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 46 毫秒
1.
Recently, a syndrome of Mutant INS-gene-induced Diabetes of Youth (MIDY, derived from one of 26 distinct mutations) has been identified as a cause of insulin-deficient diabetes, resulting from expression of a misfolded mutant proinsulin protein in the endoplasmic reticulum (ER) of insulin-producing pancreatic beta cells. Genetic deletion of one, two, or even three alleles encoding insulin in mice does not necessarily lead to diabetes. Yet MIDY patients are INS-gene heterozygotes; inheritance of even one MIDY allele, causes diabetes. Although a favored explanation for the onset of diabetes is that insurmountable ER stress and ER stress response from the mutant proinsulin causes a net loss of beta cells, in this report we present three surprising and interlinked discoveries. First, in the presence of MIDY mutants, an increased fraction of wild-type proinsulin becomes recruited into nonnative disulfide-linked protein complexes. Second, regardless of whether MIDY mutations result in the loss, or creation, of an extra unpaired cysteine within proinsulin, Cys residues in the mutant protein are nevertheless essential in causing intracellular entrapment of co-expressed wild-type proinsulin, blocking insulin production. Third, while each of the MIDY mutants induces ER stress and ER stress response; ER stress and ER stress response alone appear insufficient to account for blockade of wild-type proinsulin. While there is general agreement that ultimately, as diabetes progresses, a significant loss of beta cell mass occurs, the early events described herein precede cell death and loss of beta cell mass. We conclude that the molecular pathogenesis of MIDY is initiated by perturbation of the disulfide-coupled folding pathway of wild-type proinsulin.  相似文献   

2.
Haloperoxidases are useful oxygenases involved in halogenation of a range of water‐insoluble organic compounds and can be used without additional high‐cost cofactors. In particular, organic solvent‐stable haloperoxidases are desirable for enzymatic halogenations in the presence of organic solvents. In this study, we adopted a directed evolution approach by error‐prone polymerase chain reaction to improve the organic solvent‐stability of the homodimeric BPO‐A1 haloperoxidase from Streptomyces aureofaciens. Among 1,000 mutant BPO‐A1 haloperoxidases, an organic solvent‐stable mutant OST48 with P123L and P241A mutations and a high active mutant OST959 with H53Y and G162R mutations were selected. The residual activity of mutant OST48 after incubation in 40% (v/v) 1‐propanol for 1 h was 1.8‐fold higher than that of wild‐type BPO‐A1. In addition, the OST48 mutant showed higher stability in methanol, ethanol, dimethyl sulfoxide, and N,N‐dimethylformamide than wild‐type BPO‐A1 haloperoxidase. Moreover, after incubation at 80°C for 1 h, the residual activity of mutant OST959 was 4.6‐fold higher than that of wild‐type BPO‐A1. Based on the evaluation of single amino acid‐substituted mutant models, stabilization of the hydrophobic core derived from P123L mutation and increased numbers of hydrogen bonds derived from G162R mutation led to higher organic solvent‐stability and thermostability, respectively. © 2015 American Institute of Chemical Engineers Biotechnol. Prog., 31:917–924, 2015  相似文献   

3.
The mechanisms of interfacial folding and membrane insertion of the Alzheimer's amyloid‐β fragment Aβ(25–35) and its less toxic mutant, N27A‐Aβ(25–35) and more toxic mutant, M35A‐Aβ(25–35), are investigated using replica–exchange molecular dynamics in an implicit water‐membrane environment. This study simulates the processes of interfacial folding and membrane insertion in a spontaneous fashion to identify their general mechanisms. Aβ(25–35) and N27A‐Aβ(25–35) peptides share similar mechanisms: the peptides are first located in the membrane hydrophilic region where their C‐terminal residues form helical structures. The peptides attempt to insert themselves into the membrane hydrophobic region using the C‐terminal or central hydrophobic residues. A small portion of peptides can successfully enter the membrane's hydrophobic core, led by their C‐terminal residues, through the formation of continuous helical structures. No detectable amount of M35A‐Aβ(25–35) peptides appeared to enter the membrane's hydrophobic core. The three studied peptides share a similar helical structure for their C‐terminal five residues, and these residues mainly buried within the membrane's hydrophobic region. In contrast, their N‐terminal properties are markedly different. With respect to the Aβ(25–35), the N27A‐Aβ(25–35) forms a more structured helix and is buried deeper within the membrane, which may result in a lower degree of aggregation and a lower neurotoxicity; in contrast, the less structured and more water‐exposed M35A‐Aβ(25–35) is prone to aggregation and has a higher neurotoxicity. Understanding the mechanisms of Aβ peptide interfacial folding and membrane insertion will provide new insights into the mechanisms of neurodegradation and may give structure‐based clues for rational drug design preventing amyloid associated diseases. Proteins 2010. © 2010 Wiley‐Liss, Inc.  相似文献   

4.
Influenza A (H5N1) virus is one of the world's greatest pandemic threats. Neuraminidase (NA) inhibitors, oseltamivir and zanamivir, prevent the spread of influenza, but drug‐resistant viruses have reduced their effectiveness. Resistance depends on the binding properties of NA‐drug complexes. Key residue mutations within the active site of NA glycoproteins diminish binding, thereby resulting in drug resistance. We performed molecular simulations and calculations to characterize the mechanisms of H5N1 influenza virus resistance to oseltamivir and predict potential drug‐resistant mutations. We examined two resistant NA mutations, H274Y and N294S, and one non‐drug‐resistant mutation, E119G. Six‐nanosecond unrestrained molecular dynamic simulations with explicit solvent were performed using NA‐oseltamivir complexes containing either NA wild‐type H5N1 virus or a variant. MM_PBSA techniques were then used to rank the binding free energies of these complexes. Detailed analyses indicated that conformational change of E276 in the Pocket 1 region of NA is a key source of drug resistance in the H274Y mutant but not in the N294S mutant.  相似文献   

5.
Coiled coil is a ubiquitous structural motif in proteins, with two to seven alpha helices coiled together like the strands of a rope, and coiled coil folding and assembly is not completely understood. A GCN4 leucine zipper mutant with four mutations of K3A, D7A, Y17W, and H18N has been designed, and the crystal structure has been determined at 1.6 Å resolution. The peptide monomer shows a helix trunk with short curved N‐ and C‐termini. In the crystal, two monomers cross in 35° and form an X‐shaped dimer, and each X‐shaped dimer is welded into the next one through sticky hydrophobic ends, thus forming an extended two‐stranded, parallel, super long coiled coil rather than a discrete, two‐helix coiled coil of the wild‐type GCN4 leucine zipper. Leucine residues appear at every seventh position in the super long coiled coil, suggesting that it is an extended super leucine zipper. Compared to the wild‐type leucine zipper, the N‐terminus of the mutant has a dramatic conformational change and the C‐terminus has one more residue Glu 32 determined. The mutant X‐shaped dimer has a large crossing angle of 35° instead of 18° in the wild‐type dimer. The results show a novel assembly mode and oligomeric state of coiled coil, and demonstrate that mutations may affect folding and assembly of the overall coiled coil. Analysis of the formation mechanism of the super long coiled coil may help understand and design self‐assembling protein fibers.  相似文献   

6.
The topology of the designed protein Top7 is not found in natural proteins. Top7 shows signatures of non‐cooperative folding in both experimental studies and computer simulations. In particular, molecular dynamics of coarse‐grained structure‐based models of Top7 show a well‐populated C‐terminal folding‐intermediate. Since most similarly sized globular proteins are cooperative folders, the non‐natural topology of Top7 has been suggested as a reason for its non‐cooperative folding. Here, we computationally examine the folding of Top7 with the intent of making it cooperative. We find that its folding cooperativity can be increased in two ways: (a) Optimization of packing interactions in the N‐terminal half of the protein enables further folding of the C‐terminal intermediate. (b) Reduction in the packing density of the C‐terminal region destabilizes the intermediate. In practice, these strategies are implemented in our Top7 model through modifications to the contact‐map. These modifications do not alter the topology of Top7 but result in cooperative folding. Amino‐acid mutations that mimic these modifications also lead to a significant increase in folding cooperativity. Finally, we devise a method to randomize the sizes of amino‐acids within the same topology, and confirm that the structure of Top7 makes its folding sensitive to packing interactions. In contrast, the ribosomal protein S6, which has secondary structure similar to Top7, has folding which is much less sensitive to packing perturbations. Thus, it should be possible to make a sequence fold cooperatively to the structure of Top7, but to do so its side‐chain packing needs to be carefully designed. Proteins 2014; 82:364–374. © 2013 Wiley Periodicals, Inc.  相似文献   

7.
Liang Xu  Yonggang Chen  Xiaojuan Wang 《Proteins》2014,82(12):3286-3297
Although the N‐terminal region of Amyloid β (Aβ) peptides plays dual roles as metal‐coordinating sites and conformational modulator, few studies have been performed to explore the effects of mutations at this region on the overall conformational ensemble of Aβ and the binding propensity of metal ions. In this work, we focus on how three familial Alzheimer's disease mutations (D7H, D7N, and H6R) alter the structural characteristics and thermodynamic stabilities of Aβ42 using molecular dynamics simulations. We observe that each mutation displays increased β‐sheet structures in both N and C termini. In particular, both the N terminus and central hydrophobic region of D7H can form stable β‐hairpin structures with its C terminus. The conserved turn structure at Val24–Lys28 in all peptides and Zn2+‐bound Aβ42 is confirmed as the common structural motif to nucleate folding of Aβ. Each mutant can significantly increase the solvation free energy and thus enhance the aggregation of Aβ monomers. The correlation dynamics between Aβ(1–16) and Aβ(17–42) fragments are elucidated by linking the domain motions with the corresponding structured conformations. We characterize the different populations of correlated domain motions for each mutant from a more macroscopic perspective, and unexpectedly find that Zn2+‐bound Aβ42 ensemble shares the same populations as Aβ42, indicating that the binding of Zn2+ to Aβ follows the conformational selection mechanism, and thus is independent of domain motions, even though the structures of Aβ have been modified at a residue level. Proteins 2014; 82:3286–3297. © 2014 Wiley Periodicals, Inc.  相似文献   

8.
Ji Guo Su  Wei Zu Chen  Cun Xin Wang 《Proteins》2010,78(9):2157-2169
The impacts of three charged‐residue‐involved mutations, E46A, R3E, and R3E/L66E, on the thermostability and folding behavior of the cold shock protein from the themophile Bacillus caldolyticus (Bc‐Csp) were investigated by using a modified Gō‐like model, in which the nonspecific electrostatic interactions of charged residues were taken into account. Our simulation results show that the wild‐type Bc‐Csp and its three mutants are all two‐sate folders, which is consistent with the experimental observations. It is found that these three mutations all lead to a decrease of protein thermodynamical stability, and the effect of R3E mutation is the strongest. The lower stability of these three mutants is due to the increase of the enthalpy of the folded state and the entropy of the unfolded state. Using this model, we also studied the folding kinetics and the folding/unfolding pathway of the wild‐type Bc‐Csp as well as its three mutants and then discussed the effects of electrostatic interactions on the folding kinetics. The results indicate that the substitutions at positions 3 and 46 largely decrease the folding kinetics, whereas the mutation of residue 66 only slightly decreases the folding rate. This result agrees well with the experimental observations. It is also found that these mutations have little effects on the folding transition state and the folding pathway, in which the N‐terminal β sheet folds earlier than the C‐terminal region. We also investigated the detailed unfolding pathway and found that it is really the reverse of the folding pathway, providing the validity of our simulation results. Proteins 2010. © 2010 Wiley‐Liss, Inc.  相似文献   

9.
The insulin gene mutation c.137G>A (R46Q), which changes an arginine at the B22 position of the mature hormone to glutamine, causes the monogenic diabetes variant maturity-onset diabetes of the young (MODY). In MODY patients, this mutation is heterozygous, and both mutant and wild-type (WT) human insulin are produced simultaneously. However, the patients often depend on administration of exogenous insulin. In this study, we chemically synthesized the MODY mutant [GlnB22]-insulin and characterized its biological and structural properties. The chemical synthesis of this insulin analogue revealed that its folding ability is severely impaired. In vitro and in vivo tests showed that its binding affinity and biological activity are reduced (both approximately 20% that of human insulin). Comparison of the solution structure of [GlnB22]-insulin with the solution structure of native human insulin revealed that the most significant structural effect of the mutation is distortion of the B20-B23 β-turn, leading to liberation of the B chain C-terminus from the protein core. The distortion of the B20-B23 β-turn is caused by the extended conformational freedom of the GlnB22 side chain, which is no longer anchored in a hydrogen bonding network like the native ArgB22. The partially disordered [GlnB22]-insulin structure appears to be one reason for the reduced binding potency of this mutant and may also be responsible for its low folding efficiency in vivo. The altered orientation and flexibility of the B20-B23 β-turn may interfere with the formation of disulfide bonds in proinsulin bearing the R46Q (GlnB22) mutation. This may also have a negative effect on the WT proinsulin simultaneously biosynthesized in β-cells and therefore play a major role in the development of MODY in patients producing [GlnB22]-insulin.  相似文献   

10.
In mutant INS gene–induced diabetes of youth (MIDY), characterized by insulin deficiency, MIDY proinsulin mutants misfold and fail to exit the endoplasmic reticulum (ER). Moreover, these mutants bind and block ER exit of wild-type (WT) proinsulin, inhibiting insulin production. The ultimate fate of ER-entrapped MIDY mutants is unclear, but previous studies implicated ER-associated degradation (ERAD), a pathway that retrotranslocates misfolded ER proteins to the cytosol for proteasomal degradation. Here we establish key ERAD machinery components used to triage the Akita proinsulin mutant, including the Hrd1-Sel1L membrane complex, which conducts Akita proinsulin from the ER lumen to the cytosol, and the p97 ATPase, which couples the cytosolic arrival of proinsulin with its proteasomal degradation. Surprisingly, we find that protein disulfide isomerase (PDI), the major protein oxidase of the ER lumen, engages Akita proinsulin in a novel way, reducing proinsulin disulfide bonds and priming the Akita protein for ERAD. Efficient PDI engagement of Akita proinsulin appears linked to the availability of Hrd1, suggesting that retrotranslocation is coordinated on the lumenal side of the ER membrane. We believe that, in principle, this form of diabetes could be alleviated by enhancing the targeting of MIDY mutants for ERAD to restore WT insulin production.  相似文献   

11.
The Schistosoma juponicum 26 kDa glutathione S‐transferase (sj26GST) consists of the N‐terminal domain (N‐domain), containing three alpha‐helices (named H1‐H3) and four anti‐parallel beta‐strands (S1‐S4), and the C‐terminal domain (C‐domain), comprising five alpha‐helices (named H4‐H8). In present work, molecular dynamics simulations and fluorescence spectroscopic were used to gain insights into the unfolding process of sj26GST. The molecular dynamics simulations on sj26GST subunit both in water and in 8 M urea were carried out at 300 K, 400 K and 500 K, respectively. Spectroscopic measurements were employed to monitor structural changes. Molecular dynamics simulations of sj26GST subunit induced by urea and temperature showed that the initial unfolding step of sj26GST both in water and urea occurred on N‐domain, involving the disruption of helices H2, H3 and strands S3 and S4, whereas H6 was the last region exposed to solution and was the last helix to unfold. Moreover, simulations analyses combining with fluorescence and circular dichroism spectra indicated that N‐domain could not fold independent, suggesting that correct folding of N‐domain depended on its interactions with C‐domain. We further proposed that the folding of GSTs could begin with the hydrophobic collapse of C‐domain whose H4, H5, H6 and H7 could move close to each other and form a hydrophobic core, especially H6 wrapped in the hydrophobic center and beginning spontaneous formation of the helix. S3, S4, H3, and H2 could form in the wake of the interaction between C‐domain and N‐domain. The paper can offer insights into the molecular mechanism of GSTs unfolding. © 2014 Wiley Periodicals, Inc. Biopolymers 103: 247–259, 2015.  相似文献   

12.
Upon chronic up-regulation of proinsulin synthesis, misfolded proinsulin can accumulate in the endoplasmic reticulum (ER) of pancreatic β-cells, promoting ER stress and type 2 diabetes mellitus. In Mutant Ins-gene-induced Diabetes of Youth (MIDY), misfolded mutant proinsulin impairs ER exit of co-expressed wild-type proinsulin, limiting insulin production and leading to eventual β-cell death. In this study we have investigated the hypothesis that increased expression of ER oxidoreductin-1α (Ero1α), despite its established role in the generation of H2O2, might nevertheless be beneficial in limiting proinsulin misfolding and its adverse downstream consequences. Increased Ero1α expression is effective in promoting wild-type proinsulin export from cells co-expressing misfolded mutant proinsulin. In addition, we find that upon increased Ero1α expression, some of the MIDY mutants themselves are directly rescued from ER retention. Secretory rescue of proinsulin-G(B23)V is correlated with improved oxidative folding of mutant proinsulin. Indeed, using three different variants of Ero1α, we find that expression of either wild-type or an Ero1α variant lacking regulatory disulfides can rescue mutant proinsulin-G(B23)V, in parallel with its ability to provide an oxidizing environment in the ER lumen, whereas beneficial effects were less apparent for a redox-inactive form of Ero1. Increased expression of protein disulfide isomerase antagonizes the rescue provided by oxidatively active Ero1. Importantly, ER stress induced by misfolded proinsulin was limited by increased expression of Ero1α, suggesting that enhancing the oxidative folding of proinsulin may be a viable therapeutic strategy in the treatment of type 2 diabetes.  相似文献   

13.
Multiple MD simulations were performed for the full‐length wild‐type A1, the full length A1 mutations S27E and S27A, as well as the N‐terminal peptide (AMVSEFLKQAWFIDNEEQEYIKTVKG S 27 KGGPGSAVSPYPTFN) of wild‐type A1 and mutations S27E and S27A. The MD simulation trajectories of about 350 ns were generated and analyzed to examine the changes of core domain calcium binding affinity, core domain and N‐terminal domain structures, and N‐terminal domain orientation. Our results indicated that S27A and S27E mutations caused little changes on the calcium‐binding affinity of the core domain of A1. However, the S27A mutation made the N‐terminal domain of A1 less helical, and made the N‐terminal domain migrate faster toward the core domain; these impacts on A1 are beneficial to the membrane aggregation process. On the contrary, the S27E mutation made the N‐terminal domain of A1 more stable, and hindered the migration to the core domain; these changes on A1 are antagonistic for the membrane aggregation process. Our results using MD simulations provide an atomistic explanation for experimental observations that the S27E mutant showed a higher calcium concentration requirement and lower maximal extent of aggregation, while the wild‐type and two mutants S27E and S27A required identical calcium concentrations for liposome binding. Proteins 2014; 82:3327–3334. © 2014 Wiley Periodicals, Inc.  相似文献   

14.
The nuclease domain of colicin E7 (NColE7) promotes the nonspecific cleavage of nucleic acids at its C‐terminal HNH motif. Interestingly, the deletion of four N‐terminal residues (446–449 NColE7 = KRNK) resulted in complete loss of the enzyme activity. R447A mutation was reported to decrease the nuclease activity, but a detailed analysis of the role of the highly positive and flexible N‐terminus is still missing. Here, we present the study of four mutants, with a decreased activity in the following order: NColE7  >> KGNK > KGNG ~ GGNK > GGNG. At the same time, the folding, the metal‐ion, and the DNA‐binding affinity were unaffected by the mutations as revealed by linear and circular dichroism spectroscopy, isothermal calorimetric titrations, and gel mobility shift experiments. Semiempirical quantum chemical calculations and molecular dynamics simulations revealed that K446, K449, and/or the N‐terminal amino group are able to approach the active centre in the absence of the other positively charged residues. The results suggested a complex role of the N‐terminus in the catalytic process that could be exploited in the design of a controlled nuclease.  相似文献   

15.
To define catalytically essential residues of bacteriophage T7 RNA polymerase, we have generated five mutants of the polymerase, D537N, K631M, Y639F, H811Q and D812N, by site-directed mutagenesis and purified them to homogeneity. The choice of specific amino acids for mutagenesis was based upon photoaffinity-labeling studies with 8-azido-ATP and homology comparisons with the Klenow fragment and other DNA/RNA polymerases. Secondary structural analysis by circular dichroism indicates that the protein folding is intact in these mutants. The mutants D537N and D812N are totally inactive. The mutant K631M has 1% activity, confined to short oligonucleotide synthesis. The mutant H811Q has 25% activity for synthesis of both short and long oligonucleotides. The mutant Y639F retains full enzymatic activity although individual kinetic parameters are somewhat different. Kinetic parameters, (kcat)app and (Km)app for the nucleotides, reveal that the mutation of Lys to Met has a much more drastic effect on (kcat)app than on (Km)app, indicating the involvement of K631 primarily in phosphodiester bond formation. The mutation of His to Gln has effects on both (kcat)app and (Km)app; namely, three- to fivefold reduction in (kcat)app and two- to threefold increase in (Km)app, implying that His811 may be involved in both nucleotide binding and phosphodiester bond formation. The ability of the mutant T7 RNA polymerases to bind template has not been greatly impaired. We have shown that amino acids D537 and D812 are essential, that amino acids K631 and H811 play significant roles in catalysis, and that the active site of T7 RNA polymerase is composed of different regions of the polypeptide chain. Possible roles for these catalytically significant residues in the polymerase mechanism are discussed.  相似文献   

16.
17.
Point mutations in proteins can have different effects on protein stability depending on the mechanism of unfolding. In the most interesting case of I27, the Ig‐like module of the muscle protein titin, one point mutation (Y9P) yields opposite effects on protein stability during denaturant‐induced “global unfolding” versus “vectorial unfolding” by mechanical pulling force or cellular unfolding systems. Here, we assessed the reason for the different effects of the Y9P mutation of I27 on the overall molecular stability and N‐terminal unraveling by NMR. We found that the Y9P mutation causes a conformational change that is transmitted through β‐sheet structures to reach the central hydrophobic core in the interior and alters its accessibility to bulk solvent, which leads to destabilization of the hydrophobic core. On the other hand, the Y9P mutation causes a bend in the backbone structure, which leads to the formation of a more stable N‐terminal structure probably through enhanced hydrophobic interactions.  相似文献   

18.
In a previous study, we showed that a synthetic human insulin 1‐chain analog, named analog ( 3 ) was capable of mimicking in vitro effects of native insulin, including stimulation of cell proliferation, glucose uptake and glycogen synthesis. Here, we have synthesized three new analogs ( 6, 9, 12 ) of the human A‐chain, bearing or not their N‐ or C‐terminal residue, to determine the structural features which are responsible for their biological properties. In vitro experiments clearly demonstrated that the N‐terminal part of the peptides is required for the biological activity of the molecules, suggesting its crucial role in the mechanism underlying the cellular effect. Our findings may help to better understand the mechanism of interaction between insulin and its receptor. In addition, the present data demonstrate that some mini‐insulin derived from the A‐chain can exert similar effects as native insulin. These small peptides may offer specific advantages over insulin in the definition of new strategies for diabetes treatment. Copyright © 2009 John Wiley & Sons, Ltd. This article was published online on 17 July 2009. An error was subsequently identified. This notice is included in the online and print versions to indicate that both have been corrected 4 August 2009.  相似文献   

19.
The extrinsic proteins of photosystem II of higher plants and green algae PsbO, PsbP, PsbQ, and PsbR are essential for stable oxygen production in the oxygen evolving center. In the available X‐ray crystallographic structure of higher plant PsbQ residues S14‐Y33 are missing. Building on the backbone NMR assignment of PsbQ, which includes this “missing link”, we report the extended resonance assignment including side chain atoms. Based on nuclear Overhauser effect spectra a high resolution solution structure of PsbQ with a backbone RMSD of 0.81 Å was obtained from torsion angle dynamics. Within the N‐terminal residues 1–45 the solution structure deviates significantly from the X‐ray crystallographic one, while the four‐helix bundle core found previously is confirmed. A short α‐helix is observed in the solution structure at the location where a β‐strand had been proposed in the earlier crystallographic study. NMR relaxation data and unrestrained molecular dynamics simulations corroborate that the N‐terminal region behaves as a flexible tail with a persistent short local helical secondary structure, while no indications of forming a β‐strand are found. Proteins 2015; 83:1677–1686. © 2015 The Authors. Proteins: Structure, Function, and Bioinformatics Published by Wiley Periodicals, Inc.  相似文献   

20.
Pinak Chakrabarti 《Proteins》2015,83(4):696-710
Dystrophin is a long, rod‐shaped cytoskeleton protein implicated in muscular dystrophy (MDys). Utrophin is the closest autosomal homolog of dystrophin. Both proteins have N‐terminal actin‐binding domain (N‐ABD), a central rod domain and C‐terminal region. N‐ABD, composed of two calponin homology (CH) subdomains joined by a helical linker, harbors a few disease causing missense mutations. Although the two proteins share considerable homology (>72%) in N‐ABD, recent structural and biochemical studies have shown that there are significant differences (including stability, mode of actin‐binding) and their functions are not completely interchangeable. In this investigation, we have used extensive molecular dynamics simulations to understand the differences and the similarities of these two proteins, along with another actin‐binding protein, fimbrin. In silico mutations were performed to identify two key residues that might be responsible for the dynamical difference between the molecules. Simulation points to the inherent flexibility of the linker region, which adapts different conformations in the wild type dystrophin. Mutations T220V and G130D in dystrophin constrain the flexibility of the central helical region, while in the two known disease‐causing mutants, K18N and L54R, the helicity of the region is compromised. Phylogenetic tree and sequence analysis revealed that dystrophin and utrophin genes have probably originated from the same ancestor. The investigation would provide insight into the functional diversity of two closely related proteins and fimbrin, and contribute to our understanding of the mechanism of MDys. Proteins 2015; 83:696–710. © 2015 Wiley Periodicals, Inc.  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