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.
Aspartame (L-aspartyl L-phenylalanine methyl ester) isan artificial sweetener as shown in Fig.1 (A) [1]. Studieson its structure and function showed that its N-terminalL-aspartyl residue could only be replaced by aminomalonyl[2] or L-asparaginyl [3] residue. When its peptide bondwas replaced by an ester bond [Fig. 1(B)] or the hydrogenof amide in the peptide bond replaced by a methyl group[Fig. 1(C)], its sweetness was lost [4]. According to thecrystal structure of aspartame, between the …  相似文献   

2.
Hydrazine, hydrazone and hydrazide derivatives are nitrogen–nitrogen bond containing compounds. Such molecules are relatively scarce in nature and have been isolated from plants, marine organisms and microorganisms. These compounds exhibit remarkable structural diversity and relevant biological activities. The enzymes involved in the formation of the N–N bond are still unknown, but many lines of evidence support the involvement of N-nitrosation and N-hydroxylation activating steps. Beside the challenging N–N bond, N-acylases catalyzing the C–N bond formation contribute to the chemical diversity of N–N-containing natural products (N2NP). This review examines the state of knowledge regarding the biosynthesis of N2NP, for which only two biosynthetic gene clusters have been investigated. Biological properties and chemical synthesis of hydrazines, hydrazones and hydrazides are also reported.  相似文献   

3.
We have performed quantum chemical calculations for the MCCBr???NCH and HCCBr???NCM' (M, M'?=?Cu, Ag, and Au) halogen-bonded complexes at the MP2 level. The results showed that the transition metals have different influences on the halogen bond donor and the electron donor. The transition metal atom in the former makes the halogen bond weaker, and that in the latter causes it to enhance. Molecular electrostatic potential and natural bond orbital analysis were carried out to reveal the nature of the substitution.  相似文献   

4.
W W Cleland 《FASEB journal》1990,4(11):2899-2905
Secondary 18O isotope effects have been developed as a tool for determining transition state structures in enzymatic and nonenzymatic phosphoryl transfer reactions. 18O substitution in the nonbridge oxygens of a phosphoryl group makes the reaction go faster when the bond order is higher to these oxygens in the transition state than in the reactant, whereas the reaction goes slower if the bond order is less. The isotope effects are measured by the remote label method, using an isotope ratio mass spectrometer for analysis. The bond order to p-nitrophenolate ion when it is the leaving group is indicated by the secondary 15N isotope effect in the nitro group, with a value of 1.0028 representing nearly complete bond cleavage. It appears that the transition states for phosphoryl transfer have no more than one negative charge on the nonbridge oxygens, so that reactions of monoesters are dissociative, reactions of triesters are associative, and reactions of diesters are SN2 with half bond order to entering and leaving groups.  相似文献   

5.
The bond energy (BE) of a polyatomic molecule cannot be measured and, therefore, determination of BEs can only be done within a model using a set of assumptions. The bond strength is reflected by the intrinsic BE (IBE), which is related to the intrinsic atomization energy (IAE) and which represents the energy of dissociation under the provision that the degree of hybridization is maintained for all atoms of the molecule. IBE and BE differ in the case of CC and CH bonds by the promotion, the hybridization, and the charge reorganization energy of carbon. Since the latter terms differ from molecule to molecule, IBE and BE are not necessarily parallel and the use of BEs from thermochemical models can be misleading. The stretching force constant is a dynamical quantity and, therefore, it is related to the bond dissociation energy (BDE). Calculation and interpretation of stretching force constants for local internal coordinate modes are discussed and it is demonstrated that the best relationship between BDEs and stretching force constants is obtained within the model of adiabatic internal modes. The valence stretching force constants are less suitable since they are related to an artificial bond dissociation process with geometrical relaxation effects suppressed, which leads to an intrinsic BDE (IBDE). In the case of AXn molecules, symmetric coordinates can be used to get an appropriate stretching force constant that is related to the BE. However, in general stretching force constants determined for symmetry coordinates do not reflect the strength of a particular bond since the related dissociation processes are strongly influenced by the stability of the products formed.  相似文献   

6.
The formation of receptor ligand bonds at the interface between different cells and between cells and substrates is a widespread phenomenon in biological systems. Physical measurements of bond formation rates between cells and substrates have been exploited to increase our understanding of the biophysical mechanisms that regulate bond formation at interfaces. Heretofore, these measurements have been interpreted in terms of simple bimolecular reaction kinetics. Discrepancies between this simple framework and the behavior of neutrophils adhering to surfaces expressing vascular cell adhesion molecule 1 (VCAM-1) motivated the development of a new kinetic framework in which the explicit formation of active bond formation sites (reaction zones) are a prerequisite for bond formation to occur. Measurements of cells interacting with surfaces having a wide range of VCAM-1 concentrations, and for different durations of contact, enabled the determination of novel kinetic rate constants for the formation of reaction zones and for the intrinsic bond kinetics. Comparison of these rates with rates determined previously for other receptor-ligand pairs points to a predominant role of extrinsic factors such as surface topography and accessibility of active molecules to regions of close contact in determining forward rates of bond formation at cell interfaces.  相似文献   

7.
The disulfide bond structures established decades ago for immunoglobulins have been challenged by findings from extensive characterization of recombinant and human monoclonal IgG antibodies. Non-classical disulfide bond structure was first identified in IgG4 and later in IgG2 antibodies. Although, cysteine residues should be in the disulfide bonded states, free sulfhydryls have been detected in all subclasses of IgG antibodies. In addition, disulfide bonds are susceptible to chemical modifications, which can further generate structural variants such as IgG antibodies with trisulfide bond or thioether linkages. Trisulfide bond formation has also been observed for IgG of all subclasses. Degradation of disulfide bond through β-elimination generates free sulfhydryls disulfide and dehydroalanine. Further reaction between free sulfhydryl and dehydroalanine leads to the formation of a non-reducible cross-linked species. Hydrolysis of the dehydroalanine residue contributes substantially to antibody hinge region fragmentation. The effect of these disulfide bond variations on antibody structure, stability and biological function are discussed in this review.Key words: recombinant monoclonal antibody, disulfide bond, trisulfide bond, free sulfhydryl, dehydroalanine, thioether, aggregation  相似文献   

8.
Bis(ferrocenyl)-substituted allenylidene complexes, [(CO)5MCCCFc2] (1a-c, Fc = (C5H4)Fe(C5H5), M = Cr (a), Mo (b), W (c)) were obtained by sequential reaction of Fc2CO with Me3Si-CCH, KF/MeOH, n-BuLi, and [(CO)5M(THF)]. For the synthesis of related mono(ferrocenyl)allenylidene chromium complexes, [(CO)5CrCCC(Fc)R] (R = Ph, NMe2), three different routes were developed: (a) reaction of the deprotonated propargylic alcohol HCCC(Fc)(Ph)OH with [(CO)5Cr(THF)] followed by desoxygenation with Cl2CO, (b) Lewis acid induced alcohol elimination from alkenyl(alkoxy)carbene complexes, [(CO)5CrC(OR)CHC(NMe2)Fc], and (c) replacement of OMe in [(CO)5CrCCC(OMe)NMe2] by Fc. Complex 1a was also formed when the mono(ferrocenyl)allenylidene complex [(CO)5CrCCC(Fc)NMe2] was treated first with Li[Fc] and the resulting adduct then with SiO2. The replacement route (c) was also applied to the synthesis of an allenylidene complex (7a) with a CC spacer in between the ferrocenyl unit and Cγ of the allenylidene ligand, [(CO)5CrCCC(NMe2)-CCFc]. The related complex containing a CHCH spacer (9a) was prepared by condensation of [(CO)5CrCCC(Me)NMe2] with formylferrocene in the presence of NEt3. The bis(ferrocenyl)-substituted allenylidene complexes 1a-c added HNMe2 across the Cα-Cβ bond to give alkenyl(dimethylamino)carbene complexes and reacted with diethylaminopropyne by regioselective insertion of the CC bond into the Cβ-Cγ bond to afford alkenyl(diethylamino)allenylidene complexes, [(CO)5MCCC(NEt2)CMeCFc2]. The structures of 5a, 7a, and 9a were established by X-ray diffraction studies.  相似文献   

9.
The strain of Absidia coerulea was used to investigate the transformations of testosterone, androstenedione, progesterone and testosterone derivatives with additional C1–C2 double bond and/or 17-methyl group. All the examined substrates were transformed, mainly hydroxylated. It was found that the position and stereochemistry of the introduced hydroxyl group, as well as the yield of products, depended on the structure of the substrate. The first three substrates (hormones) underwent hydroxylation at C-14, and additional hydroxylation at 7 was observed in progesterone. The presence of the double bond (C1–C2) in 1-dehydrotestosterone did not influence the position of hydroxylation, but the product with additional C14–C15 double bond (at the same site as hydroxylation) was formed. 17-Methyltestosterone was hydroxylated at the 7 position, and also the dehydrogenated product (at the same site, with C6–C7 double bond) was obtained. The testosterone derivative with both C1–C2 double bond and 17-methyl group underwent hydroxylation at the 7 or 11β position, and a little amount of 14, 15 epoxide was formed.  相似文献   

10.
The X-ray crystal structures of the semicarbazide derivatives of D-glucose, D-galactose, and D-xylose are described. All are glycopyranosyl derivatives in the solid state. The glucose semicarbazide crystallizes as a dihydrate. Two different conformations of the galactose semicarbazide are found, one having the gg side-chain orientation unfavorable for monosaccharides having the 4C(1)-D-galacto configuration. In the two other known examples, this conformation is stabilized by an intramolecular hydrogen bond, but in the current structure, the hydrogen bonds involving the side chain are intermolecular. In these semicarbazides the N [bond] N [bond] C[double bond] O torsional angle is approximately 180 degrees, in contrast to the angles of 0.60 and 13.9 degrees in the two solid-state conformations of 2-benzoyl-1-(alpha-D-xylopyranosyl)hydrazine [Ernholt, B. V.; Thomsen, I. B.; Lohse, A.; Plesner, I. W.; Jensen, K. B.; Hazell, R. G.; Liang, X.; Jakobsen, A.; Bols, M. Chem. Eur. J. 2000, 6, 278-287]. The water molecules and the carbonyl oxygen are heavily involved in H-bonding.  相似文献   

11.
Stereoisomeric uridylyl(3',5')uridines D,L-UpU andL,D-UpU were synthesised. Their cleavage was followed in thepresence of acid, base and metal ion catalysts to studywhether the stereochemistry affects the inherent reactivity ofthe internucleosidic phosphodiester bond, and whether the lowmolecular weight catalysts can distinguish between thesubstrates. The rate constants obtained were compared to thoseof D,D-UpU. The comparison shows that the stability of thephosphodiester bond does not depend on the stereochemistry ofthe sugar rings. In contrast slight reactivity differences areobserved in the presence of metal ion catalysts, whichsuggests that selective cleavage of stereoisomeric substrateseven by small molecular weight chemical catalysts may bepossible.  相似文献   

12.
Barrick D 《Proteins》2000,39(4):291-308
The trans-substituted histidine to glycine mutant of sperm whale myoglobin (H93G Mb) is used to study energetics of proximal hydrogen bonding, proximal ligand-heme interactions, and coupling to distal ligand binding. Comparison of mono- and dimethylimidazole structural isomers shows that the hydrogen bond between the proximal ligand and the neighboring Ser92 hydroxyl (position F7) is stabilizing. The range of hydrogen bond stabilities measured here for different distal ligand complexes ranges from -0.7 kcal/mol (monomethylimidazole isomers to MbCO) to -4.1 kcal/mol (dimethylimidazole isomers to MbCN). This range of hydrogen bond stabilities, which is similar to that seen in protein mutagenesis unfolding studies, demonstrates the high sensitivity of the hydrogen bond to modest structural perturbations. The degree to which the 2-methyl group destabilizes proximal ligand binding is found to depend inversely on the total electronic spin. For monomethylimidazole proximal ligands, distal ligand binding weakens the proximal hydrogen bond compared to deoxyMb. Surprisingly, this trend is largely reversed for the dimethylimidazole proximal ligands. These results demonstrate strong coupling between the proximal protein matrix and distal ligand binding. These results provide an explanation for the strong avoidance of hydrogen bonding residues at position F7 in hemoglobin sequences.  相似文献   

13.
A distinguishing feature of camel (Camelus dromedarius) VHH domains are noncanonical disulfide bonds between CDR1 and CDR3. The disulfide bond may provide an evolutionary advantage, as one of the cysteines in the bond is germline encoded. It has been hypothesized that this additional disulfide bond may play a role in binding affinity by reducing the entropic penalty associated with immobilization of a long CDR3 loop upon antigen binding. To examine the role of a noncanonical disulfide bond on antigen binding and the biophysical properties of a VHH domain, we have used the VHH R303, which binds the Listeria virulence factor InlB as a model. Using site directed mutagenesis, we produced a double mutant of R303 (C33A/C102A) to remove the extra disulfide bond of the VHH R303. Antigen binding was not affected by loss of the disulfide bond, however the mutant VHH displayed reduced thermal stability (Tm = 12°C lower than wild‐type), and a loss of the ability to fold reversibly due to heat induced aggregation. X‐ray structures of the mutant alone and in complex with InlB showed no major changes in the structure. B‐factor analysis of the structures suggested that the loss of the disulfide bond elicited no major change on the flexibility of the CDR loops, and revealed no evidence of loop immobilization upon antigen binding. These results suggest that the noncanonical disulfide bond found in camel VHH may have evolved to stabilize the biophysical properties of the domain, rather than playing a significant role in antigen binding.  相似文献   

14.
K Hon-nami  T Oshima 《Biochemistry》1979,18(25):5693-5697
The denaturation of Thermus thermophilus cytochrome c-552 by acid, guanidine hydrochloride, and heat was studied by measuring the changes in absorption and circular dichroism. Cytochrome c-552 was remarkably resistant to acid; the pK of the transition from the low- to the high-spin form was roughly 0.3. The effect of guanidine hydrochloride on the heme iron-methionine bond of Thermus and horse cytochromes c was also investigated; a comparison of the free-energy changes for the displacement of the bond indicated that the coordination in cytochrome c-552 is highly stable. The spectra of guanidine hydrochloride unfolded cytochrome c-552 were dependent on the pH; the titration curve showed the presence of a cooperative single transition of pK = 4.7, with a one-proton dissociation, suggesting the ionization of a histidine residue. In the presence of guanidine hydrochloride, the influence of the heat on the ligand bond in cytochrome c-552 was studied. The van't Hoff plots of the reaction were biphasic. The enthalpy changes in the higher temperature range were independent on the guanidine hydrochloride concentration, while those in the lower range were not.  相似文献   

15.
Sato Y  Inaba K 《The FEBS journal》2012,279(13):2262-2271
Almost all organisms, from bacteria to humans, possess catalytic systems that promote disulfide bond formation-coupled protein folding, i.e. oxidative protein folding. These systems are necessary for the biosynthesis of many secretory and membrane proteins, such as antibodies, major histocompatibility complex molecules, growth factors, and insulin. Over the last decade, structural studies have made striking progress in this field of research, identifying how oxidative systems operate in a specific and regulated manner to maintain redox and protein homeostasis within cells. Interestingly, more and more novel catalysts that promote disulfide bond formation have been discovered in mammals, suggesting that the oxidative protein folding network is even more complicated in higher eukaryotes than previously thought. This review highlights the physiological roles and molecular bases of the disulfide bond formation pathways that have evolved in the bacterial periplasm and the endoplasmic reticulum of fungi and mammals. Accumulating knowledge about disulfide bond formation networks widely distributed throughout the biological kingdom has significantly advanced our understanding of the cellular mechanisms dedicated to protein quality control.  相似文献   

16.
The Dead-End Elimination method was used to identify 40 low energy microconformations of 16 tryptophan residues in eight proteins. Single Trp-mutants of these proteins all show a double- or triple-exponential fluorescence decay. For ten of these lifetimes the corresponding rotameric state could be identified by comparing the bimolecular acrylamide quenching constant (k(q)) and the relative solvent exposure of the side chain in that microstate. In the absence of any identifiable quencher, the origin of the lifetime heterogeneity is interpreted in terms of the electron transfer process from the indole C epsilon 3 atom to the carbonyl carbon of the peptide bond. Therefore it is expected that a shorter [C epsilon 3-C[double bond]O] distance leads to a shorter lifetime as observed for these ten rotamers. Applying the same rule to the other 30 lifetimes, a link with their corresponding rotameric state could also be made. In agreement with the theory of Marcus and Sutin, the nonradiative rate constant shows an exponential relationship with the [C epsilon 3-C[double bond]O] distance for the 40 datapoints.  相似文献   

17.
The nature of M-O bond in MOX4 compounds (where M = Ru or Os and X = F, Cl, Br or I) was analyzed by density functional theory methods at the BP86/LANL2DZ level of theory. The obtained charge density was analyzed by Fermi hole analysis, natural bond order (NBO) analysis and atoms-in-molecules (AIM)-based methods. The M-O bond is essentially a triple bond, although strongly polarized. The clearest differences in bonding between the Ru and Os compounds can be found in the M-O σ bonds, where in the Os compounds we find more charge density resting close to O.  相似文献   

18.
Heme oxygenase (HO) catalyzes heme degradation by utilizing O(2) and reducing equivalents to produce biliverdin IX alpha, iron, and CO. To avoid product inhibition, the heme[bond]HO complex (heme[bond]HO) is structured to markedly increase its affinity for O(2) while suppressing its affinity for CO. We determined the crystal structures of rat ferrous heme[bond]HO and heme[bond]HO bound to CO, CN(-), and NO at 2.3, 1.8, 2.0, and 1.7 A resolution, respectively. The heme pocket of ferrous heme-HO has the same conformation as that of the previously determined ferric form, but no ligand is visible on the distal side of the ferrous heme. Fe[bond]CO and Fe[bond]CN(-) are tilted, whereas the Fe[bond]NO is bent. The structure of heme[bond]HO bound to NO is identical to that bound to N(3)(-), which is also bent as in the case of O(2). Notably, in the CO- and CN(-)-bound forms, the heme and its ligands shift toward the alpha-meso carbon, and the distal F-helix shifts in the opposite direction. These shifts allow CO or CN(-) to bind in a tilted fashion without a collision between the distal ligand and Gly139 O and cause disruption of one salt bridge between the heme and basic residue. The structural identity of the ferrous and ferric states of heme[bond]HO indicates that these shifts are not produced on reduction of heme iron. Neither such conformational changes nor a heme shift occurs on NO or N(3)(-) binding. Heme[bond]HO therefore recognizes CO and O(2) by their binding geometries. The marked reduction in the ratio of affinities of CO to O(2) for heme[bond]HO achieved by an increase in O(2) affinity [Migita, C. T., Matera, K. M., Ikeda-Saito, M., Olson, J. S., Fujii, H., Yoshimura, T., Zhou, H., and Yoshida, T. (1998) J. Biol. Chem. 273, 945-949] is explained by hydrogen bonding and polar interactions that are favorable for O(2) binding, as well as by characteristic structural changes in the CO-bound form.  相似文献   

19.
  1. In US Pacific Northwest ponderosa pine forests the primary attraction order shown previously for red turpentine beetle, Dendroctonus valens (Coleoptera: Curculionidae: Scolytinae), is (−)-β-pinene+ethanol > (+)-3-carene+ethanol > (+)-α-pinene+ethanol. The monoterpenes are bicyclic C10H16 isomers containing one 6-carbon ring with one double bond. Both pinenes have a 4-carbon second ring and differ only by their endocyclic or exocyclic double bond. The (+)-3-carene second ring has 3-carbons; its double bond is endocyclic like (+)-α-pinene.
  2. Ring system and double bond influences on primary attraction were evaluated by hydrogenating (+)-3-carene and (+)-α-pinene to cis-carane and cis-pinane, respectively. Field test primary attraction strengths were (−)-β-pinene+ethanol > cis-carane+ethanol > cis-pinane+ethanol > ethanol.
  3. In combination with ethanol (i) a double bond is not required in either ring system to attract D. valens, (ii) the cis-carane bicyclic 3, 6-carbon ring system provides stronger beetle attraction than the cis-pinane 4, 6-carbon bicyclic ring system, and likely structural basis for stronger (+)-3-carene attraction over (+)-α-pinene, (iii) adding an exocyclic double bond to the 4, 6-carbon ring system elevates attraction above the 3, 6-carbon ring system with no double bond, and (iv) the 4, 6-carbon ring system is a much stronger attractant with an exocyclic rather than endocyclic double bond.
  相似文献   

20.
The effect of decreased protein flexibility on the stability and calcium binding properties of calbindin D9k has been addressed in studies of a disulfide bridged calbindin D9k mutant, denoted (L39C + P43M + I73C), with substitutions Leu 39-->Cys, Ile 73-->Cys, and Pro 43-->Met. Backbone 1H NMR assignments show that the disulfide bond, which forms spontaneously under air oxidation, is well accommodated. The disulfide is inserted on the opposite end of the protein molecule with respect to the calcium sites, to avoid direct interference with these sites, as confirmed by 113Cd NMR. The effect of the disulfide bond on calcium binding was assessed by titrations in the presence of a chromophoric chelator. A small but significant effect on the cooperativity was found, as well as a very modest reduction in calcium affinity. The disulfide bond increases Tm, the transition midpoint of thermal denaturation, of calcium free calbindin D9k from 85 to 95 degrees C and Cm, the urea concentration of half denaturation, from 5.3 to 8.0 M. Calbindins with one covalent bond linking the two EF-hand subdomains are equally stable regardless if the covalent link is the 43-44 peptide bond or the disulfide bond. Kinetic remixing experiments show that separated CNBr fragments of (L39C + P43M + I73C), each comprising one EF-hand, form disulfide linked homodimers. Each homodimer binds two calcium ions with positive co-operativity, and an average affinity of 10(6) M-1. Disulfide linkage dramatically increases the stability of each homodimer. For the homodimer of the C-terminal fragment Tm increases from 59 +/- 2 without covalent linkage to 91 +/- 2 degrees C with disulfide, and Cm from approximately 1.5 to 7.5 M. The overall topology of this homodimer is derived from 1H NMR assignments and a few key NOEs.  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