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.
Mimicry of structural motifs is a common feature in proteins. The 10‐membered hydrogen‐bonded ring involving the main‐chain C?O in a β‐turn can be formed using a side‐chain carbonyl group leading to Asx‐turn. We show that the N? H component of hydrogen bond can be replaced by a Cγ‐H group in the side chain, culminating in a nonconventional C? H···O interaction. Because of its shape this β‐turn mimic is designated as ω‐turn, which is found to occur ~three times per 100 residues. Three residues (i to i + 2) constitute the turn with the C? H···O interaction occurring between the terminal residues, constraining the torsion angles ?i + 1, ψi + 1, ?i + 2 and χ1(i + 2) (using the interacting Cγ atom). Based on these angles there are two types of ω‐turns, each of which can be further divided into two groups. Cβ‐branched side‐chains, and Met and Gln have high propensities to occur at i + 2; for the last two residues the carbonyl oxygen may participate in an additional interaction involving the S and amino group, respectively. With Cys occupying the i + 1 position, such turns are found in the metal‐binding sites. N‐linked glycosylation occurs at the consensus pattern Asn‐Xaa‐Ser/Thr; with Thr at i + 2, the sequence can adopt the secondary structure of a ω‐turn, which may be the recognition site for protein modification. Location between two β‐strands is the most common occurrence in protein tertiary structure, and being generally exposed ω‐turn may constitute the antigenic determinant site. It is a stable scaffold and may be used in protein engineering and peptide design. Proteins 2015; 83:203–214. © 2014 Wiley Periodicals, Inc.  相似文献   

2.
Carbon-oxygen (CH···O) hydrogen bonding represents an unusual category of molecular interactions first documented in biological structures over 4 decades ago. Although CH···O hydrogen bonding has remained generally underappreciated in the biochemical literature, studies over the last 15 years have begun to yield direct evidence of these interactions in biological systems. In this minireview, we provide a historical context of biological CH···O hydrogen bonding and summarize some major advancements from experimental studies over the past several years that have elucidated the importance, prevalence, and functions of these interactions. In particular, we examine the impact of CH···O bonds on protein and nucleic acid structure, molecular recognition, and enzyme catalysis and conclude by exploring overarching themes and unresolved questions regarding unconventional interactions in biomolecular structure.  相似文献   

3.
In this study are reported the syntheses of three bis(diarylhydrazonecarbonyl)methylene derivatives [{ArPhCNNH C(O)}2CH2] [Ar = 2 C5H4N (5), C6H5 (6), and 2‐C4H3S (7)], obtained by condensation of corresponding hydrazones with carbon suboxide, C3O2. The solid‐state self‐assembly of these carbonyl derivatives, giving rise to polymeric and dimeric networks, is described. In the formation of these structural features, in addition to N—H· · ·OC intermolecular hydrogen bonds, stabilizing intramolecular NH· · · π (systems) and intermolecular CO· · ·π (systems) interactions also seem to play an important role. Solution 1H‐nmr data of compounds 5–7 indicate that the polymeric and dimeric structures are not maintained in solution and show the occurrence of keto‐enolic equilibria. © 1999 John Wiley & Sons, Inc. Biopoly 49: 541–549, 1999  相似文献   

4.
A hydrogen bond prediction tool HBPredicT is developed for detecting structural water molecules and CH···π interactions in PDB files of protein-ligand complexes. The program adds the missing hydrogen atoms to the protein, ligands, and oxygen atoms of water molecules and subsequently all the hydrogen bonds in the complex are located using specific geometrical criteria. Hydrogen bonds are classified into various types based on (i) donor and acceptor atoms, and interactions such as (ii) protein-protein, (iii) protein-ligand, (iv) protein-water, (v) ligand-water, (vi) water-water, and (vii) protein-water-ligand. Using the information in category (vii), the water molecules which form hydrogen bonds with the ligand and the protein simultaneously–the structural water–is identified and retrieved along with the associated ligand and protein residues. For CH···π interactions, the relevant portions of the corresponding structures are also extracted in the output. The application potential of this program is tested using 19 HIV-1 protease and 11 PTP1B inhibitor complexes. All the systems showed presence of structural water molecules and in several cases, the CH···π interaction between ligand and protein are detected. A rare occurrence of CH···π interactions emanating from both faces of a phenyl ring of the inhibitor is identified in HIV-1 protease 1D4L.  相似文献   

5.
The high‐capacity cathode material V2O5·n H2O has attracted considerable attention for metal ion batteries due to the multielectron redox reaction during electrochemical processes. It has an expanded layer structure, which can host large ions or multivalent ions. However, structural instability and poor electronic and ionic conductivities greatly handicap its application. Here, in cell tests, self‐assembly V2O5·n H2O nanoflakes shows excellent electrochemical performance with either monovalent or multivalent cation intercalation. They are directly grown on a 3D conductive stainless steel mesh substrate via a simple and green hydrothermal method. Well‐layered nanoflakes are obtained after heat treatment at 300 °C (V2O5·0.3H2O). Nanoflakes with ultrathin flower petals deliver a stable capacity of 250 mA h g?1 in a Li‐ion cell, 110 mA h g?1 in a Na‐ion cell, and 80 mA h g?1 in an Al‐ion cell in their respective potential ranges (2.0–4.0 V for Li and Na‐ion batteries and 0.1–2.5 V for Al‐ion battery) after 100 cycles.  相似文献   

6.
In this study we have described the non-canonical interactions between the porphyrin ring and the protein part of porphyrin-containing proteins to better understand their stabilizing role. The analysis reported in this study shows that the predominant type of non-canonical interactions at porphyrins are CH···O and CH···N interactions, with a small percentage of CH···π and non-canonical interactions involving sulfur atoms. The majority of non-canonical interactions are formed from side-chains of charged and polar amino acids, whereas backbone groups are not frequently involved. The main-chain non-canonical interactions might be slightly more linear than the side-chain interactions, and they have somewhat shorter median distances. The analysis, performed in this study, shows that about 44% of the total interactions in the dataset are involved in the formation of multiple (furcated) non-canonical interactions. The high number of porphyrin–water interactions show importance of the inclusion of solvent in protein–ligand interaction studies. Furthermore, in the present study we have observed that stabilization centers are composed predominantly from nonpolar amino acid residues. Amino acids deployed in the environment of porphyrin rings are deposited in helices and coils. The results from this study might be used for structure-based porphyrin protein prediction and as scaffolds for future porphyrin-containing protein design.  相似文献   

7.
Hydrogen bonds are important interaction forces observed in protein structures. They can be classified as stronger or weaker depending on their energy, thereby reflecting on the type of donor. The contribution of weak hydrogen bonds is deemed as an important factor toward structure stability along with the stronger bonds. One such bond, the C‐H…O type hydrogen bond, is shown to make a contribution in maintaining three dimensional structures of proteins. Apart from their presence within protein structures, the role of these bonds in protein–ligand interactions is also noteworthy. In this study, we present a statistical analysis on the presence of C‐H…O hydrogen bonds observed between FKBPs and their cognate ligands. The FK506‐binding proteins (FKBPs) carry peptidyl cis–trans isomerase activity apart from the immunosuppressive property by binding to the immunosuppressive drugs FK506 or rapamycin. Because the active site of FKBPs is lined up by many hydrophobic residues, we speculated that the prevalence of C‐H…O hydrogen bonds will be considerable. In a total of 25 structures analyzed, a higher frequency of C‐H…O hydrogen bonds is observed in comparison with the stronger hydrogen bonds. These C‐H…O hydrogen bonds are dominated by a highly conserved donor, the Cα/β of Val55 and an acceptor, the backbone oxygen of Glu54. Both these residues are positioned in the β4‐α1 loop, whereas the other residues Tyr26, Phe36 and Phe99 with higher frequencies are lined up at the opposite face of the active site. These preferences could be implicated in FKBP pharmacophore models toward enhancing the ligand affinity. This study could be a prelude to studying other proteins with hydrophobic pockets to gain better insights into ligand recognition. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

8.
A terminally protected acyclic tetrapeptide has been synthesized, and the crystal structure of its hydrated form, Boc‐Tyr‐Aib‐Tyr‐Ile‐OMe·2H2O ( 1 ), has been determined directly from powder X‐ray diffraction data. The backbone conformation of tetrapeptide ( 1 ) exhibiting two consecutive β‐turns is stabilized by two 4 → 1 intramolecular N―H · · · O hydrogen bonds. In the crystalline state, the tetrapeptide molecules are assembled through water‐mediated O―H · · · O hydrogen bonds to form two‐dimensional molecular sheets, which are further linked by intermolecular C―H · · · O hydrogen bonds into a three‐dimensional supramolecular framework. The molecular electrostatic potential (MEP) surface of ( 1 ) has been used to supplement the crystallographic observations. The nature of intermolecular interactions in ( 1 ) has been analyzed quantitatively through the Hirshfeld surface and two‐dimensional fingerprint plot. The DFT optimized molecular geometry of ( 1 ) agrees closely with that obtained from the X‐ray structure analysis. The present structure analysis of Boc‐Tyr‐Aib‐Tyr‐Ile‐OMe·2H2O ( 1 ) represents a case where ab‐initio crystal structure of an acyclic tetrapeptide with considerable molecular flexibility has been accomplished from laboratory X‐ray powder diffraction data. Copyright © 2015 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

9.
Sulfur‐containing odorants and flavors play an important role in flavor and food industry, especially when meaty, garlic, onion, and vegetable scents are needed. Still, many S‐containing flavors also possess fruity scents and may be used in compositions of perfumes that require a fresh and fruity odor perception. They are naturally abundant in various fruits, essential oils, and food. Most of these compounds possess strong scents, and their scent composition is highly dependent on the concentration applied. At higher concentrations, they usually feature repellent off‐odors, while their sweet and fruity nature is only experienced at very low concentrations. This represents a challenge for their application in perfumery, as well as in food industry. From a molecular point of view, the endless structural and scent variety of these compounds makes them fascinating, especially as their O‐analogs are usually free of any malodors. Here, we report on the investigation of the gas‐phase structure and dynamics of the S‐containing blackcurrant odorant cat ketone (4‐methyl‐4‐sulfanylpentan‐2‐one). The work was performed by combining quantum‐chemical calculations and molecular‐beam Fourier‐transform microwave spectroscopy as complementary tools. Using this technique, we are able to determine the structures of sizeable molecules where energy differences are small and conformational distinction is not always possible by quantum‐chemical calculations alone. The presented results can be used for further structure? activity correlation studies, as well as for benchmarks to improve theoretical models, especially, as there is still significant interest in characterizing the various conformers of organic molecules in terms of relative energies, structures, and dipole moments.  相似文献   

10.
Tetraalloxygermanium(IV), (CH2·CH·CH2·O)4Ge, has been synthesized from germanium tetrachloride, allyl alcohol, and ammonia. The alloxides [(CH2·CH· CH2·O)4Ti]2and[(CH2·CH·CH2O)5M]2 (M = Nb and Ta) have been synthesized by reactions of the corresponding metal isopropoxides with allyl alcohol followed by removal of the isopropanol by azeotropic distillation with benzene. These four metal alloxides can be purified by distillation under reduced pressure. The spectroscopic properties of these new compounds are discussed.  相似文献   

11.
The structure of Peptide T was determined by solution NMR spectroscopy, under strong structure‐inducing conditions: 40% hexafluoro‐2‐propanol aqueous solution at 5 °C. Under these conditions it was possible to detect medium‐range NOEs for the first time for this peptide. This allowed a much better‐defined structure to be determined for Peptide T in comparison with earlier NMR and computational studies. Peptide structures consistent with the experimental restraints were generated using a restrained MD simulation with a full empirical force field. Residues 4–8 of Peptide T take on a well‐defined structure with a heavy atom RMSD of 0.78 Å. The structure is stabilized by hydrogen bonding to side‐chain oxygen atoms of Thr 4 and Thr 8, as well as backbone hydrogen bonding between residues 5 and 7 that forms this region into a classic γ‐turn. Copyright © 2009 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

12.
Ethanol is one of the most commonly used solvents to extract flavonoids from propolis. Hydrogen bonding interactions play an important role in the properties of liquid system. The main objective of the work is to study the hydrogen bonding interactions between flavonoid and ethanol. Luteolin is a very common flavonoid that has been found in different geographical and botanical propolis. In this work, it was selected as the representative flavonoid to do detailed research. The study was performed from a theoretical perspective using density functional theory (DFT) method. After careful optimization, there exist nine optimized geometries for the luteolin ? CH3CH2OH complex. The binding distance of X ? H···O, and the bond length, vibrational frequency, and electron density changes of X ? H all indicate the formation of the hydrogen bond in the optimized geometries. In the optimized geometries, it is found that: (1) except for the H2’, H5’, and H6’, CH3CH2OH has formed hydrogen bonds with all the hydrogen and oxygen atoms in luteolin. The hydrogen atoms in the hydroxyl groups of luteolin form the strongest hydrogen bonds with CH3CH2OH; (2) all of the hydrogen bonds are closed-shell interactions; (3) the strongest hydrogen bond is the O3’ ? H3’···O in structure A, while the weakest one is the C3 ? H3···O in structure E; (4) the hydrogen bonds of O3’ ? H3’···O, O ? H···O4, O ? H···O3’ and O ? H···O7 are medium strength and covalent dominant in nature. While the other hydrogen bonds are weak strength and possess a dominant character of the electrostatic interactions in nature.  相似文献   

13.
We report on an unrestrained molecular dynamics simulation of the flavin mononucleotide (FMN)–RNA aptamer. The simulated average structure maintains both cross‐strand and intermolecular FMN–RNA nuclear Overhauser effects from the nmr experiments and has all qualitative features of the nmr structure including the G10–U12–A25 base triple and the A13–G24, A8–G28, and G9–G27 mismatches. However, the relative orientation of the hairpin loop to the remaining part of the molecule differs from the nmr structure. The simulation predicts that the flexible phosphoglycerol part of FMN moves toward G27 and forms hydrogen bonds. There are structurally long‐lived water molecules in the FMN binding pocket forming hydrogen bonds within FMN and between FMN and RNA. In addition, long‐lived water is found bridging primarily RNA backbone atoms. A general feature of the environment of long‐lived “structural” water is at least two and in most cases three or four potential acceptor atoms. The 2′‐OH group of RNA usually acts as an acceptor in interactions with the solvent. There are almost no intrastrand O2′H(n)⋮O4′(n + 1) hydrogen bonds within the RNA backbone. In the standard case the preferred orientation of the 2′‐OH hydrogen atoms is approximately toward O3′ of the same nucleotide. However, a relatively large number of conformations with the backbone torsional angle γ in the trans orientation is found. A survey of all experimental RNA x‐ray structures shows that this backbone conformation occurs but is less frequent than found in the simulation. Experimental nmr RNA aptamer structures have a higher fraction of this conformation as compared to the x‐ray structures. The backbone conformation of nucleotide n + 1 with the torsional angle γ in the trans orientation leads to a relatively short distance between 2′‐OH(n) and O5′(n + 1), enabling hydrogen‐bond formation. In this case the preferred orientation of the 2′‐OH hydrogen atom is approximately toward O5′(n + 1). We find two relatively short and dynamically stable types of backbone–backbone next‐neighbor contacts, namely C2′(H)(n)⋮O4′(n + 1) and C5′(H)(n + 1)⋮O2′(n). These interactions may affect both backbone rigidity and thermodynamic stability of RNA helical structures. © 1999 John Wiley & Sons, Inc. Biopoly 50: 287–302, 1999  相似文献   

14.
15.
Common structural elements in proteins such as α‐helices or β‐sheets are characterized by uniformly repeating, energetically favorable main chain conformations which additionally exhibit a completely saturated hydrogen‐bonding network of the main chain NH and CO groups. Although polyproline or polyglycine type II helices (PPII or PGII) are frequently found in proteins, they are not considered as equivalent secondary structure elements because they do not form a similar self‐contained hydrogen‐bonding network of the main chain atoms. In this context our finding of an unusual motif of glycine‐rich PGII‐like helices in the structure of the acetophenone carboxylase core complex is of relevance. These PGII‐like helices form hexagonal bundles which appear to fulfill the criterion of a (largely) saturated hydrogen‐bonding network of the main‐chain groups and therefore may be regarded in this sense as a new secondary structure element. It consists of a central PGII‐like helix surrounded by six nearly parallel PGII‐like helices in a hexagonal array, plus an additional PGII‐like helix extending the array outwards. Very related structural elements have previously been found in synthetic polyglycine fibers. In both cases, all main chain NH and CO groups of the central PGII‐helix are saturated by either intra‐ or intermolecular hydrogen‐bonds, resulting in a self‐contained hydrogen‐bonding network. Similar, but incomplete PGII‐helix patterns were also previously identified in a GTP‐binding protein and an antifreeze protein.  相似文献   

16.
Backbone mimicry by the formation of closed-loop C7, C10 and C13 (mimics of gamma-, beta- and alpha-turns) conformations through side chain-main chain hydrogen bonds by polar groups is a frequent observation in protein structures. A data set of 250 non-homologous and high-resolution protein crystal structures was used to analyze these conformations for their characteristic features. Seven out of the nine polar residues (Ser, Thr, Asn, Asp, Gln, Glu and His) have hydrogen bonding groups in their side chains which can participate in such mimicry and as many as 15% of all these polar residues engage in such conformations. The distributions of dihedral angles of these mimics indicate that only certain combinations of the dihedral angles involved aid the formation of these mimics. The observed examples were categorized into various classes based on these combinations, resulting in well defined motifs. Asn and Asp residues show a very high capability to perform such backbone secondary structural mimicry. The most highly mimicked backbone structure is of the C10 conformation by the Asx residues. The mimics formed by His, Ser, Thr and Glx residues are also discussed. The role of such conformations in initiating the formation of regular secondary structures during the course of protein folding seems significant.  相似文献   

17.
This review, including some new experimental results, is the summary of a talk at the ‘flavors & fragrances 2013’ conference in Leipzig, organized jointly by the GDCh, the Liebig‐Vereinigung, and the EuCheMS. After times of searching for natural odor principles and serendipitous discoveries by chemical inspiration, directed odorant design today offers the highest hit rates for the discovery of new odorants, although serendipity still plays a role. Keeping intact the electronic shape required for a certain olfactophore‐binding geometry, one can add or subtract structural elements, rigidify molecular structures, or introduce more structural flexibility. To find out which structural features are critical, the ‘seco‐approach’, in which different fragments are removed by cutting strategic bonds, is the most analytical. Following this approach, such ingredients as Serenolide, Sylkolide, and Pomarose were designed. Transferring this design principle from the family of damascones to that of the theaspiranes led to the discovery of Cassyrane, though completely different structural features turned out to be relevant. Application of the seco‐concept to a 3,7a‐substituted 2,6,7,7a‐tetrahydro‐1H‐inden‐5‐yl musk lead structure derived from carotol resulted in the discovery of a new family of dienone musks with novel structure? odor correlations. However, cutting the C(2)? O bond of Cassyrane and oxidizing the resulting secostructure to the 1,2,5,1″‐tetradehydro derivative links the family of dienone musks with that of blackcurrant odorants, but the resulting target structures turned out to be potent orris odorants. (3E,5E)‐5‐(tert‐Butyl)octadeca‐3,5‐dien‐2‐one even possesses the lowest odor threshold in the whole ionone family (0.036 ng/l air), which could be rationalized by a superposition analysis on (?)‐cisγ‐irone. In the course of the synthesis of these high‐impact orris odorants, we discovered that, depending on the reaction conditions, the dehydration step of the intermediate 5‐hydroxyalk‐3‐yn‐2‐ones was accompanied by a carbenium‐ion rearrangement. Depending on the substitution pattern, these rearrangement products and their derivatives possessed interesting musky‐woody olfactory properties reminiscent of Cashmeran, demonstrating that the same structural elements can code for completely different odors, i.e., cassis, musk, orris, violet, and Cashmeran‐type, depending only on their spatial arrangement.  相似文献   

18.
Reliable computational prediction of protein side chain conformations and the energetic impact of amino acid mutations are the key aspects for the optimization of biotechnologically relevant enzymatic reactions using structure‐based design. By improving the protein stability, higher yields can be achieved. In addition, tuning the substrate selectivity of an enzymatic reaction by directed mutagenesis can lead to higher turnover rates. This work presents a novel approach to predict the conformation of a side chain mutation along with the energetic effect on the protein structure. The HYDE scoring concept applied here describes the molecular interactions primarily by evaluating the effect of dehydration and hydrogen bonding on molecular structures in aqueous solution. Here, we evaluate its capability of side‐chain conformation prediction in classic remutation experiments. Furthermore, we present a new data set for evaluating “cross‐mutations,” a new experiment that resembles real‐world application scenarios more closely. This data set consists of protein pairs with up to five point mutations. Thus, structural changes are attributed to point mutations only. In the cross‐mutation experiment, the original protein structure is mutated with the aim to predict the structure of the side chain as in the paired mutated structure. The comparison of side chain conformation prediction (“remutation”) showed that the performance of HYDEprotein is qualitatively comparable to state‐of‐the art methods. The ability of HYDEprotein to predict the energetic effect of a mutation is evaluated in the third experiment. Herein, the effect on protein stability is predicted correctly in 70% of the evaluated cases. Proteins 2017; 85:1550–1566. © 2017 Wiley Periodicals, Inc.  相似文献   

19.
Transthyretin (TTR) is a tetrameric protein associated with human amyloidosis. In vitro, the formation of amyloid fibrils by TTR is known to be promoted by low pH. Here we show the neutron structure of TTR, focusing on the hydrogen bonds, protonation states and pH sensitivities. A large crystal was prepared at pD 7.4 for neutron protein crystallography. Neutron diffraction studies were conducted using the IBARAKI Biological Crystal Diffractometer with the time-of-flight method. The neutron structure solved at 2.0 Å resolution revealed the protonation states of His88 and the detailed hydrogen-bond network depending on the protonation states of His88. This hydrogen-bond network is composed of Thr75, Trp79, His88, Ser112, Pro113, Thr118-B and four water molecules, and is involved in both monomer–monomer and dimer–dimer interactions, suggesting that the double protonation of His88 by acidification breaks the hydrogen-bond network and causes the destabilization of the TTR tetramer. In addition, the comparison with X-ray structure at pH 4.0 indicated that the protonation occurred to Asp74, His88 and Glu89 at pH 4.0. Our neutron model provides insights into the molecular stability of TTR related to the hydrogen-bond network, the pH sensitivity and the CH···O weak hydrogen bond.  相似文献   

20.
Three series of novel urushiol derivatives were designed by introducing a hydroxamic acid moiety into the tail of an alkyl side chain and substituents with differing electronic properties or steric bulk onto the benzene ring and alkyl side chain. The compounds’ binding affinity toward HDAC8 was screened by Glide docking. The highest-scoring compounds were processed further with molecular docking, MD simulations, and binding free energy studies to analyze the binding modes and mechanisms. Ten compounds had Glide scores of ?8.2 to ?10.2, which revealed that introducing hydroxy, carbonyl, amino, or methyl ether groups into the alkyl side chain or addition of –F, –Cl, sulfonamide, benzamido, amino, or hydroxy substituents on the benzene ring could significantly increase binding affinity. Molecular docking studies revealed that zinc ion coordination, hydrogen bonding, and hydrophobic interactions contributed to the high calculated binding affinities of these compounds toward HDAC8. MD simulations and binding free energy studies showed that all complexes possessed good stability, as characterized by low RMSDs, low RMSFs of residues, moderate hydrogen bonding and zinc ion coordination and low values of binding free energies. Hie147, Tyr121, Phe175, Hip110, Phe119, Tyr273, Lys21, Gly118, Gln230, Leu122, Gly269, and Gly107 contributed favorably to the binding; and Van der Waals and electrostatic interactions provided major contributions to the stability of these complexes. These results show the potential of urushiol derivatives as HDAC8 binding lead compounds, which have great therapeutic potential in the treatment of various malignancies, neurological disorders, and human parasitic diseases.  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