hosted by
publicationslist.org
    

Jüri jyri.jarvet@gmail.com Jarvet


jyri.jarvet@gmail.com

Journal articles

2008
Anna Wahlström, Loïc Hugonin, Alex Perálvarez-Marín, Jüri Jarvet, Astrid Gräslund (2008)  Secondary structure conversions of Alzheimer's Abeta(1-40) peptide induced by membrane-mimicking detergents.   FEBS J 275: 20. 5117-5128 Oct  
Abstract: The amyloid beta peptide (Abeta) with 39-42 residues is the major component of amyloid plaques found in brains of Alzheimer's disease patients, and soluble oligomeric peptide aggregates mediate toxic effects on neurons. The Abeta aggregation involves a conformational change of the peptide structure to beta-sheet. In the present study, we report on the effect of detergents on the structure transitions of Abeta, to mimic the effects that biomembranes may have. In vitro, monomeric Abeta(1-40) in a dilute aqueous solution is weakly structured. By gradually adding small amounts of sodium dodecyl sulfate (SDS) or lithium dodecyl sulfate to a dilute aqueous solution, Abeta(1-40) is converted to beta-sheet, as observed by CD at 3 degrees C and 20 degrees C. The transition is mainly a two-state process, as revealed by approximately isodichroic points in the titrations. Abeta(1-40) loses almost all NMR signals at dodecyl sulfate concentrations giving rise to the optimal beta-sheet content (approximate detergent/peptide ratio = 20). Under these conditions, thioflavin T fluorescence measurements indicate a maximum of aggregated amyloid-like structures. The loss of NMR signals suggests that these are also involved in intermediate chemical exchange. Transverse relaxation optimized spectroscopy NMR spectra indicate that the C-terminal residues are more dynamic than the others. By further addition of SDS or lithium dodecyl sulfate reaching concentrations close to the critical micellar concentration, CD, NMR and FTIR spectra show that the peptide rearranges to form a micelle-bound structure with alpha-helical segments, similar to the secondary structures formed when a high concentration of detergent is added directly to the peptide solution.
Notes:
Kristina Westerlund, Sean D Moran, Heidi K Privett, Sam Hay, Jüri Jarvet, Brian R Gibney, Cecilia Tommos (2008)  Making a single-chain four-helix bundle for redox chemistry studies.   Protein Eng Des Sel 21: 11. 645-652 Nov  
Abstract: The construction and characteristics of the stable and well-structured alpha(4)W protein are described. The 117-residue, single-chain protein has a molecular weight of 13.1 kDa and is designed to fold into a four-helix bundle. Experimental characterization of the expressed and purified protein shows a 69.8 +/- 0.8% helical content over a 5.5-10.0 pH range. The protein is thermostable with a T(M) > 355 K and has a free energy of unfolding as measured by chemical denaturation of -4.7 kcal mol(-1) at 25 degrees C and neutral pH. One-dimensional (1D) proton and 2D (15)N-HSQC spectra show narrow, well-dispersed spectral lines consistent with a uniquely structured alpha-helical protein. Analytical ultracentrifugation and NMR data show that the protein is monomeric over a broad protein concentration range. The 324 nm emission maximum of the unique Trp-106 is consistent with a sequestered position of the aromatic residue. Additionally, differential pulse voltammetry characterization indicates an elevated peak potential for Trp-106 when the protein is folded (pH range 7.0-8.5) relative to partly unfolded (pH range 11.4-13.2). The oxidation of Trp-106 is coupled to proton release as shown by a 53 +/- 3 mV/pH unit dependence of the peak potential over the 7.0-8.5 pH range.
Notes:
2007
Tariq Massad, Jüri Jarvet, Risto Tanner, Katrin Tomson, Julia Smirnova, Peep Palumaa, Mariko Sugai, Toshiyuki Kohno, Kalju Vanatalu, Peter Damberg (2007)  Maximum entropy reconstruction of joint phi, psi-distribution with a coil-library prior: the backbone conformation of the peptide hormone motilin in aqueous solution from phi and psi-dependent J-couplings.   J Biomol NMR 38: 2. 107-123 Jun  
Abstract: In this paper, we present a new method for structure determination of flexible "random-coil" peptides. A numerical method is described, where the experimentally measured 3J(H(alpha)Nalpha) and [3J(H(alpha)Nalpha+1 couplings, which depend on the phi and psi dihedral angles, are analyzed jointly with the information from a coil-library through a maximum entropy approach. The coil-library is the distribution of dihedral angles found outside the elements of the secondary structure in the high-resolution protein structures. The method results in residue specific joint phi,psi-distribution functions, which are in agreement with the experimental J-couplings and minimally committal to the information in the coil-library. The 22-residue human peptide hormone motilin, uniformly 15N-labeled was studied. The 3J(H(alpha)-N(i+1)) were measured from the E.COSY pattern in the sequential NOESY cross-peaks. By employing homodecoupling and an in-phase/anti-phase filter, sharp H(alpha)-resonances (about 5 Hz) were obtained enabling accurate determination of the coupling with minimal spectral overlap. Clear trends in the resulting phi,psi-distribution functions along the sequence are observed, with a nascent helical structure in the central part of the peptide and more extended conformations of the receptor binding N-terminus as the most prominent characteristics. From the phi,psi-distribution functions, the contribution from each residue to the thermodynamic entropy, i.e., the segmental entropies, are calculated and compared to segmental entropies estimated from 15N-relaxation data. Remarkable agreement between the relaxation and J-couplings based methods is found. Residues belonging to the nascent helix and the C-terminus show segmental entropies, of approximately -20 J K(-1) mol(-1) and -12 J K(-1) mol(-1), respectively, in both series. The agreement between the two estimates of the segmental entropy, the agreement with the observed J-couplings, the agreement with the CD experiments, and the assignment of population to sterically allowed conformations show that the phi,psi-distribution functions are indeed meaningful and useful descriptions of the conformational preferences for each residue in this flexible peptide.
Notes:
Jüri Jarvet, Jens Danielsson, Peter Damberg, Marta Oleszczuk, Astrid Gräslund (2007)  Positioning of the Alzheimer Abeta(1-40) peptide in SDS micelles using NMR and paramagnetic probes.   J Biomol NMR 39: 1. 63-72 Sep  
Abstract: NMR spectroscopy combined with paramagnetic relaxation agents was used to study the positioning of the 40-residue Alzheimer Amyloid beta-peptide Abeta(1-40) in SDS micelles. 5-Doxyl stearic acid incorporated into the micelle or Mn(2+) ions in the aqueous solvent were used to determine the position of the peptide relative to the micelle geometry. In SDS solvent, the two alpha-helices induced in Abeta(1-40), comprising residues 15-24, and 29-35, respectively, are surrounded by flexible unstructured regions. NMR signals from these unstructured regions are strongly attenuated in the presence of Mn(2+) showing that these regions are positioned mostly outside the micelle. The central helix (residues 15-24) is significantly affected by 5-doxyl stearic acid however somewhat less for residues 16, 20, 22 and 23. This alpha-helix therefore resides in the SDS headgroup region with the face with residues 16, 20, 22 and 23 directed away from the hydrophobic interior of the micelle. The C-terminal helix is protected both from 5-doxyl stearic acid and Mn(2+), and should be buried in the hydrophobic interior of the micelle. The SDS micelles were characterized by diffusion and (15)N-relaxation measurements. Comparison of experimentally determined translational diffusion coefficients for SDS and Abeta(1-40) show that the size of SDS micelle is not significantly changed by interaction with Abeta(1-40).
Notes:
2006
Jens Danielsson, August Andersson, Jüri Jarvet, Astrid Gräslund (2006)  15N relaxation study of the amyloid beta-peptide: structural propensities and persistence length.   Magn Reson Chem 44 Spec No: S114-S121 Jul  
Abstract: The dynamics of monomeric Alzheimer Abeta1-40 in aqueous solution was studied using heteronuclear NMR experiments. 15N NMR relaxation rates of amide groups report on the dynamics in the peptide chain and make it possible to estimate structural propensities from temperature-dependent relaxation data and chemical shifts change analysis. The persistence length of the polypeptide chain was determined using a model in which the influence of neighboring residue relaxation is assumed to decay exponentially as a function of distance. The persistence length of the Abeta1-40 monomer was found to decrease from eight to three residues when temperature was increased from 3 to 18 degrees C. At 3 degrees C the peptide shows structural propensities that correlate well with the suggested secondary structure regions of the peptide to be present in the fibrils, and with the alpha-helical structure in membrane-mimicking systems. Our data leads to a structural model for the monomeric soluble beta-peptide with six different regions of secondary structure propensities. The peptide has two regions with beta-strand propensity (residues 16-24 and 31-40), two regions with high PII-helix propensity (residues 1-4 and 11-15) and two unstructured regions with higher mobility (residues 5-10 and 25-30) connecting the structural elements.
Notes:
Evangelos Papadopoulos, Kamila Oglecka, Lena Mäler, Jüri Jarvet, Peter E Wright, H Jane Dyson, Astrid Gräslund (2006)  NMR solution structure of the peptide fragment 1-30, derived from unprocessed mouse Doppel protein, in DHPC micelles.   Biochemistry 45: 1. 159-166 Jan  
Abstract: The downstream prion-like Doppel (Dpl) protein is a homologue related to the prion protein (PrP). Dpl is expressed in the brains of mice that do not express PrP, and Dpl is known to be toxic to neurons. One mode of toxicity has been suggested to involve direct membrane interactions. PrP under certain conditions of cell trafficking retains an uncleaved signal peptide, which may also hold for the much less studied Dpl. For a peptide with a sequence derived from the N-terminal part (1-30) of mouse Dpl (mDpl(1-30)) CD spectroscopy shows about 40% alpha-helical structure in DHPC and SDS micelles. In aqueous solution it is mostly a random coil. The three-dimensional solution structure was determined by NMR for mDpl(1-30) associated with DHPC micelles. 2D 1H NMR spectra of the peptide in q = 0.25 DMPC/DHPC bicelles only showed signals from the unstructured termini, indicating that the structured part of the peptide resides within the lipid bilayer. Together with 2H2O exchange data in the DHPC micelle solvent, these results show an alpha-helix protected from solvent exchange between residues 7 and 19, and suggest that the alpha-helical segment can adopt a transmembrane localization also in a membrane. Leakage studies with entrapped calcein in large unilamellar phospholipid vesicles showed that the peptide is almost as membrane perturbing as melittin, known to form pores in membranes. The results suggest a possible channel formation mechanism for the unprocessed Dpl protein, which may be related to toxicity through direct cell membrane interaction and damage.
Notes:
Merike Vaher, Säde Viirlaid, Kersti Ehrlich, Riina Mahlapuu, Jüri Jarvet, Ursel Soomets, Mihkel Kaljurand (2006)  Characterization of the antioxidative activity of novel nontoxic neuropeptides by using capillary electrophoresis.   Electrophoresis 27: 13. 2582-2589 Jul  
Abstract: In the present study, we have monitored the oxidation process of novel nontoxic neuropeptides and determined its rate constants, which describe the antioxidative potential of the peptides. A capillary electrophoretic method was implemented which ensures the simultaneity of analysis of reactants and products in a short time of analysis. The rate constants of oxidation of the four novel peptides, 4-methoxy-L-tyrosinyl-gamma-L-glutamyl-L-cysteinyl-glycine (UPF1), D-serinyl-gamma-L-glutamyl-L-cysteinyl-glycine (UPF6), 4-methoxy-L-tyrosinyl-alpha-L-glutamyl-L-cysteinyl-glycine and D-serinyl-alpha-L-glutamyl-L-cysteinyl-glycine, designed by us, were compared with those of oxidation of glutathione (reduced glutathione) by using capillary electrophoresis. The second-order rate constants were similar for all peptides if the oxidation was carried out with hydrogen peroxide (k(II) = 0.208 - 0.236 x 10(3)/M.min). The rate constants were also determined for the mixtures of peptides. When the oxidation is caused by hydroxyl radical (OH*), the gamma-glutamate containing peptides (UPF1 and UPF6) exhibited two to four times higher antioxidative activity (k(II) = 4.428 and 2.152 x 10(3)/M.min, respectively). The results suggest that the antioxidative potential of the peptides studied is not determined by the formation of disulphide bridge alone.
Notes:
2005
Magnus Lindskog, Christian Spenger, Tomas Klason, Jüri Jarvet, Astrid Gräslund, John Inge Johnsen, Frida Ponthan, Lena Douglas, Bo Nordell, Per Kogner (2005)  Proton magnetic resonance spectroscopy in neuroblastoma: current status, prospects and limitations.   Cancer Lett 228: 1-2. 247-255 Oct  
Abstract: Non-invasive biological information about residual neuroblastoma tumour tissue could allow treatment monitoring without the need for repeated biopsies. Magnetic resonance spectroscopy (MRS) can be performed with standard MR-scanners, providing specific biochemical information from selected tumour regions. By proton 1H-MRS, lipids, certain amino acids and lactate can be detected and their relative concentrations estimated in vivo. Using experimental models of neuroblastoma, we have described the potential of 1H-MRS for the prediction of tumour tissue viability and treatment response. Whereas viable neuroblastoma tissue is dominated by the choline 1H-MRS resonance, cell death as a consequence of spontaneous necrosis or successful treatment with chemotherapy, angiogenesis inhibitors, or NSAIDs is associated with decreased choline content. Therapy-induced neuroblastoma cell death is also associated with enhanced 1H-MRS resonances from mobile lipids and polyunsaturated fatty acids. The mobile lipid/choline ratio correlates significantly with cell death and based on the dynamics of this ratio tumour regression or continued growth (drug resistance) after chemotherapy can be predicted in vivo. The implications of these findings are discussed with focus on the potentials and limitations of introducing 1H-MRS for clinical assessment of treatment response in children with neuroblastoma. Biochemical monitoring of neuroblastoma with 1H-MRS could enable tailoring of individual therapy as well as provide early pharmacodynamic evaluation of novel therapeutic modalities.
Notes:
Jens Danielsson, Jüri Jarvet, Peter Damberg, Astrid Gräslund (2005)  The Alzheimer beta-peptide shows temperature-dependent transitions between left-handed 3-helix, beta-strand and random coil secondary structures.   FEBS J 272: 15. 3938-3949 Aug  
Abstract: The temperature-induced structural transitions of the full length Alzheimer amyloid beta-peptide [A(beta)(1-40) peptide] and fragments of it were studied using CD and 1H NMR spectroscopy. The full length peptide undergoes an overall transition from a state with a prominent population of left-handed 3(1) (polyproline II; PII)-helix at 0 degrees C to a random coil state at 60 degrees C, with an average DeltaH of 6.8 +/- 1.4 kJ.mol(-1) per residue, obtained by fitting a Zimm-Bragg model to the CD data. The transition is noncooperative for the shortest N-terminal fragment A(beta)(1-9) and weakly cooperative for A(beta)(1-40) and the longer fragments. By analysing the temperature-dependent 3J(HNH(alpha)) couplings and hydrodynamic radii obtained by NMR for A(beta)(1-9) and A(beta)(12-28), we found that the structure transition includes more than two states. The N-terminal hydrophilic A(beta)(1-9) populates PII-like conformations at 0 degrees C, then when the temperature increases, conformations with dihedral angles moving towards beta-strand at 20 degrees C, and approaches random coil at 60 degrees C. The residues in the central hydrophobic (18-28) segment show varying behaviour, but there is a significant contribution of beta-strand-like conformations at all temperatures below 20 degrees C. The C-terminal (29-40) segment was not studied by NMR, but from CD difference spectra we concluded that it is mainly in a random coil conformation at all studied temperatures. These results on structural preferences and transitions of the segments in the monomeric form of A(beta) may be related to the processes leading to the aggregation and formation of fibrils in the Alzheimer plaques.
Notes:
Peter Damberg, Jüri Jarvet, Astrid Gräslund (2005)  Limited variations in 15N CSA magnitudes and orientations in ubiquitin are revealed by joint analysis of longitudinal and transverse NMR relaxation.   J Am Chem Soc 127: 6. 1995-2005 Feb  
Abstract: The site-specific magnitudes and orientations of the chemical shift tensors have been estimated for 70 backbone (15)N-nuclei in human ubiquitin from the field dependence of dynamic independent ratios between relaxation rates, both longitudinal and transverse, measured at 9.4, 11.7, 14.1, and 18.8 T. The results were jointly analyzed with previously published relaxation data [Fushman; Tjandra; Cowburn. J.Am. Chem. Soc. 1998, 120, 10947-10952] [Kover; Batta. J. Mag. Reson. 2001, 150, 137-146]. The effective magnitudes of the anisotropies distribute around 169 ppm with a variability of 5 ppm. The orientation factors, reflecting the orientation of the CSA relative to the NH bond, distribute around -0.80 with a variability of 0.04, which corresponds to an angle between the symmetry axis of an assumed axially symmetric shielding tensor and the NH bond of 21.4 degrees, and a variability of 2.3 degrees. Correlations with the isotropic (15)N-chemical shifts are observed. Variations in the shielding anisotropies add uncertainty to the obtained order parameters proportional to the square of the magnetic field, when data are analyzed using an assumed invariant CSA tensor for all sites. Around 3% additional uncertainty in the order parameters for 800 MHz data is expected. The optimal TROSY field for amide nitrogen TROSY is estimated, with only marginal variations due to site-to-site variations. Variations in the shielding tensors add uncertainty to the exchange terms calculated from cross-correlation rates. An approach for estimating the exchange terms is suggested, where the uncertainty due to CSA-variations is minimized.
Notes:
2004
A Päiviö, J Jarvet, A Gräslund, L Lannfelt, A Westlind-Danielsson (2004)  Unique physicochemical profile of beta-amyloid peptide variant Abeta1-40E22G protofibrils: conceivable neuropathogen in arctic mutant carriers.   J Mol Biol 339: 1. 145-159 May  
Abstract: A new early-onset form of Alzheimer's disease (AD) was described recently where a point mutation was discovered in codon 693 of the beta-amyloid (Abeta) precursor protein gene, the Arctic mutation. The mutation translates into a single amino acid substitution, glutamic acid-->glycine, in position 22 of the Abeta peptide. The mutation carriers have lower plasma levels of Abeta than normal, while in vitro studies show that Abeta1-40E22G protofibril formation is significantly enhanced. We have explored the nature of the Abeta1-40E22G peptide in more detail, in particular the protofibrils. Using size-exclusion chromatography (SEC) and circular dichroism spectroscopy (CD) kinetic and secondary structural characteristics were compared with other Abeta1-40 peptides and the Abeta12-28 fragment, all having single amino acid substitutions in position 22. We have found that Abeta1-40E22G protofibrils are a group of comparatively stabile beta-sheet-containing oligomers with a heterogeneous size distribution, ranging from >100 kDa to >3000 kDa. Small Abeta1-40E22G protofibrils are generated about 400 times faster than large ones. Salt promotes their formation, which significantly exceeds all the other peptides studied here, including the Dutch mutation Abeta1-40E22Q. Position 22 substitutions had significant effects on aggregation kinetics of Abeta1-40 and in Abeta12-28, although the qualitative aspects of the effects differed between the native peptide and the fragment, as no protofibrils were formed by the fragments. The rank order of protofibril formation of Abeta1-40 and its variants was the same as the rank order of the length of the nucleation/lag phase of the Abeta12-28 fragments, E22V>E22A?E22G>E22Q?E22, and correlated with the degree of hydrophobicity of the position 22 substituent. The molecular mass of peptide monomers and protofibrils were estimated better in SEC studies using linear rather than globular calibration standards. The characteristics of the Abeta1-40E22G suggest an important role for the peptide in the neuropathogenesis in the Arctic form of AD.
Notes:
Jens Danielsson, Jüri Jarvet, Peter Damberg, Astrid Gräslund (2004)  Two-site binding of beta-cyclodextrin to the Alzheimer Abeta(1-40) peptide measured with combined PFG-NMR diffusion and induced chemical shifts.   Biochemistry 43: 20. 6261-6269 May  
Abstract: The interactions of Alzheimer's amyloid beta-peptide with cyclodextrins were studied by (1)H NMR: the translational diffusion coefficient of the peptide and chemical shift changes were studied by the presence of variable concentrations of cyclodextrins. For the full-length peptide, Abeta(1-40), the combined results of translational diffusion and chemical shift changes are consistent with a model where aromatic side chains interact with beta-cyclodextrin with dissociation constants in the millimolar range. The diffusion data were consistent with two beta-cyclodextrin molecules bound per peptide. The binding occurs at two sites, at F(19) and/or F(20) and at Y(10), with dissociation constants K(d)(F) = 4.7 mM and K(d)(Y) = 6.6 mM, respectively, in 10 mM sodium phosphate, pH 7.4 and 298 K. Shorter Alzheimer peptide fragments were studied to measure specific affinities for different binding sites. The N-terminal fragment Abeta(1-9) with a putative binding site at F(4) does not show measurable affinity for beta-cyclodextrin. The fragment Abeta(12-28) has similar apparent affinity (K(d) = 3.8 mM) to beta-cyclodextrin as the full-length peptide Abeta(1-40). Here, the diffusion data suggests a one-to-one stoichiometry, and the binding site is F(19) and/or F(20). Both diffusion results and chemical shift changes give the same affinity. A variant Abeta(12-28)G(19)G(20) without phenylalanines does not bind to beta-cyclodextrin. Other potential ligands, alpha-cyclodextrin, gamma-cyclodextrin, nicotine, and nornicotine do not bind to the Abeta(12-28) fragment. This study shows that combined (1)H NMR diffusion and chemical shift changes may be used to quantitatively determine affinities and stoichiometries of weak interactions, using unlabeled ligands and hosts of comparable sizes.
Notes:
Magnus Lindskog, Christian Spenger, Jüri Jarvet, Astrid Gräslund, Per Kogner (2004)  Predicting resistance or response to chemotherapy by proton magnetic resonance spectroscopy in neuroblastoma.   J Natl Cancer Inst 96: 19. 1457-1466 Oct  
Abstract: We previously showed that proton magnetic resonance spectroscopy (1H-MRS) enables estimation of neuroblastoma tumor viability. Here we investigated if 1H-MRS can predict response or resistance to chemotherapy in neuroblastoma.
Notes:
2003
J Jarvet, P Damberg, J Danielsson, I Johansson, L E G Eriksson, A Gräslund (2003)  A left-handed 3(1) helical conformation in the Alzheimer Abeta(12-28) peptide.   FEBS Lett 555: 2. 371-374 Dec  
Abstract: We show for the first time that the secondary structure of the Alzheimer beta-peptide is in a temperature-dependent equilibrium between an extended left-handed 3(1) helix and a flexible random coil conformation. Circular dichroism spectra, recorded at 0.03 mM peptide concentration, show that the equilibrium is shifted towards increasing left-handed 3(1) helix structure towards lower temperatures. High resolution nuclear magnetic resonance (NMR) spectroscopy has been used to study the Alzheimer peptide fragment Abeta(12-28) in aqueous solution at 0 degrees C and higher temperatures. NMR translation diffusion measurements show that the observed peptide is in monomeric form. The chemical shift dispersion of the amide protons increases towards lower temperatures, in agreement with the increased population of a well-ordered secondary structure. The solvent exchange rates of the amide protons at 0 degrees C and pH 4.5 vary within at least two orders of magnitude. The lowest exchange rates (0.03-0.04 min(-1)) imply that the corresponding amide protons may be involved in hydrogen bonding with neighboring side chains.
Notes:
2002
P Lundberg, M Magzoub, M Lindberg, M Hällbrink, J Jarvet, L E G Eriksson, U Langel, A Gräslund (2002)  Cell membrane translocation of the N-terminal (1-28) part of the prion protein.   Biochem Biophys Res Commun 299: 1. 85-90 Nov  
Abstract: The N-terminal (1-28) part of the mouse prion protein (PrP) is a cell penetrating peptide, capable of transporting large hydrophilic cargoes through a cell membrane. Confocal fluorescence microscopy shows that it transports the protein avidin (67kDa) into several cell lines. The (1-28) peptide has a strong tendency for aggregation and beta-structure formation, particularly in interaction with negatively charged phospholipid membranes. The findings have implications for how prion proteins with uncleaved signal peptides in the N-termini may enter into cells, which is important for infection. The secondary structure conversion into beta-structure may be relevant as a seed for the conversion into the scrapie (PrP(Sc)) form of the protein and its amyloidic transformation.
Notes:
Peter Damberg, Jüri Jarvet, Peter Allard, Ulo Mets, Rudolf Rigler, Astrid Gräslund (2002)  (13)C-(1)H NMR relaxation and fluorescence anisotropy decay study of tyrosine dynamics in motilin.   Biophys J 83: 5. 2812-2825 Nov  
Abstract: Tyrosine ring dynamics of the gastrointestinal hormone motilin was studied using two independent physical methods: fluorescence polarization anisotropy decay and NMR relaxation. Motilin, a 22-residue peptide, was selectively (13)C labeled in the ring epsilon-carbons of the single tyrosine residue. To eliminate effects of differences in peptide concentration, the same motilin sample was used in both experiments. NMR relaxation rates of the tyrosine ring C(epsilon)-H(epsilon) vectors, measured at four magnetic field strengths (9.4, 11.7, 14.1, and 18.8 Tesla) were used to map the spectral density function. When the data were analyzed using dynamic models with the same number of components, the dynamic parameters from NMR and fluorescence are in excellent agreement. However, the estimated rotational correlation times depend on the choice of dynamic model. The correlation times estimated from the two-component model-free approach and the three-component models were significantly different (1.7 ns and 2.2 ns, respectively). Various earlier studies of protein dynamics by NMR and fluorescence were compared. The rotational correlation times estimated by NMR for samples with high protein concentration were on average 18% longer for folded monomeric proteins than the corresponding times estimated by fluorescence polarization anisotropy decay, after correction for differences in viscosity due to temperature and D(2)O/H(2)O ratio.
Notes:
2001
M Lindberg, J Jarvet, U Langel, A Gräslund (2001)  Secondary structure and position of the cell-penetrating peptide transportan in SDS micelles as determined by NMR.   Biochemistry 40: 10. 3141-3149 Mar  
Abstract: Transportan is a 27-residue peptide (GWTLN SAGYL LGKIN LKALA ALAKK IL-amide) which has the ability to penetrate into living cells carrying a hydrophilic load. Transportan is a chimeric peptide constructed from the 12 N-terminal residues of galanin in the N-terminus with the 14-residue sequence of mastoparan in the C-terminus and a connecting lysine. Circular dichroism studies of transportan and mastoparan show that both peptides have close to random coil secondary structure in water. Sodium dodecyl sulfate (SDS) micelles induce 60% helix in transportan and 75% helix in mastoparan. The 600 MHz (1)H NMR studies of secondary structure in SDS micelles confirm the helix in mastoparan and show that in transportan the helix is localized to the mastoparan part. The less structured N-terminus of transportan has a secondary structure similar to that of the same sequence in galanin [Ohman, A., et al. (1998) Biochemistry 37, 9169-9178]. The position of mastoparan and transportan relative to the SDS micelle surface was studied by adding spin-labeled 5-doxyl- or 12-doxyl-stearic acid or Mn2+ to the peptide/micelle system. The combined results show that the peptides are for the most part buried in the SDS micelles. Only the C-terminal parts of both peptides and the central segment connecting the two parts of transportan are clearly surface exposed. For mastoparan, the secondary chemical shifts of the amide protons were found to vary periodically and display a pattern almost identical to those reported for mastoparan in phospholipid bicelles [Vold, R., et al. (1997) J. Biomol. NMR 9, 329-335], indicating similar structures and interactions in the two membrane-mimicking environments.
Notes:
P Damberg, J Jarvet, A Gräslund (2001)  Accurate measurement of translational diffusion coefficients: a practical method to account for nonlinear gradients.   J Magn Reson 148: 2. 343-348 Feb  
Abstract: For NMR probes equipped with pulsed field gradient coils, which are not optimized for gradient linearity, the precision and accuracy of experimentally measured translational diffusion coefficients are limited by the linearity of the gradient pulses over the sample volume. This study shows that the accuracy and precision of measured diffusion coefficients by the Stejskal--Tanner spin-echo pulsed field gradient experiment can be significantly improved by mapping the gradient z-profile and by using the mapped calibration parameters in the data analysis. For practical applications the gradient distribution may be approximated by a truncated linear distribution defined by minimum and maximum values of the gradient. By including the truncated linear gradient distribution function in the Stejskal--Tanner equation, the systematic deviation between the fitted curve and the experimental attenuation curve decreases by an order of magnitude. The gradient distribution may be calibrated using an intense NMR signal from a sample with a known diffusion coefficient. The diffusion coefficient of an unknown sample may then be determined from a two-parameter fit, using the known gradient distribution function.
Notes:
1999
A M Blom, M Mörgelin, M Oyen, J Jarvet, E Fries (1999)  Structural characterization of inter-alpha-inhibitor. Evidence for an extended shape.   J Biol Chem 274: 1. 298-304 Jan  
Abstract: Inter-alpha-inhibitor (IalphaI) is a 180-kDa serum protein consisting of three polypeptides. Two of these, the heavy chains 1 and 2 (H1 and H2), are of 75-80 kDa and have similar amino acid sequences. The third polypeptide, bikunin, has a molecular mass of 25 kDa and contains a 7-kDa chondroitin sulfate chain that is covalently linked to the C-terminal amino acid residues of H1 and H2. IalphaI has been shown to be required for the formation of the hyaluronan-containing extracellular matrix of certain cell types. How IalphaI exerts this function is not known, but it appears that upon interaction with cells, the heavy chains are released and become covalently linked to hyaluronan. Our results indicate that IalphaI and its heavy chains are extended molecules; thus, upon electron microscopy, IalphaI appeared to consist of two globular domains connected by a thin structure 31-nm long and the isolated heavy chains of a globular domain and a "tail" about 15-nm long. Analysis of the heavy chains by partial proteolysis showed that the C-terminal halves are particularly sensitive to hydrolysis indicating that they are loosely folded. Furthermore, electron microscopy showed that partially degraded heavy chains lacked the extended regions. Taken together, these results suggest that the N-terminal half of the heavy chains forms a globular domain, whereas the other half has an extended and loosely folded structure.
Notes:
U Soomets, R Mahlapuu, R Tehranian, J Jarvet, E Karelson, M Zilmer, K Iverfeldt, M Zorko, A Gräslund, U Langel (1999)  Regulation of GTPase and adenylate cyclase activity by amyloid beta-peptide and its fragments in rat brain tissue.   Brain Res 850: 1-2. 179-188 Dec  
Abstract: Modulation of GTPase and adenylate cyclase (ATP pyrophosphate-lyase, EC 4.6.1.1) activity by Alzheimer's disease related amyloid beta-peptide, A beta (1-42), and its shorter fragments, A beta (12-28), A beta (25-35), were studied in isolated membranes from rat ventral hippocampus and frontal cortex. In both tissues, the activity of GTPase and adenylate cyclase was upregulated by A beta (25-35), whereas A beta (12-28) did not have any significant effect on the GTPase activity and only weakly influenced adenylate cyclase. A beta (1-42), similar to A beta (25-35), stimulated the GTPase activity in both tissues and adenylate cyclase activity in ventral hippocampal membranes. Surprisingly, A beta (1-42) did not have a significant effect on adenylate cyclase activity in the cortical membranes. At high concentrations of A beta (25-35) and A beta (1-42), decreased or no activation of adenylate cyclase was observed. The activation of GTPase at high concentrations of A beta (25-35) was pertussis toxin sensitive, suggesting that this effect is mediated by Gi/G(o) proteins. Addition of glutathione and N-acetyl-L-cysteine, two well-known antioxidants, at 1.5 and 0.5 mM, respectively, decreased A beta (25-35) stimulated adenylate cyclase activity in both tissues. Lys-A beta (16-20), a hexapeptide shown previously to bind to the same sequence in A beta-peptide, and prevent fibril formation, decreased stimulation of adenylate cyclase activity by A beta (25-35), however, NMR diffusion measurements with the two peptides showed that this effect was not due to interactions between the two and that A beta (25-35) was active in a monomeric form. Our data strongly suggest that A beta and its fragments may affect G-protein coupled signal transduction systems, although the mechanism of this interaction is not fully understood.
Notes:
P Damberg, J Jarvet, P Allard, A Gräslund (1999)  Quantitative estimation of magnitude and orientation of the CSA tensor from field dependence of longitudinal NMR relaxation rates.   J Biomol NMR 15: 1. 27-37 Sep  
Abstract: A method is presented that makes it possible to estimate both the orientation and the magnitude of the chemical shift anisotropy (CSA) tensor in molecules with a pair of spin 1/2 nuclei, typically (13)C-(1)H or (15) N-(1)H. The method relies on the fact that the longitudinal cross-correlation rate as well as a linear combination of the autorelaxation rates of longitudinal heterospin magnetization, longitudinal two-spin order and longitudinal proton magnetization are proportional to the spectral density at the Larmor frequency of the heterospin. Therefore the ratio between the cross-correlation rate and the above linear combination is independent of the dynamics. From the field dependence of the ratio both the magnitude and the orientation of the CSA tensor can be estimated. The method is applicable to molecules in all motional regimes and is not limited to molecules in extreme narrowing or slow tumbling, nor is it sensitive to chemical exchange broadening. It is tested on the 22 amino acid residue peptide motilin, selectively (13) C labeled in the ortho positions in the ring of the single tyrosine residue. In the approximation of an axially symmetric (13)C CSA tensor, the symmetry axis of the CSA tensor makes an angle of 23 degrees +/- 1 degrees to the (13) C-(1)H bond vector, and has a magnitude of 156 +/- 5 ppm. This is in close agreement with solid-state NMR data on tyrosine powder [Frydman et al. (1992) Isr. J. Chem., 32, 161-164].
Notes:
1997
J Jarvet, J Zdunek, P Damberg, A Gräslund (1997)  Three-dimensional structure and position of porcine motilin in sodium dodecyl sulfate micelles determined by 1H NMR.   Biochemistry 36: 26. 8153-8163 Jul  
Abstract: The solution structure of the porcine gastrointestinal peptide hormone motilin was determined in the presence of sodium dodecyl sulfate (SDS) micelles at 28 degrees C using 1H nuclear magnetic resonance, full relaxation matrix analysis, and structure calculations based on restrained molecular dynamics. The structure of motilin in SDS micelles is described by a reverse gamma-turn and a beta-turn of type II in the N terminal end, an alpha-helical region in the middle of the molecule, and an extended structure at the C terminus. The position of the motilin molecule relative to the SDS micelles was probed by adding spin-labeled stearic acids, containing 12-doxyl or 5-doxyl spin-labels. We observed selective broadening of the proton resonances of residues 3-5 and concluded that they must be located in the interior of the micelle. These experiments suggest a structural model in which the hydrophobic N terminus consists of two well-defined turns buried in the interior of the micelle, whereas the amphiphilic alpha-helical part is located at the surface of the micelle. Spectral density mapping using a 13C label on the alphaC of Leu10 gave overall rotational correlation times taum of 6.6 and 4.5 ns at 35 and 45 degrees C, respectively. The long correlation time in combination with a high order parameter (S = 0.92) indicates that motilin has a rigid structure in the complex with the SDS micelle.
Notes:
1996
Jarvet, Allard (1996)  Phase-Sensitive Two-Dimensional Heteronuclear Zero- and Double-Quantum-Coherence Spectroscopy   J Magn Reson B 112: 3. 240-244 Sep  
Abstract: New pulse sequences for two-dimensional phase-sensitive detection of heteronuclear zero- and double-quantum coherence are presented for use with and without pulsed field gradients. The magnetization is phase modulated during t1 and then transferred to a proton for detection using a novel approach similar to the sensitivity-enhancement technique [Palmer et al. J. Magn. Reson. 93, 151 (1991)]. These pulse sequences are useful for the measurement of size and relative sign of passive J couplings as well as for the measurement of the relaxation rates of zero- and double-quantum coherence. They can also be used as building blocks in multidimensional pulse sequences. The pulse sequences are tested on the peptide hormone motilin with a selectively 13C-enriched alpha carbon.
Notes:
Jarvet, Allard, Ehrenberg, Graslund (1996)  Spectral-Density Mapping of 13Calpha-1Halpha Vector Dynamics Using Dipolar Relaxation Rates Measured at Several Magnetic Fields   J Magn Reson B 111: 1. 23-30 Apr  
Abstract: The spectral-density mapping of a 13Calpha-1Halpha vector of Leu10 in the 22-residue peptide hormone motilin [P. Allard, J. Jarvet, A. Ehrenberg, and A. Graslund, J. Biomol. NMR 5, 133-146 (1995)] is extended in this paper to three polarizing fields 9.4, 11.7, and 14.1 T in order to improve the accuracy of the calculated spectral-density function J(omega) and to extend the sampling range up to 750 MHz. The problem with a usually large relative error in J(omegaH) is eliminated since the generally more precise J(omegaH - omegaC) and J(omegaH + omegaC) determined at other fields appear at nearly the same frequencies. The fitting of dynamic models to the points of spectral density was made with error weighting, and the influence of J(omegaH) was found to be negligible. Therefore, the high-frequency part of the spectral-density function is determined essentially without influence from the two transverse-type relaxation rates. In the case of a carbon-proton vector, the relaxation is mainly determined by dipolar interaction and is only weakly influenced by other relaxation mechanisms, which makes it particularly suitable for the spectral-density mapping technique. The measured relaxation rates in the time domain are transformed into the frequency domain by spectral-density mapping, and the slopes in different frequency regions are important parameters when comparing experimental data with theoretical models of motion. Using an adjustable internuclear distance reff, combined with the model-free approach, it is possible to obtain a reasonable fit to measured spectral-density points at J(0) and around J(omegaC). At the same time, however, the high-frequency slope of the spectral-density function defined by J(omegaH - omegaC) and J(omegaH + omegaC) could not be reproduced.
Notes:
1995
P Allard, J Jarvet, A Ehrenberg, A Gräslund (1995)  Mapping of the spectral density function of a C alpha-H alpha bond vector from NMR relaxation rates of a 13C-labelled alpha-carbon in motilin.   J Biomol NMR 5: 2. 133-146 Feb  
Abstract: The peptide hormone motilin was synthesised with a 13C-enriched alpha-carbon in the leucine at position 10. In aqueous solution, six different relaxation rates were measured for the 13C alpha-H alpha fragment as a function of temperature and with and without the addition of 30% (v/v) of the cosolvent d2-1,1,1,3,3,3-hexafluoro-2-propanol (HFP). The relaxation rates were analysed employing the spectral density mapping technique introduced by Peng and Wagner [(1992) J. Magn. Reson., 98, 308-332] and using the model-free approach by Lipari and Szabo [(1982) J. Am. Chem. Soc., 104, 4546-4570]. The fit to various models of dynamics was also considered. Different procedures to evaluate the overall rotational correlation time were compared. A single exponential time correlation function was found to give a good fit to the measured spectral densities only for motilin in 30% (v/v) HFP at low temperatures, whereas at high temperatures in this solvent, and in D2O at all temperatures, none of the considered models gave an acceptable fit. A new empirical spectral density function was tested and found to accurately fit the experimental spectral density mapping points. The application of spectral density mapping based on NMR relaxation data for a specific 13C-1H vector is shown to be a highly useful method to study biomolecular dynamics. Advantages are high sensitivity, high precision and uniform sampling of the spectral density function over the frequency range.
Notes:
1994
K Arvidsson, J Jarvet, P Allard, A Ehrenberg (1994)  Solution structure by 1H and dynamics by natural abundance 13C NMR of a receptor recognising peptide derived from a C-terminal fragment of neuropeptide Y.   J Biomol NMR 4: 5. 653-672 Sep  
Abstract: A peptide consisting of 20 amino acid residues, derived from a C-terminal fragment of neuropeptide Y (NPY) and showing high affinity to NPY receptors, was synthesised. Its sequence is PAADLARYRHYINLITRQRY-NH2, and the solution structure was calculated from NMR-derived distance and torsion angle restraints, obtained at 15 degrees C in a solvent mixture of water and 30% (v/v) 1,1,1,3,3,3-hexafluoro-2-propanol, by using DIANA and restrained energy minimisation. The structure was found to consist of a well-defined alpha-helix in the centre, with a few residues at the termini having less well defined conformations. The spin-lattice and spin-spin relaxation rates of alpha-carbons have been determined on 13C at natural abundance. From 1D experiments the global rotational correlation time was determined and from 2D experiments the dynamics of each individual residue was obtained. The results demonstrate that the C alpha-H alpha vectors in the alpha-helix essentially follow the global motion. Towards the termini, contributions from local dynamics increase. This tendency is correlated to the increasing uncertainty of the structure towards the peptide ends. An effective molecular volume was calculated from the temperature dependence of the global rotational correlation time. This is well compatible with a monomeric peptide, solvated by water and 1,1,1,3,3,3-hexafluoro-2-propanol. The presence of peptide dimers was ruled out as being inconsistent with the relaxation data.
Notes:
1989
J Jarvet, A Olivson, U Mets, M Pooga, R Aguraiuja, E Lippmaa (1989)  13C and 15N NMR and time-resolved fluorescence depolarization study of bovine carbonic anhydrase--4-methylbenzenesulfonamide complex.   Eur J Biochem 186: 1-2. 287-290 Dec  
Abstract: The influence of the binding of the high-affinity inhibitor, 4-methylbenzenesulfonamide, to the active site of bovine carbonic anhydrase B was studied by 15N- and 13C-NMR spectroscopy. The rotational correlation time dependence on temperature and concentration of the complex was determined by time-resolved fluorescence depolarization measurements. Our experiment provides evidence that the stoichiometry of the interaction of 4-methylbenzenesulfonamide with carbonic anhydrase B is 1:1 and the inhibitor is bound in anionic form. The 15N-NMR relaxation parameters confirm our previous conclusions about the presence of librational motions in the active site of carbonic anhydrase and indicate that the internal motion in the enzyme-inhibitor complex is more restricted than the backbone motion in the uncomplexed native enzyme.
Notes:
Powered by PublicationsList.org.