis/trans isomerization of peptidyl-prolyl peptide bonds is often the bottleneck of the refolding reaction for proteins containing cis proline residues in the native state. Proline (Pro) analogues, especially C4-substituted fluoroprolines, have been widely used in protein engineering to enhance the thermodynamic stability of peptides and proteins and to investigate folding kinetics. 4-thiaproline (Thp) has been shown to bias the ring pucker of Pro, to increase the cis population percentage of model peptides in comparison to Pro, and to diminish the activation energy barrier for the cis/trans isomerization reaction. Despite its intriguing properties, Thp has been seldom incorporated into proteins. Moreover, the impact of Thp on the folding kinetics of globular proteins has never been reported. In this study, we show that upon incorporation of Thp at cisPro76 into the thioredoxin variant Trx1P the half-life of the refolding reaction decreased from ~2 h to ~35 s. A dramatic acceleration of the refolding rate could be observed also for the protein pseudo wild-type barstar upon replacement of cisPro48 with Thp. Quantum chemical calculations suggested that the replacement of the CγH2 group by a sulfur atom in the pyrrolidine ring, might lower the barrier for cis/trans rotation due to a weakened peptide bond. The protein variants retained their thermodynamic stability upon incorporation of Thp, while the catalytic and enzymatic activities of the modified Trx1P remained unchanged. Our results show that the Pro isostere Thp might accelerate the rate of the slow refolding reaction for proteins containing cis proline residues in the native state, independent from the local structural environment.  相似文献   

7.
    
To investigate the structural origin of decreased pressure and temperature stability, the crystal structure of bovine pancreatic ribonuclease A variants V47A, V54A, V57A, I81A, I106A, and V108A was solved at 1.4–2.0 Å resolution and compared with the structure of wild‐type protein. The introduced mutations had only minor influence on the global structure of ribonuclease A. The structural changes had individual character that depends on the localization of mutated residue, however, they seemed to expand from mutation site to the rest of the structure. Several different parameters have been evaluated to find correlation with decrease of free energy of unfolding ΔΔGT, and the most significant correlation was found for main cavity volume change. Analysis of the difference distance matrices revealed that the ribonuclease A molecule is organized into five relatively rigid subdomains with individual response to mutation. This behavior consistent with results of unfolding experiments is an intrinsic feature of ribonuclease A that might be surviving remnants of folding intermediates and reflects the dynamic nature of the molecule. Proteins 2009. © 2009 Wiley‐Liss, Inc.  相似文献   

8.
    
Photoactive yellow protein (PYP), a blue-light photoreceptor for Ectothiorhodospira halophila, has provided a unique system for studying protein folding that is coupled with a photocycle. Upon receptor activation by blue light, PYP proceeds through a photocycle that includes a partially folded signaling state. The last-step photocycle is a thermal recovery reaction from the signaling state to the native state. Bi-exponential kinetics had been observed for the last-step photocycle; however, the slow phase of the bi-exponential kinetics has not been extensively studied. Here we analyzed both fast and slow phases of the last-step photocycle in PYP. From the analysis of the denaturant dependence of the fast and slow phases, we found that the last-step photocycle proceeds through parallel channels of the folding pathway. The burial of the solvent-accessible area was responsible for the transition state of the fast phase, while structural rearrangement from the compact state to the native state was responsible for the transition state of the slow phase. The photocycle of PYP was linked to the thermodynamic cycle that includes both unfolding and refolding of the fast- and slow-phase intermediates. In order to test the hypothesis of proline-limited folding for the slow phase, we constructed two proline mutants: P54A and P68A. We found that only a single phase of the last-step photocycle was observed in P54A. This suggests that there is a low energy barrier between trans to cis conformation in P54 in the light-induced state of PYP, and the resulting cis conformation of P54 generates a slow-phase kinetic trap during the photocycle-coupled folding pathway of PYP.  相似文献   

9.
    
Hypoxanthine-guanine phosphoribosyltransferase (HGPRT) is a key enzyme of the purine recycling pathway that catalyzes the conversion of 5-phospho-ribosyl-α-1-pyrophosphate and guanine or hypoxanthine to guanosine monophosphate (GMP) or inosine monophosphate (IMP), respectively, and pyrophosphate (PPi). We report the first crystal structure of a fungal 6-oxopurine phosphoribosyltransferase, the Saccharomyces cerevisiae HGPRT (Sc-HGPRT) in complex with GMP. The crystal structures of full length protein with (WT1) or without (WT2) sulfate that mimics the phosphate group in the PPi binding site were solved by molecular replacement using the structure of a truncated version (Δ7) solved beforehand by multiwavelength anomalous diffusion. Sc-HGPRT is a dimer and adopts the overall structure of class I phosphoribosyltransferases (PRTs) with a smaller hood domain and a short two-stranded parallel β-sheet linking the N- to the C-terminal end. The catalytic loops in WT1 and WT2 are in an open form while in Δ7, due to an inter-subunit disulfide bridge, the catalytic loop is in either an open or closed form. The closure is concomitant with a peptide plane flipping in the PPi binding loop. Moreover, owing the flexibility of a GGGG motif conserved in fungi, all the peptide bonds of the phosphate binding loop are in trans conformation whereas in nonfungal 6-oxopurine PRTs, one cis-peptide bond is required for phosphate binding. Mutations affecting the enzyme activity or the previously characterized feedback inhibition by GMP are located at the nucleotide binding site and the dimer interface.  相似文献   

10.
    
Zhang J  Germann MW 《Biopolymers》2011,95(11):755-762
Secondary amide cis peptide bonds are of even lower abundance than the cis tertiary amide bonds of prolines, yet they are of biochemical importance. Using 2D NMR exchange spectroscopy (EXSY) we investigated the formation of cis peptide bonds in several oligopeptides: Ac-G-G-G-NH(2) , Ac-I-G-G-NH(2) , Ac-I-G-G-N-NH(2) and its cyclic form: I-G-G-N in dimethylsulfoxide (DMSO). From the NMR studies, using the amide protons as monitors, an occurrence of 0.13-0.23% of cis bonds was obtained at 296 K. The rate constants for the trans to cis conversion determined from 2D EXSY spectroscopy were 4-9 × 10(-3) s(-1) . Multiple minor conformations were detected for most peptide bonds. From their thermodynamic and kinetic properties the cis isomers are distinguished from minor trans isomers that appear because of an adjacent cis peptide bond. Solvent and sequence effects were investigated utilizing N-methylacetamide (NMA) and various peptides, which revealed a unique enthalpy profile in DMSO. The cyclization of a tetrapeptide resulted in greatly lowered cis populations and slower isomerization rates compared to its linear counterpart, further highlighting the impact of structural constraints.  相似文献   

11.
    
In proteins and peptides, the vast majority of peptide bonds occurs in trans conformation, but a considerable fraction (about 5%) of X-Pro bonds adopts the cis conformation. Here we study the conservation of cis prolyl residues in evolutionary related proteins. We find that overall, in contrast to local, protein sequence similarity is a clear indicator for the conformation of prolyl residues. We observe that cis prolyl residues are more often conserved than trans prolyl residues, and both are more conserved than the surrounding amino acids, which show the same extent of conservation as the whole protein. The pattern of amino acid exchanges differs between cis and trans prolyl residues. Also, the cis prolyl bond is maintained in proteins with sequence identity as low as 20%. This finding emphasizes the importance of cis peptide bonds in protein structure and function.  相似文献   

12.
Visual sensitivity can be tuned by differential expression of opsin genes. Among African cichlid fishes, seven cone opsin genes are expressed in different combinations to produce diverse visual sensitivities. To determine the genetic architecture controlling these adaptive differences, we analysed genetic crosses between species expressing different complements of opsin genes. Quantitative genetic analyses suggest that expression is controlled by only a few loci with correlations among some genes. Genetic mapping identifies clear evidence of trans‐acting factors in two chromosomal regions that contribute to differences in opsin expression as well as one cis‐regulatory region. Therefore, both cis and trans regulation are important. The simple genetic architecture suggested by these results may explain why opsin gene expression is evolutionarily labile, and why similar patterns of expression have evolved repeatedly in different lineages.  相似文献   

13.
    
The pseudoproline residue (ΨPro, L ‐2,2‐dimethyl‐1,3‐thiazolidine‐4‐carboxylic acid) has been introduced into heterochiral diproline segments that have been previously shown to facilitate the formation of β‐hairpins, containing central two and three residue turns. NMR studies of the octapeptide Boc‐Leu‐Phe‐Val‐DPro‐ΨPro‐Leu‐Phe‐Val‐OMe ( 1 ), Boc‐Leu‐Val‐Val‐DPro‐ΨPro‐Leu‐Val‐Val‐OMe ( 2 ), and the nonapeptide sequence Boc‐Leu‐Phe‐Val‐DPro‐ΨPro‐DAla‐Leu‐Phe‐Val‐OMe ( 3 ) established well‐registered β‐hairpin structures in chloroform solution, with the almost exclusive population of the trans conformation for the peptide bond preceding the ΨPro residue. The β‐hairpin conformation of 1 is confirmed by single crystal X‐ray diffraction. Truncation of the strand length in Boc‐Val‐DPro‐ΨPro‐Leu‐OMe ( 4 ) results in an increase in the population of the cis conformer, with a cis/trans ratio of 3.65. Replacement of ΨPro in 4 by LPro in 5 , results in almost exclusive population of the trans form, resulting in an incipient β‐hairpin conformation, stabilized by two intramolecular hydrogen bonds. Further truncation of the sequence gives an appreciable rise in the population of cis conformers in the tripeptide Piv‐DPro‐ΨPro‐Leu‐OMe ( 6 ). In the homochiral segment Piv‐Pro‐ΨPro‐Leu‐OMe ( 7 ) only the cis form is observed with the NMR evidence strongly supporting a type VIa β‐turn conformation, stabilized by a 4→1 hydrogen bond between the Piv (CO) and Leu (3) NH groups. The crystal structure of the analog peptide 7a (Piv‐Pro‐ΨH,CH3Pro‐Leu‐NHMe) confirms the cis peptide bond geometry for the Pro‐ΨH,CH3Pro peptide bond, resulting in a type VIa β‐turn conformation. © 2009 Wiley Periodicals, Inc. Biopolymers (Pept Sci) 92: 405–416, 2009. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

14.
    
Peptomers are oligomeric molecules composed of both α‐amino acids and N‐substituted glycine monomers, thus creating a hybrid of peptide and peptoid units. Peptomers have been used in several applications such as antimicrobials, protease inhibitors, and antibody mimics. Despite the considerable promise of peptomers as chemically diverse molecular scaffolds, we know little about their conformational tendencies. This lack of knowledge limits the ability to implement computational approaches for peptomer design. Here we computationally evaluate the local structural propensities of the peptide–peptoid linkage. We find some general similarities between the peptide residue conformational preferences and the Ramachandran distribution of residues that precede proline in folded protein structures. However, there are notable differences. For example, several β‐turn motifs are disallowed when the i+2 residue is also a peptoid monomer. Significantly, the lowest energy geometry, when dispersion forces are accounted for, corresponds to a “cis‐Pro touch‐turn” conformation, an unusual turn motif that has been observed at protein catalytic centers and binding sites. The peptomer touch‐turn thus represents a useful design element for the construction of folded oligomers capable of molecular recognition and as modules in the assembly of structurally complex peptoid–protein hybrid macromolecules. © 2014 Wiley Periodicals, Inc. Biopolymers (Pept Sci) 102: 369–378, 2014.  相似文献   

15.
    
YbbR domains are widespread throughout Eubacteria and are expressed as monomeric units, linked in tandem repeats or cotranslated with other domains. Although the precise role of these domains remains undefined, the location of the multiple YbbR domain‐encoding ybbR gene in the Bacillus subtilis glmM operon and its previous identification as a substrate for a surfactin‐type phosphopantetheinyl transferase suggests a role in cell growth, division, and virulence. To further characterize the YbbR domains, structures of two of the four domains (I and IV) from the YbbR‐like protein of Desulfitobacterium hafniense Y51 were solved by solution nuclear magnetic resonance and X‐ray crystallography. The structures show the domains to have nearly identical topologies despite a low amino acid identity (23%). The topology is dominated by β‐strands, roughly following a “figure 8” pattern with some strands coiling around the domain perimeter and others crossing the center. A similar topology is found in the C‐terminal domain of two stress‐responsive bacterial ribosomal proteins, TL5 and L25. Based on these models, a structurally guided amino acid alignment identifies features of the YbbR domains that are not evident from naïve amino acid sequence alignments. A structurally conserved cis‐proline (cis‐Pro) residue was identified in both domains, though the local structure in the immediate vicinities surrounding this residue differed between the two models. The conservation and location of this cis‐Pro, plus anchoring Val residues, suggest this motif may be significant to protein function.  相似文献   

16.
β2‐Microglobulin has been a model system for the study of fibril formation for 20 years. The experimental study of β2‐microglobulin structure, dynamics, and thermodynamics in solution, at atomic detail, along the pathway leading to fibril formation is difficult because the onset of disorder and aggregation prevents signal resolution in Nuclear Magnetic Resonance experiments. Moreover, it is difficult to characterize conformers in exchange equilibrium. To gain insight (at atomic level) on processes for which experimental information is available at molecular or supramolecular level, molecular dynamics simulations have been widely used in the last decade. Here, we use molecular dynamics to address three key aspects of β2‐microglobulin, which are known to be relevant to amyloid formation: (1) 60 ns molecular dynamics simulations of β2‐microglobulin in trifluoroethanol and in conditions mimicking low pH are used to study the behavior of the protein in environmental conditions that are able to trigger amyloid formation; (2) adaptive biasing force molecular dynamics simulation is used to force cis‐trans isomerization at Proline 32 and to calculate the relative free energy in the folded and unfolded state. The native‐like trans‐conformer (known as intermediate 2 and determining the slow phase of refolding), is simulated for 10 ns, detailing the possible link between cis‐trans isomerization and conformational disorder; (3) molecular dynamics simulation of highly concentrated doxycycline (a molecule able to suppress fibril formation) in the presence of β2‐microglobulin provides details of the binding modes of the drug and a rationale for its effect. Proteins 2011. © 2010 Wiley‐Liss, Inc.  相似文献   

17.
【背景】作为降解木聚糖的核心酶种,木聚糖酶可以有效促进木质纤维素的消化水解,在动物养殖领域应用广泛。来源于嗜热细菌贝斯其热解纤维素菌(Caldicellulosiruptorbescii)的GH10家族木聚糖酶Cb Xyn10C最适温度为85℃,在80℃条件下具有良好的热稳定性,具有饲料工业应用潜力。【目的】为满足饲料制粒尤其是水产饲料加工过程的工艺要求,进一步提高木聚糖酶Cb Xyn10C的热稳定性并阐明其耐热机理。【方法】以Cb Xyn10C晶体结构为基础,采用刚性氨基酸引入、疏水作用网络重排2种策略对其热稳定性进行理性设计,获得在100℃条件下比活提高的单点突变体后,通过有益突变位点叠加策略进一步提升酶的热稳定性,最后采用分子动力学模拟技术分析其热稳定性提高的分子机制。【结果】共获得了4个稳定性提高的单点突变体A45P、T69P、F309V和A325P,其中突变体A45P效果最优。随着在A45P基础上另外3个突变位点的叠加,酶的热稳定性在不损失酶活的前提下得到了逐步提升。获得的四点突变体A45P/F309V/A325P/T69P的耐热性最好,其最适反应温度和熔解温度Tm值较野生型...  相似文献   

18.
β-Site amyloid precursor protein (APP) cleaving enzyme 1 (BACE1) is the transmembrane aspartyl protease that catalyzes the first cleavage step in the proteolysis of the APP to the amyloid β-protein (Aβ), a process involved in the pathogenesis of Alzheimer disease. BACE1 pre-mRNA undergoes complex alternative splicing, the regulation of which is not well understood. We identified a G-rich sequence within exon 3 of BACE1 involved in controlling splice site selection. Mutation of the G-rich sequence decreased use of the normal 5' splice site of exon 3, which leads to full-length and proteolytically active BACE1, and increased use of an alternative splice site, which leads to a shorter, essentially inactive isoform. Nuclease protection assays, nuclear magnetic resonance, and circular dichroism spectroscopy revealed that this sequence folds into a G-quadruplex structure. Several proteins were identified as capable of binding to the G-rich sequence, and one of these, heterogeneous nuclear ribonucleoprotein H, was found to regulate BACE1 exon 3 alternative splicing and in a manner dependent on the G-rich sequence. Knockdown of heterogeneous nuclear ribonucleoprotein H led to a decrease in the full-length BACE1 mRNA isoform as well as a decrease in Aβ production from APP, suggesting new possibilities for therapeutic approaches to Alzheimer's disease.  相似文献   

19.
    
Escherichia coli (strain K‐12, substrain MG1655) glycerol dehydrogenase (GldA) is required to catalyze the first step in fermentative glycerol metabolism. The protein was expressed and purified to homogeneity using a simple combination of heat‐shock and chromatographic methods. The high yield of the protein (∼250 mg per litre of culture) allows large‐scale production for potential industrial applications. Purified GldA exhibited a homogeneous tetrameric state (∼161 kDa) in solution and relatively high thermostability (Tm = 65.6°C). Sitting‐drop sparse‐matrix screens were used for protein crystallization. An optimized condition with ammonium sulfate (2 M) provided crystals suitable for diffraction, and a binary structure containing glycerol in the active site was solved at 2.8 Å resolution. Each GldA monomer consists of nine β‐strands, thirteen α‐helices, two 310‐helices and several loops organized into two domains, the N‐ and C‐terminal domains; the active site is located in a deep cleft between the two domains. The N‐terminal domain contains a classic Rossmann fold for NAD+ binding. The O1 and O2 atoms of glycerol serve as ligands for the tetrahedrally coordinated Zn2+ ion. The orientation of the glycerol within the active site is mainly stabilized by van der Waals and electrostatic interactions with the benzyl ring of Phe245. Computer modeling suggests that the glycerol molecule is sandwiched by the Zn2+ and NAD+ ions. Based on this, the mechanism for the relaxed substrate specificity of this enzyme is also discussed.  相似文献   

20.
A series of 18 differently substituted new aryl hetaryl ketones and thioketones were synthesized in four to six steps from commercial starting materials. The new ketones were evaluated as inhibitors of the peptidyl‐prolyl cis‐trans isomerase hPin1 with Ki values ranging in the one‐digit micromolar to sub‐micromolar numbers. A crystal structure revealed the non‐planar arrangement of the aryl residues at the carbonyl compound and supports the hypothesis that the new compounds might mimic the transition state of the enzymatic conversion.  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