Rescue of protein splicing activity from a Magnetospirillum magnetotacticum intein-like element

2004 ◽  
Vol 32 (2) ◽  
pp. 250-254 ◽  
Author(s):  
M.W. Southworth ◽  
J. Yin ◽  
F.B. Perler

The self-catalytic protein splicing mechanism is mediated by the intein plus the first amino acid following the intein C-terminus (termed the +1 residue). Although polymorphisms of conserved residues elsewhere in inteins have been widely reported, no splicing-competent intein has been observed without a Ser, Thr or Cys in this functionally essential +1 position. This residue is the nucleophile in two steps of the protein splicing pathway: ligation of the extein fragments during transesterification and formation of a peptide bond between the exteins by an acyl rearrangement. An intein-like element in a hypothetical protein (gene Magn8951) from Magnetospirillum magnetotacticum has all intein signature sequences except the +1 residue, where it has a Tyr. Although the Tyr side-chain hydroxyl can potentially mediate the transesterification reaction, an acyl shift has never been observed with this residue. When the activities of this bacterial intein-like element were studied, protein splicing was not observed and N-terminal cleavage predominated. Mutation of Tyr+1 to Phe or Ala indicated that the Tyr side-chain hydroxyl was not necessary for N-terminal cleavage. Protein splicing activity could be rescued by ‘reversion’ of Tyr+1 to Cys.

Catalysts ◽  
2021 ◽  
Vol 11 (8) ◽  
pp. 926
Author(s):  
Maria C. Martins ◽  
Susana F. Fernandes ◽  
Bruno A. Salgueiro ◽  
Jéssica C. Soares ◽  
Célia V. Romão ◽  
...  

Flavodiiron proteins (FDPs) are a family of modular and soluble enzymes endowed with nitric oxide and/or oxygen reductase activities, producing N2O or H2O, respectively. The FDP from Escherichia coli, which, apart from the two core domains, possesses a rubredoxin-like domain at the C-terminus (therefore named flavorubredoxin (FlRd)), is a bona fide NO reductase, exhibiting O2 reducing activity that is approximately ten times lower than that for NO. Among the flavorubredoxins, there is a strictly conserved amino acids motif, -G[S,T]SYN-, close to the catalytic diiron center. To assess its role in FlRd’s activity, we designed several site-directed mutants, replacing the conserved residues with hydrophobic or anionic ones. The mutants, which maintained the general characteristics of the wild type enzyme, including cofactor content and integrity of the diiron center, revealed a decrease of their oxygen reductase activity, while the NO reductase activity—specifically, its physiological function—was almost completely abolished in some of the mutants. Molecular modeling of the mutant proteins pointed to subtle changes in the predicted structures that resulted in the reduction of the hydration of the regions around the conserved residues, as well as in the elimination of hydrogen bonds, which may affect proton transfer and/or product release.


1988 ◽  
Vol 53 (11) ◽  
pp. 2810-2824 ◽  
Author(s):  
Ilmars Sekacis ◽  
Mark Shenderovich ◽  
Gregory Nikiforovich ◽  
Edvards Liepinš ◽  
Ludmila Polevaya ◽  
...  

A group of synthetic peptides including Boc-Lys-Phe-X-Y, X = Ala (I, III) or Thr (II), Y = Pro (I, II) or Ala (III) was studied by means of 1H NMR spectroscopy and theoretical conformational analysis. Compound I in DMSO shows two conformers with the trans- and cis-configuration of the peptide bond Ala-Pro. The salt bridge between the Lys ε-amino group and the C-terminal carboxyl is featured by magnetic nonequivalence of the Lys CεH2 protons. The space structure of I and II was found to possess a salt bridge fixed by an unusual turn in the chain formed by the Lys side chain and the C-terminal dipeptide with the trans-peptide bond X-Pro. Since a stable ionic bond in III and in the cis-conformer of I has not been observed, its contribution to stabilization of the space structure of the peptides in DMSO appears rather small.


1988 ◽  
Vol 66 (11) ◽  
pp. 2733-2750 ◽  
Author(s):  
Saul Wolfe ◽  
Kiyull Yang ◽  
Maged Khalil

Using the MMPEN parameters of Allinger's MMP2(85) force field, a conformational analysis has been performed on four biologically active penicillins; D-ampicillin, L-α-phenoxyethylpenicillin, penicillin G, and penicillin V, and on five biologically inactive or much less active penicillins: L-ampicillin, D-α-phenoxyethylpenicillin, N-methylpenicillin G, 6α-methylpenicillin G, and bisnorpenicillin G. Antibacterial activity is found to be associated with the existence of a global minimum having a compact structure, whose convex face is accessible to a penicillin binding protein (PBP), with the C3-carboxyl group and the side-chain N-H exposed on this face. Using the MMPEP parameters of MMP2(85), a conformational analysis has been performed on phenylacetyl-D-Ala-D-Ala-O−, a peptide model of the normal substrate of a PBP. Labischinski's global minimum has been reproduced, along with structures that correspond to Tipper and Strominger's proposal that the N4—C7 bond of a penicillin corresponds to the Ala–Ala peptide bond, and to Hasan's proposal that the N4—C5 bond of penicillin corresponds to the peptide bond. For both models, conformations of the peptide related to the pseudoaxial and pseudoequatorial conformations of the thiazolidine ring of penicillin G have been examined. It is concluded that penicillin is not a structural analog of the global minimum of the peptide; however, comparisons based on unbound conformations of PBP substrates are unable to determine which model is more appropriate, or which conformation of penicillin G is the biologically significant one. Using the ECEPP/MMPEP strategy, a model of the active site of a PBP has been obtained, following a search of 200,000 structures of the peptide Ac-NH-Val-Gly-Ser-Val-Thr-Lys-NH-Me. This peptide contains the sequence at the active site of a PBP of Streptomyces R61, for which it is also known that the C3-carboxyl group of penicillin binds to the ε-amino group of lysine, and the β-lactam reacts chemically with the serine OH. The lysine and serine side chains and the C-terminal carbonyl group are found to occupy the concave face of the active site model.A strategy for the docking of penicillins or peptides to this model, with full minimization of the conformational energies of the complexes, has been devised. All active penicillins bind through strong hydrogen bonds to the C3-carboxyl group and the side-chain N-H, and with a four-centered relationship between the O-H of serine and the (O)C-N of the β-lactam ring. The geometrical parameters of this relationship are reminiscent of those found in the gas phase transition state of neutral hydration of a carbonyl group. When the energies of formation and geometries of the pseudoaxial and pseudoequatorial penicillin G complexes are examined, there is now a clear preference for the binding of the pseudoaxial conformation, which is the global minimum of the uncomplexed penicillin in this case. A similar examination of the peptide complexes reveals that only the conformation of the peptide that corresponds to Tipper and Strominger's model, and is based on the pseudoaxial conformation of penicillin G, can form a complex with a geometry and energy comparable to those of a biologically active penicillin.


2021 ◽  
Vol 18 ◽  
Author(s):  
Sarah Kappler ◽  
Andreas Siebert ◽  
Uli Kazmaier

Introduction: Miuraenamides belong to marine natural compounds with interesting biological properties. Materials and Methods: They initiate polymerization of monomeric actin and therefore show high cytotoxicity by influencing the cytoskeleton. New derivatives of the miuraenamides have been synthesized containing a N-methylated amide bond instead of the more easily hydrolysable ester in the natural products. Results: Incorporation of an aromatic side chain onto the C-terminal amino acid of the tripeptide fragment also led to highly active new miuraenamides. Conclusion: We could show that the ester bond of the natural product miuraenamide can be replaced by an N-methyl amide. The yields in the cyclization step are high and generally much better that with the corresponding esters. On the other hand, the biological activity of the new amide analogs are lower compared to the natural products, but the activity can significantly be increased by incorporation of a p-nitrophenyl group at the C-terminus of the peptide fragment.


RSC Advances ◽  
2018 ◽  
Vol 8 (20) ◽  
pp. 11134-11144 ◽  
Author(s):  
Lanyan He ◽  
Pingmei Wang ◽  
Lipeng He ◽  
Zhou Qu ◽  
Jianhui Luo ◽  
...  

The self-organization of five model side-chain decorated polyaromatic asphaltene molecules with or without toluene solvent was investigated by means of molecular dynamic (MD) simulations.


Amino Acids ◽  
2020 ◽  
Vol 52 (10) ◽  
pp. 1425-1438
Author(s):  
Johann Sajapin ◽  
Michael Hellwig

Abstract Oxidative stress, an excess of reactive oxygen species (ROS), may lead to oxidative post-translational modifications of proteins resulting in the cleavage of the peptide backbone, known as α-amidation, and formation of fragments such as peptide amides and α-ketoacyl peptides (α-KaP). In this study, we first compared different approaches for the synthesis of different model α-KaP and then investigated their stability compared to the corresponding unmodified peptides. The stability of peptides was studied at room temperature or at temperatures relevant for food processing (100 °C for cooking and 150 °C as a simulation of roasting) in water, in 1% (m/v) acetic acid or as the dry substance (to simulate the thermal treatment of dehydration processes) by HPLC analysis. Oxidation of peptides by 2,5-di-tert-butyl-1,4-benzoquinone (DTBBQ) proved to be the most suited method for synthesis of α-KaPs. The acyl side chain of the carbonyl-terminal α-keto acid has a crucial impact on the stability of α-KaPs. This carbonyl group has a catalytic effect on the hydrolysis of the neighboring peptide bond, leading to the release of α-keto acids. Unmodified peptides were significantly more stable than the corresponding α-KaPs. The possibility of further degradation reactions was shown by the formation of Schiff bases from glyoxylic or pyruvic acids with glycine and proven through detection of transamination products and Strecker aldehydes of α-keto acids by HPLC–MS/MS. We propose here a mechanism for the decomposition of α-ketoacyl peptides.


2007 ◽  
Vol 40 (26) ◽  
pp. 9398-9405 ◽  
Author(s):  
M. Knaapila ◽  
F. B. Dias ◽  
V. M. Garamus ◽  
L. Almásy ◽  
M. Torkkeli ◽  
...  

ChemistryOpen ◽  
2017 ◽  
Vol 6 (2) ◽  
pp. 266-272 ◽  
Author(s):  
Jurgen Schill ◽  
Lech-Gustav Milroy ◽  
Jody A. M. Lugger ◽  
Albertus P. H. J. Schenning ◽  
Luc Brunsveld

Author(s):  
Jennifer Hochscherf ◽  
Markus Pietsch ◽  
William Tieu ◽  
Kevin Kuan ◽  
Andrew D. Abell ◽  
...  

Glycosylated human leukocyte elastase (HLE) was crystallized and structurally analysed in complex with a 1,3-thiazolidine-2,4-dione derivative that had been identified as an HLE inhibitor in preliminary studies. In contrast to previously described HLE structures with small-molecule inhibitors, in this structure the inhibitor does not bind to the S1 and S2 substrate-recognition sites; rather, this is the first HLE structure with a synthetic inhibitor in which the S2′ site is blocked that normally binds the second side chain at the C-terminal side of the scissile peptide bond in a substrate protein. The inhibitor also induces the formation of crystalline HLE dimers that block access to the active sites and that are also predicted to be stable in solution. Neither such HLE dimers nor the corresponding crystal packing have been observed in previous HLE crystal structures. This novel crystalline environment contributes to the observation that comparatively large parts of the N-glycan chains of HLE are defined by electron density. The final HLE structure contains the largest structurally defined carbohydrate trees among currently available HLE structures.


2003 ◽  
Vol 278 (40) ◽  
pp. 39133-39142 ◽  
Author(s):  
Yi Ding ◽  
Ming-Qun Xu ◽  
Inca Ghosh ◽  
Xuehui Chen ◽  
Sebastien Ferrandon ◽  
...  

Sign in / Sign up

Export Citation Format

Share Document