Your browser doesn't support javascript.
loading
Show: 20 | 50 | 100
Results 1 - 8 de 8
Filter
Add more filters










Database
Language
Publication year range
1.
Allergy ; 71(5): 621-8, 2016 05.
Article in English | MEDLINE | ID: mdl-26608594

ABSTRACT

BACKGROUND: Allergic reactions to mosquito bites are an increasing clinical concern. Due to the lack of availability of mosquito salivary allergens, they are underdiagnosed. Here, we reported a newly cloned mosquito Aedes (Ae.) aegypti salivary allergen. METHODS: A cDNA encoding a 30-kDa Ae. aegypti salivary protein, designated Aed a 3, was isolated from an expression library. The full-length cDNA was cloned into a baculovirus expression vector, and recombinant Aed a 3 (rAed a 3) was expressed, purified, and characterized. Skin prick tests with purified rAed a 3 and Ae. aegypti bite tests were performed in 43 volunteers. Serum rAed a 3-specific IgE levels were measured in 28 volunteers. RESULTS: The primary nucleotide sequence, deduced amino acid sequence, and IgE-binding sites of Aed a 3 were identified. rAed a 3-selected antibodies recognized a 30-kDa Ae. aegypti saliva protein. rAed a 3 bound IgE in mosquito-allergic volunteers and the binding could be inhibited by the addition of natural mosquito extract dose dependently. Immediate skin test reactions to rAed a 3 correlated significantly with mosquito bite-induced reactions. Of the bite test-positive volunteers, 32% had a positive rAed a 3 skin test and 46% had specific IgE. No bite test-negative volunteers reacted to rAed a 3 in either the skin tests or the IgE assays, confirming the specificity of the assay. CONCLUSIONS: Aed a 3 that corresponds to the Aegyptin protein is a major mosquito salivary allergen. Its recombinant form has biological activity and is suitable for use in skin tests and specific IgE assays in mosquito-allergic individuals.


Subject(s)
Allergens/genetics , Allergens/immunology , Insect Proteins/genetics , Insect Proteins/immunology , Salivary Proteins and Peptides/genetics , Salivary Proteins and Peptides/immunology , Adult , Aedes , Allergens/chemistry , Amino Acid Sequence , Animals , Cloning, Molecular , DNA, Complementary , Enzyme-Linked Immunosorbent Assay , Female , Humans , Immunoglobulin E/immunology , Insect Bites and Stings/diagnosis , Insect Bites and Stings/immunology , Insect Proteins/chemistry , Models, Molecular , Protein Binding , Protein Conformation , Recombinant Proteins/chemistry , Recombinant Proteins/genetics , Recombinant Proteins/immunology , Salivary Proteins and Peptides/chemistry , Sensitivity and Specificity , Sequence Analysis, DNA , Skin Tests
2.
Proteins ; 44(4): 418-27, 2001 Sep 01.
Article in English | MEDLINE | ID: mdl-11484219

ABSTRACT

Traditionally, molecular disorder has been viewed as local or global instability. Molecules or regions displaying disorder have been considered inherently unstructured. The term has been routinely applied to cases for which no atomic coordinates can be derived from crystallized molecules. Yet, even when it appears that the molecules are disordered, prevailing conformations exist, with population times higher than those of all alternate conformations. Disordered molecules are the outcome of rugged energy landscapes away from the native state around the bottom of the funnel. Ruggedness has a biological function, creating a distribution of structured conformers that bind via conformational selection, driving association and multimolecular complex formation, whether chain-linked in folding or unlinked in binding. We classify disordered molecules into two types. The first type possesses a hydrophobic core. Here, even if the native conformation is unstable, it still has a large enough population time, enabling its experimental detection. In the second type, no such hydrophobic core exists. Hence, the native conformations of molecules belonging to this category have shorter population times, hindering their experimental detection. Although there is a continuum of distribution of hydrophobic cores in proteins, an empirical, statistically based hydrophobicity function may be used as a guideline for distinguishing the two disordered molecule types. Furthermore, the two types relate to steps in the protein folding reaction. With respect to protein design, this leads us to propose that engineering-optimized specific electrostatic interactions to avoid electrostatic repulsion would reduce the type I disordered state, driving the molten globule (MG) --> native (N) state. In contrast, for overcoming the type II disordered state, in addition to specific interactions, a stronger hydrophobic core is also indicated, leading to the denatured --> MG --> N state.


Subject(s)
Protein Folding , Proteins/chemistry , Enzyme Stability , Models, Molecular , Protein Conformation , Protein Denaturation , Protein Renaturation , Proteins/metabolism , Structure-Activity Relationship , Thermodynamics
3.
Biophys J ; 80(5): 2439-54, 2001 May.
Article in English | MEDLINE | ID: mdl-11325743

ABSTRACT

Three-dimensional protein folds range from simple to highly complex architectures. In complex folds, some building block fragments are more important for correct protein folding than others. Such fragments are typically buried in the protein core and mediate interactions between other fragments. Here we present an automated, surface area-based algorithm that is able to indicate which, among all local elements of the structure, is critical for the formation of the native fold, and apply it to structurally well-characterized proteins. In particular, we focus on adenylate kinase. The fragment containing the phosphate binding, P-loop (the "giant anion hole") flanked by a beta-strand and an alpha-helix near the N-terminus, is identified as a critical building block. This building block shows a high degree of sequence and structural conservation in all adenylate kinases. The results of our molecular dynamics simulations are consistent with this identification. In its absence, the protein flips to a stable, non-native state. In this misfolded conformation, the other local elements of the structure are in their native-like conformations; however, their association is non-native. Furthermore, this element is critically important for the function of the enzyme, coupling folding, and function.


Subject(s)
Adenylate Kinase/chemistry , Algorithms , Amino Acid Motifs , Amino Acid Sequence , Anions , Biophysical Phenomena , Biophysics , Computer Simulation , Conserved Sequence , Dimerization , Fungal Proteins/chemistry , Models, Molecular , Models, Theoretical , Molecular Sequence Data , Peptides/chemistry , Phosphates , Phylogeny , Protein Binding , Protein Conformation , Protein Folding , Protein Structure, Secondary , Protein Structure, Tertiary , Sequence Homology, Amino Acid , Software , Time Factors
4.
Protein Sci ; 10(1): 135-48, 2001 Jan.
Article in English | MEDLINE | ID: mdl-11266602

ABSTRACT

We have carried out molecular dynamics simulations of the native dihydrofolate reductase from Escherichia coli and several of its folded protein fragments at standard temperature. The simulations have shown fragments 1--36, 37--88, and 89--159 to be unstable, with a C(alpha)RMSD (C(alpha) root mean squared deviation) >5 A after 3.0 nsec of simulation. The unfolding of fragment 1--36 was immediate, whereas fragments 37--88 and 89--159 gradually unfolded because of the presence of the beta-sheet core structure. In the absence of residues 1--36, the two distinct domains comprising fragment 39--159 associated with each other, resulting in a stable conformation. This conformation retained most of its native structural elements. We have further simulated fragments derived from computational protein cutting. These were also found to be unstable, with the exception of fragment 104--159. In the absence of alpha(4), the loose loop region of residues 120--127 exhibited a beta-strand-like behavior, associating itself with the beta-sheet core of the protein fragment. The current study suggests that the folding of dihydrofolate reductase involves cooperative folding of distinct domains which otherwise would have been unstable as independent folded units in solution. Finally, the critical role of residues 1--36 in allowing the two distinct domains of fragment 104--159 to fold into the final native conformation is discussed.


Subject(s)
Escherichia coli/enzymology , Tetrahydrofolate Dehydrogenase/chemistry , Computer Simulation , Enzyme Stability , Models, Molecular , Molecular Chaperones/chemistry , Peptide Fragments/chemistry , Peptide Fragments/metabolism , Peptide Hydrolases/metabolism , Protein Conformation , Protein Folding , Tetrahydrofolate Dehydrogenase/metabolism
5.
Proc Natl Acad Sci U S A ; 97(22): 11899-904, 2000 Oct 24.
Article in English | MEDLINE | ID: mdl-11050223

ABSTRACT

The idea that enzymes accelerate their reactions by entropic effects has played a major role in many prominent proposals about the origin of enzyme catalysis. This idea implies that the binding to an enzyme active site freezes the motion of the reacting fragments and eliminates their entropic contributions, (delta S(cat)(double dagger))', to the activation energy. It is also implied that the binding entropy is equal to the activation entropy, (delta S(w)(double dagger))', of the corresponding solution reaction. It is, however, difficult to examine this idea by experimental approaches. The present paper defines the entropic proposal in a rigorous way and develops a computer simulation approach that determines (delta S(double dagger))'. This approach allows us to evaluate the differences between (delta S(double dagger))' of an enzymatic reaction and of the corresponding reference reaction in solution. Our approach is used in a study of the entropic contribution to the catalytic reaction of subtilisin. It is found that this contribution is much smaller than previously thought. This result is due to the following: (i) Many of the motions that are free in the reactants state of the reference solution reaction are also free at the transition state. (ii) The binding to the enzyme does not completely freeze the motion of the reacting fragments so that (delta S(double dagger))' in the enzymes is not zero. (iii) The binding entropy is not necessarily equal to (delta S(w)(double dagger))'.


Subject(s)
Entropy , Enzymes/metabolism , Catalysis , Dipeptides/metabolism
6.
Proteins ; 39(4): 393-407, 2000 Jun 01.
Article in English | MEDLINE | ID: mdl-10813821

ABSTRACT

Several strategies for evaluation of the protein-ligand binding free energies are examined. Particular emphasis is placed on the Linear Response Approximation (LRA) (Lee et. al., Prot Eng 1992;5:215-228) and the Linear Interaction Energy (LIE) method (Aqvist et. al., Prot Eng 1994;7:385-391). The performance of the Protein Dipoles Langevin Dipoles (PDLD) method and its semi-microscopic version (the PDLD/S method) is also considered. The examination is done by using these methods in the evaluating of the binding free energies of neutral C2-symmetric cyclic urea-based molecules to Human Immunodeficiency Virus (HIV) protease. Our starting point is the introduction of a thermodynamic cycle that decomposes the total binding free energy to electrostatic and non-electrostatic contributions. This cycle is closely related to the cycle introduced in our original LRA study (Lee et. al., Prot Eng 1992;5:215-228). The electrostatic contribution is evaluated within the LRA formulation by averaging the protein-ligand (and/or solvent-ligand) electrostatic energy over trajectories that are propagated on the potentials of both the polar and non-polar (where all residual charges are set to zero) states of the ligand. This average involves a scaling factor of 0.5 for the contributions from each state and this factor is being used in both the LRA and LIE methods. The difference is, however, that the LIE method neglects the contribution from trajectories over the potential of the non-polar state. This approximation is entirely valid in studies of ligands in water but not necessarily in active sites of proteins. It is found in the present case that the contribution from the non-polar states to the protein-ligand binding energy is rather small. Nevertheless, it is clearly expected that this term is not negligible in cases where the protein provides preorganized environment to stabilize the residual charges of the ligand. This contribution can be particularly important in cases of charged ligands. The analysis of the non-electrostatic term is much more complex. It is concluded that within the LRA method one has to complete the relevant thermodynamic cycle by evaluating the binding free energy of the "non-polar" ligand, l;, where all the residual charges are set to zero. It is shown that the LIE term, which involves the scaling of the van der Waals interaction by a constant beta (usually in the order of 0.15 to 0.25), corresponds to this part of the cycle. In order to elucidate the nature of this non-electrostatic term and the origin of the scaling constant beta, it is important to evaluate explicitly the different contributions to the binding energy of the non-polar ligand, DeltaG(bind,l;). Since this cannot be done at present (for relatively large ligands) by rigorous free energy perturbation approaches, we evaluate DeltaG(bind,l;) by the PDLD approach, augmented by microscopic calculations of the change in configurational entropy upon binding. This evaluation takes into account the van der Waals, hydrophobic, water penetration and entropic contributions, which are the most important free energy contributions that make up the total DeltaG(bind,l;). The sum of these contributions is scaled by a factor straight theta and it is argued that obtaining a quantitative balance between these contributions should result in straight theta = 1. By doing so we should have a reliable estimate of the value of the LIE beta and a way to understand its origin. The present approach gives straight theta values between 0.5 and 0.73, depending on the approximation used. This is encouraging but still not satisfying. Nevertheless, one might be able to use our PDLD approach to estimate the change of the LIE straight theta between different protein active sites. It is pointed out that the LIE method is quite similar to our original approach where the electrostatic term was evaluated by the LRA method and the non-electrostatic term by the PDLD method (with its vdw, solvation,


Subject(s)
HIV Protease/metabolism , Energy Transfer , Humans , Ligands , Linear Models , Mathematical Computing , Models, Molecular
7.
Proteins ; 36(4): 484-500, 1999 Sep 01.
Article in English | MEDLINE | ID: mdl-10450091

ABSTRACT

A general method for simulating proton translocations in proteins and for exploring the role of different proton transfer pathways is developed and examined. The method evaluates the rate constants for proton transfer processes using the energetics of the relevant proton configurations. The energies (DeltaG((m))) of the different protonation states are evaluated in two steps. First, the semimicroscopic version of the protein dipole Langevin dipole (PDLD/S) method is used to evaluate the intrinsic energy of bringing the protons to their protein sites, when the charges of all protein ionized residues are set to zero. Second, the interactions between the charged groups are evaluated by using a Coulomb's Law with an effective dielectric constant. This approach, which was introduced in an earlier study by one of the authors of the current report, allows for a very fast determination of any DeltaG((m)) and for practical evaluation of the time-dependent proton population: That is, the rate constants for proton transfer processes are evaluated by using the corresponding DeltaG((m)) values and a Marcus type relationship. These rate constants are then used to construct a master equation, the integration of which by a fourth-order Runge-Kutta method yields the proton population as a function of time. The integration evaluates, 'on the fly,' the changes of the rate constants as a result of the time-dependent changes in charge-charge interaction, and this feature benefits from the fast determination of DeltaG((m)). The method is demonstrated in a preliminary study of proton translocation processes in the reaction center of Rhodobacter sphaeroides. It is found that proton transfer across water chains involves significant activation barriers and that ionized protein residues probably are involved in the proton transfer pathways. The potential of the present method in analyzing mutation experiments is discussed briefly and illustrated. The present study also examines different views of the nature of proton translocations in proteins. It is shown that such processes are controlled mainly by the electrostatic interaction between the proton site and its surroundings rather than by the local bond rearrangements of water molecules that are involved in the proton pathways. Thus, the overall rate of proton transport frequently is controlled by the highest barrier along the conduction pathway. Proteins 1999;36:484-500.


Subject(s)
Computer Simulation , Photosynthetic Reaction Center Complex Proteins/metabolism , Protons , Rhodobacter sphaeroides/metabolism , Electron Transport , Ion Transport , Kinetics , Models, Chemical , Models, Molecular , Photosynthetic Reaction Center Complex Proteins/chemistry , Protein Conformation , Proton Pumps/chemistry , Proton Pumps/metabolism , Rhodobacter sphaeroides/chemistry , Solvents , Static Electricity , Thermodynamics , Water/chemistry , Water/metabolism
8.
Biophys J ; 74(4): 1744-53, 1998 Apr.
Article in English | MEDLINE | ID: mdl-9545037

ABSTRACT

The effect of the reorganization of the protein polar groups on charge-charge interaction and the corresponding effective dielectric constant (epsilon(eff)) is examined by the semimicroscopic version of the Protein Dipole Langevin Dipoles (PDLD/S) method within the framework of the Linear Response Approximation (LRA). This is done by evaluating the interactions between ionized residues in the reaction center of Rhodobacter sphaeroides, while taking into account the protein reorganization energy. It is found that an explicit consideration of the protein relaxation leads to a significant increase in epsilon(eff) and that semimicroscopic models that do not take this relaxation into account force one to use a large value for the so-called "protein dielectric constant," epsilon(p), of the Poisson-Boltzmann model or for the corresponding epsilon(in) in the PDLD/S model. An additional increase in epsilon(eff) is expected from the reorganization of ionized residues and from changes in the degree of water penetration. This finding provides further support for the idea that epsilon(in) (or epsilon(p)) represents contributions that are not considered explicitly. The present study also provides a systematic illustration of the nature of epsilon(eff), supporting our previously reported view that charge-charge interactions correspond to a large value of this "dielectric constant," even in protein interiors. It is also pointed out that epsilon(eff) for the interaction between ionizable groups in proteins is very different from the effective dielectric constant, epsilon'(eff), that determines the free energy of ion pairs in proteins (epsilon'(eff) reflects the effect of preoriented protein dipoles). Finally, the problems associated with the search for a general epsilon(in) are discussed. It is clarified that the epsilon(in) that reproduces the effect of protein relaxation on charge-charge interaction is not equal to the epsilon(in) that reproduces the corresponding effect upon formation of individual charges. This reflects fundamental inconsistencies in attempts to cast microscopic concepts in a macroscopic model. Thus one should either use a large epsilon(in) for charge-charge interactions and a small epsilon(in) for charge-dipole interactions or consider the protein relaxation microscopically.


Subject(s)
Proteins/chemistry , Binding Sites , Biophysical Phenomena , Biophysics , Ions , Models, Chemical , Photosynthetic Reaction Center Complex Proteins/chemistry , Rhodobacter sphaeroides/chemistry , Static Electricity , Thermodynamics
SELECTION OF CITATIONS
SEARCH DETAIL
...