Arne O. Smalås
University of Tromsø
Network
Latest external collaboration on country level. Dive into details by clicking on the dots.
Publication
Featured researches published by Arne O. Smalås.
Comparative Biochemistry and Physiology B | 1996
Heidi Outzen; Gunnar I. Berglund; Arne O. Smalås; Nils Peder Willassen
Four differently charged trypsins were purified from pyloric caeca of Atlantic salmon (Salmo salar). The isoelectric points of three anionic isoforms were 4.70, 4.60, and 4.55 (anionic trypsin I, II and III, respectively). And for the first time a cationic isoform (isoelectric point above 9.3) has been isolated from a marine species. The apparent molecular weights of all four isoforms were about 25 kDa as determined by SDS-PAGE. The salmon enzymes were inhibited by serine proteinase inhibitors in general and also by specific trypsin inhibitors. Anionic trypsin I and the cationic isoform were further examined. Anionic trypsin I showed the typical cold-adaptation features, low pH and temperature stability (also lower Gibbs free energy of GdnHCl-induced unfolding) and high catalytic efficiency as compared to the mammalian trypsins. The cationic isoform did not show these features, but resembled the mammalian trypsins.
Protein Science | 2004
Hanna-Kirsti S. Leiros; Bjørn Olav Brandsdal; Ole Andreas Andersen; Ingar Leiros; Ronny Helland; Jacek Otlewski; Nils Peder Willassen; Arne O. Smalås
The variation in inhibitor specificity for five different amine inhibitors bound to CST, BT, and the cold‐adapted AST has been studied by use of association constant measurements, structural analysis of high‐resolution crystal structures, and the LIE method. Experimental data show that AST binds the 1BZA and 2BEA inhibitors 0.8 and 0.5 kcal/mole more strongly than BT. However, structural interactions and orientations of the inhibitors within the S1 site have been found to be virtually identical in the three enzymes studied. For example, the four water molecules in the inhibitor‐free structures of AST and BT are channeled into similar positions in the S1 site, and the nitrogen atom(s) of the inhibitors are found in two cationic binding sites denoted Position1 and Position2. The hydrophobic binding contributions for all five inhibitors, estimated by the LIE calculations, are also in the same order (−2.1 ± 0.2 kcal/mole) for all three enzymes. Our hypothesis is therefore that the observed variation in inhibitor binding arises from different electrostatic interactions originating from residues outside the S1 site. This is well illustrated by AST, in which Asp 150 and Glu 221B, despite some distance from the S1 binding site, lower the electrostatic potential of the S1 site and thus enhance substrate binding. Because the trends in the experimentally determined binding energies were reproduced by the LIE calculations after adding the contribution from long‐range interactions, we find this method very suitable for rational studies of protein–substrate interactions.
Acta Crystallographica Section D-biological Crystallography | 2003
Ingar Leiros; Elin Moe; Olav Lanes; Arne O. Smalås; Nils Peder Willassen
Uracil-DNA glycosylase (UDG; EC 3.2.2.3) is a DNA-repair protein that catalyses the hydrolysis of promutagenic uracil residues from single- or double-stranded DNA, generating free uracil and abasic DNA. The crystal structure of the catalytic domain of cod uracil-DNA glycosylase (cUDG) has been determined to 1.9 A resolution, with final R factors of 18.61 and 20.57% for the working and test sets of reflections, respectively. This is the first crystal structure of a uracil-DNA glycosylase from a cold-adapted species and a detailed comparison with the human enzyme is performed in order to rationalize the cold-adapted behaviour of the cod enzyme at the structural level. The catalytic domain of cUDG comprises 223 residues, with a sequence identity to the human UDG of 75%. The tertiary structures of the two enzymes are also similar, with an overall displacement in main-chain atomic positions of 0.63 A. The amino-acid substitutions and the differences in intramolecular hydrogen bonds, hydrophobic interactions, ion-pair interactions and electrostatic potentials are compared and discussed in order to gain insight into the factors that cause the increased activity and reduced thermostability of the cod enzyme. In particular, the reduced number of strong ion-pair interactions in the C-terminal half of cUDG is believed to greatly affect the flexibility and/or stability. Increased positive electrostatic surface potential on the DNA-facing side of cUDG seems to be responsible for increasing the affinity for the negatively charged DNA compared with that of hUDG.
Acta Crystallographica Section D-biological Crystallography | 2001
Hanna-Kirsti S. Leiros; Sean McSweeney; Arne O. Smalås
Radiation damage is an inherent problem in protein X-ray crystallography and the process has recently been shown to be highly specific, exhibiting features such as cleavage of disulfide bonds, decarboxylation of acidic residues, increase in atomic B factors and increase in unit-cell volume. Reported here are two trypsin structures at atomic resolution (1.00 and 0.95 A), the data for which were collected at a third-generation synchrotron (ESRF) at two different beamlines. Both trypsin structures exhibit broken disulfide bonds; in particular, the bond from Cys191 to Cys220 is very sensitive to synchrotron radiation. The data set collected at the most intense beamline (ID14-EH4) shows increased structural radiation damage in terms of lower occupancies for cysteine residues, more breakage in the six disulfide bonds and more alternate conformations. It appears that high intensity and not only the total X-ray dose is most harmful to protein crystals.
Protein Science | 2001
Bjørn Olav Brandsdal; Johan Åqvist; Arne O. Smalås
The binding of P1 variants of bovine pancreatic trypsin inhibitor (BPTI) to trypsin has been investigated by means of molecular dynamics simulations. The specific interaction formed between the amino acid at the primary binding (P1) position of the binding loop of BPTI and the specificity pocket of trypsin was estimated by use of the linear interaction energy (LIE) method. Calculations for 13 of the naturally occurring amino acids at the P1 position were carried out, and the results obtained were found to correlate well with the experimental binding free energies. The LIE calculations rank the majority of the 13 variants correctly according to the experimental association energies and the mean error between calculated and experimental binding free energies is only 0.38 kcal/mole, excluding the Glu and Asp variants, which are associated with some uncertainties regarding protonation and the possible presence of counter‐ions. The three‐dimensional structures of the complex with three of the P1 variants (Asn, Tyr, and Ser) included in this study have not at present been solved by any experimental techniques and, therefore, were modeled on the basis of experimental data from P1 variants of similar size. Average structures were calculated from the MD simulations, from which specific interactions explaining the broad variation in association energies were identified. The present study also shows that explicit treatment of the complex water‐mediated hydrogen bonding network at the protein–protein interface is of crucial importance for obtaining reliable binding free energies. The successful reproduction of relative binding energies shows that this type of methodology can be very useful as an aid in rational design and redesign of biologically active macromolecules.
FEBS Letters | 2001
Bjørn Olav Brandsdal; Arne O. Smalås; Johan Åqvist
Organisms that live in constantly cold environments have to adapt their metabolism to low temperatures, but mechanisms of enzymatic adaptation to cold environments are not fully understood. Cold active trypsin catalyses reactions more efficiently and binds ligands more strongly in comparison to warm active trypsin. We have addressed this issue by means of comparative free energy calculations studying the binding of positively charged ligands to two trypsin homologues. Stronger inhibition of the cold active trypsin by benzamidine and positively charged P1‐variants of BPTI is caused by rather subtle electrostatic effects. The different affinity of benzamidine originates solely from long range interactions, while the increased binding of P1–Lys and –Arg variants of BPTI is attributed to both long and short range effects that are enhanced in the cold active trypsin compared to the warm active counterpart. Electrostatic interactions thus provide an efficient strategy for cold adaptation of trypsin.
Proteins | 2000
Alemayehu A. Gorfe; Bjørn Olav Brandsdal; Hanna-Kirsti S. Leiros; Ronny Helland; Arne O. Smalås
A qualitative evaluation of electrostatic features of the substrate binding region of seven isoenzymes of trypsin has been performed by using the continuum electrostatic model for the solution of the Poisson‐Boltzmann equation. The sources of the electrostatic differences among the trypsins have been sought by comparative calculations on selective charges: all charges, conserved charges, partial charges, unique cold trypsin charges, and a number of charge mutations. As expected, most of the negative potential at the S1 region of all trypsins is generated from Asp189, but the potential varies significantly among the seven trypsin isoenzymes. The three cold active enzymes included in this study possess a notably lower potential at and around the S1‐pocket compared with the warm active counterparts; this finding may be the main contribution to the increased binding affinity. The source of the differences are nonconserved charged residues outside the specificity pocket, producing electric fields at the S1‐pocket that are different in both sign and magnitude. The surface charges of the mesophilic trypsins generally induce the S1 pocket positively, whereas surface charges of the cold trypsins produce a negative electric field of this region. Calculations on mutants, where charged amino acids were substituted between the trypsins, showed that mutations in Loop2 (residues 221B and 224) and residue 175, in particular, were responsible for the low potential of the cold enzymes. Proteins 2000;40:207–217.
Journal of Biomolecular Structure & Dynamics | 1999
Bjørn Olav Brandsdal; Eldbjørg Sofie Heimstad; Ingebrigt Sylte; Arne O. Smalås
It is well established that the dynamic motion of proteins plays an important functional role, and that the adaptation of a protein molecule to its environment requires optimization of internal non-covalent interactions and protein-solvent interactions. Serine proteinases in general, and trypsin in particular has been used as a model system in exploring possible structural features for cold adaptation. In this study, a 500 p.s. and a 1200 p.s. molecular dynamics (MD) simulation at 300 K of both anionic salmon trypsin and cationic bovine trypsin are analyzed in terms of molecular flexibility, internal non-covalent interactions and protein-solvent interactions. The present MD simulations do not indicate any increased flexibility of the cold adapted enzyme on an overall basis. However, the apparent higher flexibility and deformability of the active site of anionic salmon trypsin may lower the activation energy for ligand binding and for catalysis, and might be a reason for the increased binding affinity and catalytic efficiency compared to cationic bovine trypsin.
Acta Crystallographica Section D-biological Crystallography | 1995
Gunnar I. Berglund; Nils Peder Willassen; A. Hordvik; Arne O. Smalås
The crystal structure of native salmon pancreatic elastase (SPE) has been solved by molecular-replacement methods, and refined by conventional conjugate-gradient methods and simulated-annealing techniques. The final R value is 17.2% for 21 389 reflections between 8.0 and 1.61 A, and the corresponding free R value is 23.9%. The overall tertiary structure of SPE is remarkably similar to that of porcine pancreatic elastase I (PPE), to which it shows about 67% sequence identity. The primary structure of SPE is determined from the electron-density maps, and only about 15 side chains are somewhat uncertain. Interesting differences between SPE and PPE, are one sequence deletion assigned to position 186, the residue 192 at the entrance of the specificity pocket is substituted from a Gln in PPE to Asn in SPE, and one of the calcium ligands is different. Furthermore, electron density is missing in SPE for the last three residues of the C-terminal helix. A comparison of the present amino-acid sequence of SPE with other sequences available indicates that SPE belongs to the class 1 pancreatic elastases.
FEBS Journal | 2006
Ronny Helland; Atle Noralf Larsen; Arne O. Smalås; Nils Peder Willassen
Proteins from organisms living in extreme conditions are of particular interest because of their potential for being templates for redesign of enzymes both in biotechnological and other industries. The crystal structure of a proteinase K‐like enzyme from a psychrotroph Serratia species has been solved to 1.8 Å. The structure has been compared with the structures of proteinase K from Tritirachium album Limber and Vibrio sp. PA44 in order to reveal structural explanations for differences in biophysical properties. The Serratia peptidase shares around 40 and 64% identity with the Tritirachium and Vibrio peptidases, respectively. The fold of the three enzymes is essentially identical, with minor exceptions in surface loops. One calcium binding site is found in the Serratia peptidase, in contrast to the Tritirachium and Vibrio peptidases which have two and three, respectively. A disulfide bridge close to the S2 site in the Serratia and Vibrio peptidases, an extensive hydrogen bond network in a tight loop close to the substrate binding site in the Serratia peptidase and different amino acid sequences in the S4 sites are expected to cause different substrate specificity in the three enzymes. The more negative surface potential of the Serratia peptidase, along with a disulfide bridge close to the S2 binding site of a substrate, is also expected to contribute to the overall lower binding affinity observed for the Serratia peptidase. Clear electron density for a tripeptide, probably a proteolysis product, was found in the S’ sites of the substrate binding cleft.