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 31 毫秒
1.
Tear lipocalin and β‐lactoglobulin are members of the lipocalin superfamily. They have similar tertiary structures but unusually low overall sequence similarity. Non‐native helical structures are formed during the early stage of β‐lactoglobulin folding. To address whether the non‐native helix formation is found in the folding of other lipocalin superfamily proteins, the folding kinetics of a tear lipocalin variant were investigated by stopped‐flow methods measuring the time‐dependent changes in circular dichroism (CD) spectrum and small‐angle X‐ray scattering (SAXS). CD spectrum showed that extensive secondary structures are not formed during a burst‐phase (within a measurement dead time). The SAXS data showed that the radius of gyration becomes much smaller than in the unfolded state during the burst‐phase, indicating that the molecule is collapsed during an early stage of folding. Therefore, non‐native helix formation is not general for folding of all lipocalin family members. The non‐native helix content in the burst‐phase folding appears to depend on helical propensities of the amino acid sequence. Proteins 2009. © 2008 Wiley‐Liss, Inc.  相似文献   

2.
β‐Lactoglobulin has been shown to interact with carotenoids from sea buckthorn berries. However, previously, no studies have taken into account the effect of calcium and magnesium on the β‐lactoglobulin–carotenoids complex. This study aims to determine the effect of calcium and magnesium on the interaction between β‐lactoglobulin and carotenoids from sea buckthorn berries extract, during heating from the perspective of deepening interaction mechanisms as prerequisites for micro‐ and nanoencapsulation. Phase diagram, intrinsic fluorescence spectra, quenching experiments and synchronous spectra were employed to acquire information regarding the conformation of protein in the presence of calcium chloride and magnesium chloride. Intrinsic fluorescence data showed that, between 25°C and 60°C, the presence of calcium chloride in the complex favoured the movement of tryptophan residues to domains located at the protein–water interface, while magnesium chloride favoured the burial of tryptophan residues. Higher temperatures generated blue shifts regardless of which salt was present, suggesting exposure of tryptophan residues to the hydrophobic core of the protein. Extrinsic fluorescence intensity of the non‐heat‐treated complex with magnesium chloride was significantly higher (P < 0.01) than of the complex with calcium chloride, suggesting that 1‐anilino‐8‐naphtalenesulphonic acid was bound to a higher proportion of the β‐lactoglobulin–carotenoids complex. Calcium chloride increased extrinsic fluorescence to a greater extent than magnesium chloride at temperatures above 70°C and was related to small structural changes induced by preheating β‐lactoglobulin.  相似文献   

3.
The crystal structures of six different fibronectin Type III consensus‐derived Tencon domains, whose solution properties exhibit no, to various degrees of, aggregation according to SEC, have been determined. The structures of the five variants showing aggregation reveal 3D domain swapped dimers. In all five cases, the swapping involves the C‐terminal β‐strand resulting in the formation of Tencon dimers in which the target‐binding surface is blocked. All of the variants differ in sequence in the FG loop, which is the hinge loop in the β‐strand‐swapped dimers. The six tencon variants have between 0 and 5 residues inserted between positions 77 and 78 in the FG loop. Analysis of the structures suggests that a non‐glycine residue at position 77 and insertions of <4 residues may destabilize the β‐turn in the FG loop promoting β‐strand swapping. Swapped dimers with an odd number of inserted residues may be less stable, particularly if they contain proline residues, because they cannot form perfect β‐bridges in the FG regions that link the swapped dimers. The Tencon β‐swapped variants with the longest FG sequences are observed to form higher order hexameric or helical oligomeric structures in the crystal correlating well with the aggregation properties of these domains observed in solution. Understanding the structural basis for domain‐swapped dimerization and oligomerization will support engineering efforts of the Tencon domain to produce variants with desired biophysical properties. Proteins 2014; 82:1359–1369. © 2013 Wiley Periodicals, Inc.  相似文献   

4.
Sorafenib tosylate (SORt) is an oral multikinase inhibitor used for treatment of advanced renal cell, liver, and thyroid cancers. In this study, this drug was synthesized and its antiproliferative activities against HCT116 and CT26 cells were assessed. The interaction of SORt with β‐lactoglobulin (BLG) was studied using different fluorescence techniques, circular dichroism (CD), zeta potential measurements, and docking simulation. The results of infrared (IR), mass, HNMR, and CNMR spectra demonstrated that the drug was produced with high quality, purity, and efficiency. SORt showed potent cytotoxicity against HCT116 and CT26 cells with IC50 of 8.12 and 5.42 μM, respectively. For BLG binding of SORt, the results showed that static quenching was the cause of the high affinity drug–protein interaction. Three‐dimensional fluorescence and synchronous spectra indicated that SORt conformation was changed at different levels. CD suggested that the α‐helix content remained almost constant in the BLG–SORt complex, whereas random coil content decreased. Zeta potential values of BLG were more positive after binding with SORt, due to electrostatic interactions between BLG and SORt. Thermodynamic parameters confirmed van der Waals and hydrogen bond interactions in the complex formation. Molecular modelling predicted the presence of hydrogen bonds and electrostatic forces in the BLG–SORt system, which was consistent with the experimental results.  相似文献   

5.
Martiniano Bello 《Biopolymers》2014,101(10):1010-1018
The bovine dairy protein β‐lactoglobulin (βlg) is a promiscuous protein that has the ability to bind several hydrophobic ligands. In this study, based on known experimental data, the dynamic interaction mechanism between bovine βlg and four fatty acids was investigated by a protocol combining molecular dynamics (MD) simulations and molecular mechanics generalized Born surface area (MMGBSA) binding free energy calculations. Energetic analyses revealed binding free energy trends that corroborated known experimental findings; larger ligand size corresponded to greater binding affinity. Finally, binding free energy decomposition provided detailed information about the key residues stabilizing the complex. © 2014 Wiley Periodicals, Inc. Biopolymers 101: 1010–1018, 2014.  相似文献   

6.
Endo‐β‐1,4‐d ‐mannanase from the Antarctic springtail, Cryptopygus antarcticus (CaMan), is a cold‐adapted β‐mannanase that has the lowest optimum temperature (30°C) of all known β‐mannanases. Here, we report the apo‐ and mannopentaose (M5) complex structures of CaMan. Structural comparison of CaMan with other β‐mannanases from the multicellular animals reveals that CaMan has an extended loop that alters topography of the active site. Structural and mutational analyses suggest that this extended loop is linked to the cold‐adapted enzymatic activity. From the CaMan‐M5 complex structure, we defined the mannose‐recognition subsites and observed unreported M5 binding site on the surface of CaMan. Proteins 2014; 82:3217–3223. © 2014 Wiley Periodicals, Inc.  相似文献   

7.
The structural properties of a 10‐residue and a 15‐residue peptide in aqueous solution were investigated by molecular dynamics simulation. The two designed peptides, SYINSDGTWT and SESYINSDGTWTVTE, had been studied previously by NMR at 278 K and the resulting model structures were classified as 3:5 β‐hairpins with a type I + G1 β‐bulge turn. In simulations at 278 K, starting from the NMR model structure, the 3:5 β‐hairpin conformers proved to be stable over the time period evaluated (30 ns). Starting from an extended conformation, simulations of the decapeptide at 278 K, 323 K and 353 K were also performed to study folding. Over the relatively short time scales explored (30 ns at 278 K and 323 K, 56 ns at 353 K), folding to the 3:5 β‐hairpin could only be observed at 353 K. At this temperature, the collapse to β‐hairpin‐like conformations is very fast. The conformational space accessible to the peptide is entirely dominated by loop structures with different degrees of β‐hairpin character. The transitions between different types of ordered loops and β‐hairpins occur through two unstructured loop conformations stabilized by a single side‐chain interaction between Tyr2 and Trp9, which facilitates the changes of the hydrogen‐bond register. In agreement with previous experimental results, β‐hairpin formation is initially driven by the bending propensity of the turn segment. Nevertheless, the fine organization of the turn region appears to be a late event in the folding process. Copyright © 2004 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

8.
The packed‐bed adsorption and elution of aqueous solutions of whey concentrate powders were investigated at pH 3.7 using a 5‐mL SP Sepharose FF column to separate and isolate two major proteins namely, α‐lactalbumin (ALA) and β‐lactoglobulin (BLG) from these solutions. ALA displaced and eluted BLG from the column in a pure form. Pure ALA could then be eluted with good recovery. A novel consecutive two‐stage separation process was developed to separate ALA and BLG from whey concentrate mixtures. Almost all of the BLG in the feed was recovered, with 78% being recovered at 95% purity and a further 20% at 86% purity. In addition, 67% of ALA was recovered, 48% at 54% purity and 19% at 60% purity. © 2009 American Institute of Chemical Engineers Biotechnol. Prog., 2010  相似文献   

9.
Structural changes ensuing from the non‐covalent absorption of bovine beta‐lactoglobulin (BLG) on the surface of polystyrene nanoparticles were investigated by using spectroscopic approaches, by assessing the reactivity of specific residues, and by limited proteolysis/mass spectrometry. Also, the immunoreactivity of absorbed and free BLG was compared. All these approaches indicated substantial rearrangements of the protein structure in the absorbed state, in spite of the reported structural rigidity of BLG. Changes made evident by experimental measurements were confirmed by computational approaches. These indicate that adsorption‐related changes are most marked in the area between the main C‐terminal alpha helix and the beta‐barrel, and lead to full exposure of the thiol on Cys121, consistent with experimental measurements. In the computational model of bound BLG, both Trp61 and Trp19 also move away from their neighboring quenchers and become solvent‐exposed, as indicated by fluorescence measurement. Upon binding, the beta‐barrel also loosens, with a substantial increase in immunoreactivity and with noticeable changes in the trypsinolytic pattern. The possible general significance of the structural changes reported here for non‐covalently adsorbed BLG is discussed with respect to recognition events involving surface‐bound proteins, as are aspects related to the carrier function(s) of BLG, and to its use as a common ingredient in many food systems. Proteins 2014; 82:1272–1282. © 2013 Wiley Periodicals, Inc.  相似文献   

10.
Deposition of insoluble fibrillar aggregates of β‐amyloid (Aβ) peptides in the brain is a hallmark of Alzheimer's disease. Apart from forming fibrils, these peptides also exist as soluble aggregates. Fibrillar and a variety of nonfibrillar aggregates of Aβ have also been obtained in vitro. Hexafluoroisopropanol (HFIP) has been widely used to dissolve Aβ and other amyloidogenic peptides. In this study, we show that the dissolution of Aβ40, 42, and 43 in HFIP followed by drying results in highly ordered aggregates. Although α‐helical conformation is observed, it is not stable for prolonged periods. Drying after prolonged incubation of Aβ40, 42, and 43 peptides in HFIP leads to structural transition from α‐helical to β‐conformation. The peptides form short fibrous aggregates that further assemble giving rise to highly ordered ring‐like structures. Aβ16–22, a highly amyloidogenic peptide stretch from Aβ, also formed very similar rings when dissolved in HFIP and dried. HFIP could not induce α‐helical conformation in Aβ16–22, and rings were obtained from freshly dissolved peptide. The rings formed by Aβ40, 42, 43, and Aβ16–22 are composed of the peptides in β‐conformation and cause enhancement in thioflavin T fluorescence, suggesting that the molecular architecture of these structures is amyloid‐like. Our results clearly indicate that dissolution of Aβ40, 42 and 43 and the amyloidogenic fragment Aβ16–22 in HFIP results in the formation of annular amyloid‐like structures. Copyright © 2012 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

11.
NagZ is an N‐acetyl‐β‐d ‐glucosaminidase that participates in the peptidoglycan (PG) recycling pathway of Gram‐negative bacteria by removing N‐acetyl‐glucosamine (GlcNAc) from PG fragments that have been excised from the cell wall during growth. The 1,6‐anhydromuramoyl‐peptide products generated by NagZ activate β‐lactam resistance in many Gram‐negative bacteria by inducing the expression of AmpC β‐lactamase. Blocking NagZ activity can thereby suppress β‐lactam antibiotic resistance in these bacteria. The NagZ active site is dynamic and it accommodates distortion of the glycan substrate during catalysis using a mobile catalytic loop that carries a histidine residue which serves as the active site general acid/base catalyst. Here, we show that flexibility of this catalytic loop also accommodates structural differences in small molecule inhibitors of NagZ, which could be exploited to improve inhibitor specificity. X‐ray structures of NagZ bound to the potent yet non‐selective N‐acetyl‐β‐glucosaminidase inhibitor PUGNAc (O‐(2‐acetamido‐2‐deoxy‐d ‐glucopyranosylidene) amino‐N‐phenylcarbamate), and two NagZ‐selective inhibitors – EtBuPUG, a PUGNAc derivative bearing a 2‐N‐ethylbutyryl group, and MM‐156, a 3‐N‐butyryl trihydroxyazepane, revealed that the phenylcarbamate moiety of PUGNAc and EtBuPUG completely displaces the catalytic loop from the NagZ active site to yield a catalytically incompetent form of the enzyme. In contrast, the catalytic loop was found positioned in the catalytically active conformation within the NagZ active site when bound to MM‐156, which lacks the phenylcarbamate extension. Displacement of the catalytic loop by PUGNAc and its N‐acyl derivative EtBuPUG alters the active site conformation of NagZ, which presents an additional strategy to improve the potency and specificity of NagZ inhibitors.  相似文献   

12.
The lipocalin β‐lactoglobulin (β‐LG) exists in different natural genetic variants—of which β‐LG A and B are predominant in bovine milk. At physiological conditions the protein dimerizes—building homodimers of β‐LG A and β‐LG B and heterodimers of β‐LG AB. Although β‐LG is one of the most intensely characterized lipocalins, the interaction behavior of ligands with hetero‐ and homodimers of β‐LG is largely unknown. The present findings revealed significant differences for hetero‐ and homodimers regarding ligand binding capacity as tested with a model ligand (i.e. surface binding (?)‐epigallocatechin gallate (EGCG)). These findings were confirmed using FT‐IR, where the addition of EGCG influenced the β‐sheet backbone of homodimer A and B with significantly higher intensity compared to heterodimer AB. Further, shape analysis by SAXS revealed oligomerization of both types of dimers upon addition of EGCG; however, homodimer A and B produced significantly larger aggregates compared to the heterodimer AB. In summary, the present study revealed that EGCG showed significantly different interaction reactivity (binding sites, aggregation size and conformational changes) to the hetero and homodimers of β‐LG in the order β‐LG A > B > AB. The results suggest that conformational differences between homodimers and heterodimers strongly influence the EGCG binding ability. This may also occur with other polyphenols and ligands of β‐LG and gives not only important information for β‐LG binding studies, but may also apply for polymorphisms of other self‐aggregating lipocalins. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

13.
The increasing interest in click chemistry and its use to stabilize turn structures led us to compare the propensity for β‐turn stabilization of different analogs designed as mimics of the β‐turn structure found in tendamistat. The β‐turn conformation of linear β‐amino acid‐containing peptides and triazole‐cyclized analogs were compared to ‘conventional’ lactam‐ and disulfide‐bridged hexapeptide analogs. Their 3D structures and their propensity to fold in β‐turns in solution, and for those not structured in solution in the presence of α‐amylase, were analyzed by NMR spectroscopy and by restrained molecular dynamics with energy minimization. The linear tetrapeptide Ac‐Ser‐Trp‐Arg‐Tyr‐NH2 and both the amide bond‐cyclized, c[Pro‐Ser‐Trp‐Arg‐Tyr‐D ‐Ala] and the disulfide‐bridged, Ac‐c[Cys‐Ser‐Trp‐Arg‐Tyr‐Cys]‐NH2 hexapeptides adopt dominantly in solution a β‐turn conformation closely related to the one observed in tendamistat. On the contrary, the β‐amino acid‐containing peptides such as Ac‐(R)‐β3‐hSer‐(S)‐Trp‐(S)‐β3‐hArg‐(S)‐β3‐hTyr‐NH2, and the triazole cyclic peptide, c[Lys‐Ser‐Trp‐Arg‐Tyr‐βtA]‐NH2, both specifically designed to mimic this β‐turn, do not adopt stable structures in solution and do not show any characteristics of β‐turn conformation. However, these unstructured peptides specifically interact in the active site of α‐amylase, as shown by TrNOESY and saturation transfer difference NMR experiments performed in the presence of the enzyme, and are displaced by acarbose, a specific α‐amylase inhibitor. Thus, in contrast to amide‐cyclized or disulfide‐bridged hexapeptides, β‐amino acid‐containing peptides and click‐cyclized peptides may not be regarded as β‐turn stabilizers, but can be considered as potential β‐turn inducers. Copyright © 2011 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

14.
Unlike most small globular proteins, lipocalins lack a compact hydrophobic core. Instead, they present a large central cavity that functions as the primary binding site for hydrophobic molecules. Not surprisingly, these proteins typically exhibit complex structural dynamics in solution, which is intricately modified by intermolecular recognition events. Although many lipocalins are monomeric, an increasing number of them have been proven to form oligomers. The coupling effects between self‐association and ligand binding in these proteins are largely unknown. To address this issue, we have calorimetrically characterized the recognition of dodecyl sulfate by bovine β‐lactoglobulin, which forms weak homodimers at neutral pH. A thermodynamic analysis based on coupled‐equilibria revealed that dimerization exerts disparate effects on the ligand‐binding capacity of β‐lactoglobulin. Protein dimerization decreases ligand affinity (or, reciprocally, ligand binding promotes dimer dissociation). The two subunits in the dimer exhibit a positive, entropically driven cooperativity. To investigate the structural determinants of the interaction, the crystal structure of β‐lactoglobulin bound to dodecyl sulfate was solved at 1.64 Å resolution. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

15.
Although the thermodynamic principles that control the binding of drug molecules to their protein targets are well understood, the detailed process of how a ligand reaches a protein binding site has been an intriguing question over decades. The short time interval between the encounter between a ligand and its receptor to the formation of the stable complex has prevented experimental observations. Bovine β‐lactoglobulin (βlg) is a lipocalin member that carries fatty acids (FAs) and other lipids in the cellular environment. Βlg accommodates a FA molecule in its highly hydrophobic cavity and exhibits the capability of recognizing a wide variety of hydrophobic ligands. To elucidate the ligand entry process on βlg, we report molecular dynamics simulations of the encounter between palmitate (PA) or laurate (LA) and βlg. Our results show that residues localized in loops at the cavity entrance play an important role in the ligand penetration process. Analysis of the short‐term interaction energies show that the forces operating on the systems lead to average conformations very close to the crystallographic holo‐forms. Whereas the binding free energy analysis using the molecular mechanics Generalized Born surface area method shows that these conformations were thermodynamically favorable. © 2013 Wiley Periodicals, Inc. Biopolymers 101: 744–757, 2014.  相似文献   

16.
Protein loops make up a large portion of the secondary structure in nature. But very little is known concerning loop closure dynamics and the effects of loop composition on fold stability. We have designed a small system with stable β‐sheet structures, including features that allow us to probe these questions. Using paired Trp residues that form aromatic clusters on folding, we are able to stabilize two β‐strands connected by varying loop lengths and composition (an example sequence: R W ITVTI – loop – KKIRV W E). Using NMR and CD, both fold stability and folding dynamics can be investigated for these systems. With the 16 residue loop peptide (sequence: R W ITVTI‐(GGGGKK)2GGGG‐KKIRV W E) remaining folded (ΔGU = 1.6 kJ/mol at 295K). To increase stability and extend the series to longer loops, we added an additional Trp/Trp pair in the loop flanking position. With this addition to the strands, the 16 residue loop (sequence: R W ITVRI W ‐(GGGGKK)2GGGG‐ W KTIRV W E) supports a remarkably stable β‐sheet (ΔGU = 6.3 kJ/mol at 295 K, Tm = ~55°C). Given the abundance of loops in binding motifs and between secondary structures, these constructs can be powerful tools for peptide chemists to study loop effects; with the Trp/Trp pair providing spectroscopic probes for assessing both stability and dynamics by NMR.  相似文献   

17.
The small heat shock protein (sHSP) from Methanococcus jannaschii (Mj Hsp16.5) forms a monodisperse 24mer and each of its monomer contains two flexible N‐ and C‐terminals and a rigid α‐crystallin domain with an extruding β‐strand exchange loop. The minimal α‐crystallin domain with a β‐sandwich fold is conserved in sHSP family, while the presence of the β‐strand exchange loop is divergent. The function of the β‐strand exchange loop and the minimal α‐crystallin domain of Mj Hsp16.5 need further study. In the present study, we constructed two fragment‐deletion mutants of Mj Hsp16.5, one with both the N‐ and C‐terminals deleted (ΔNΔC) and the other with a further deletion of the β‐strand exchange loop (ΔNΔLΔC). ΔNΔC existed as a dimer in solution. In contrast, the minimal α‐crystallin domain ΔNΔLΔC became polydisperse in solution and exhibited more efficient chaperone‐like activities to prevent amorphous aggregation of insulin B chain and fibril formation of the amyloidogenic peptide dansyl‐SSTSAA‐W than the mutant ΔNΔC and the wild type did. The hydrophobic probe binding experiments indicated that ΔNΔLΔC exposed much more hydrophobic surface than ΔNΔC. Our study also demonstrated that Mj Hsp16.5 used different mechanisms for protecting different substrates. Though Mj Hsp16.5 formed stable complexes with substrates when preventing thermal aggregation, no complexes were detected when preventing aggregation under non‐heat‐shock conditions. Proteins 2014; 82:1156–1167. © 2013 Wiley Periodicals, Inc.  相似文献   

18.
The conformation of oligomers of β‐amino acids of the general type Ac‐[β‐Xaa]n‐NHMe (β‐Xaa = β‐Ala, β‐Aib, and β‐Abu; n = 1–4) was systematically examined at different levels of ab initio molecular orbital theory (HF/6‐31G*, HF/3‐21G). The solvent influence was considered employing two quantum‐mechanical self‐consistent reaction field models. The results show a wide variety of possibilities for the formation of characteristic elements of secondary structure in β‐peptides. Most of them can be derived from the monomer units of blocked β‐peptides with n = 1. The stability and geometries of the β‐peptide structures are considerably influenced by the side‐chain positions, by the configurations at the Cα‐ and Cβ‐atoms of the β‐amino acid constituents, and especially by environmental effects. Structure peculiarities of β‐peptides, in particular those of various helix alternatives, are discussed in relation to typical elements of secondary structure in α‐peptides. © 1999 John Wiley & Sons, Inc. Biopoly 50: 167–184, 1999  相似文献   

19.
Secondary structural transitions from α‐helix to β‐sheet conformations are observed in several misfolding diseases including Alzheimer's and Parkinson's. Determining factors contributing favorably to the formation of each of these secondary structures is therefore essential to better understand these disease states. β‐hairpin peptides form basic components of anti‐parallel β‐sheets and are suitable model systems for characterizing the fundamental forces stabilizing β‐sheets in fibrillar structures. In this study, we explore the free energy landscape of the model β‐hairpin peptide GB1 and its E2 isoform that preferentially adopts α‐helical conformations at ambient conditions. Umbrella sampling simulations using all‐atom models and explicit solvent are performed over a large range of end‐to‐end distances. Our results show the strong preference of GB1 and the E2 isoform for β‐hairpin and α‐helical conformations, respectively, consistent with previous studies. We show that the unfolded states of GB1 are largely populated by misfolded β‐hairpin structures which differ from each other in the position of the β‐turn. We discuss the energetic factors contributing favorably to the formation of α‐helix and β‐hairpin conformations in these peptides and highlight the energetic role of hydrogen bonds and non‐bonded interactions. Proteins 2014; 82:2394–2402. © 2014 Wiley Periodicals, Inc.  相似文献   

20.
Clostridium propionicum is the only organism known to ferment β‐alanine, a constituent of coenzyme A (CoA) and the phosphopantetheinyl prosthetic group of holo‐acyl carrier protein. The first step in the fermentation is a CoA‐transfer to β‐alanine. Subsequently, the resulting β‐alanyl‐CoA is deaminated by the enzyme β‐alanyl‐CoA:ammonia lyase (Acl) to reversibly form ammonia and acrylyl‐CoA. We have determined the crystal structure of Acl in its apo‐form at a resolution of 0.97 Å as well as in complex with CoA at a resolution of 1.59 Å. The structures reveal that the enyzme belongs to a superfamily of proteins exhibiting a so called “hot dog fold” which is characterized by a five‐stranded antiparallel β‐sheet with a long α‐helix packed against it. The functional unit of all “hot dog fold” proteins is a homodimer containing two equivalent substrate binding sites which are established by the dimer interface. In the case of Acl, three functional dimers combine to a homohexamer strongly resembling the homohexamer formed by YciA‐like acyl‐CoA thioesterases. Here, we propose an enzymatic mechanism based on the crystal structure of the Acl·CoA complex and molecular docking. Proteins 2014; 82:2041–2053. © 2014 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