Big Chemical Encyclopedia

Chemical substances, components, reactions, process design ...

Articles Figures Tables About

Tertiary structure in protein

One of the best known examples of reversibility in bond formation is the cross-linking of cysteine, a sulfur-containing amino acid, that affects tertiary structure in proteins and, ultimately, macroscale phenomena such as the degree of curl in hair. Other examples include the imine bond, formed by the reaction of an amine group with an aldehyde, and metal coordinate bonds to atoms such as nitrogen as found in many enzymes. [Pg.9]

It is helpful to know the chemistry of fixatives in order to understand their action and avoid artifacts (4). Most commonly studied antigens are either proteins or carbohydrates. Many of these molecules are soluble in aqueous solutions and need to be fixed in place in cells. Insoluble antigens also need to be structurally preserved (1). All chemical fixatives will cause chemical and conformational changes in the protein structure of cells with lesser changes noted for carbohydrate antigens (5). Secondary and tertiary structures in proteins are the most important for eliciting antigenicity, and chemical fixatives usually disturb these structures (3). [Pg.56]

Three-dimensional tertiary structure in proteins is maintained by ionic bonds, hydrogen bonds, -S-S- bridges, van der Waals forces, and hydrophobic interactions. [Pg.57]

O What are the differences among primary, secondary, and tertiary structure in proteins ... [Pg.761]

In extracellular proteins, the oxidizing environment leads to formation of disulfide bonds whenever two free cysteine side chains can assume the appropriate configuration. Not surprisingly, the range of tertiary structures in proteins with disulfide bonds is more varied than that normally observed for intracellular proteins. [Pg.124]

Protein tertiar-y structure is also influenced by the environment. In water a globular- protein usually adopts a shape that places its hydrophobic groups toward the interior, with its polar- groups on the surface, where they are solvated by water molecules. About 65% of the mass of most cells is water, and the proteins present in cells are said to be in their native state—the tertiary structure in which they express their biological activity. When the tertiar-y structure of a protein is disrupted by adding substances that cause the protein chain to unfold, the protein becomes denatured and loses most, if not all, of its activity. Evidence that supports the view that the tertiary structure is dictated by the primary structure includes experiments in which proteins are denatured and allowed to stand, whereupon they are observed to spontaneously readopt then native-state conformation with full recovery of biological activity. [Pg.1146]

Section 28.9 Within the cell nucleus, double-helical DNA adopts a supercoiled tertiary structure in which short sections are wound around proteins called histones. This reduces the effective length of the DNA and maintains it in an ordered anangement. [Pg.1188]

When the polypeptide chains of protein molecules bend and fold in order to assume a more compact three-dimensional shape, a tertiary (3°) level of structure is generated (Figure 5.9). It is by virtue of their tertiary structure that proteins adopt a globular shape. A globular conformation gives the lowest surface-to-volume ratio, minimizing interaction of the protein with the surrounding environment. [Pg.118]

Role of the Amino Acid Sequence in Protein Structure Secondary Structure in Protein.s Protein Folding and Tertiary Structure Subunit Interaction.s and Quaternary Structure... [Pg.158]

Figure 5-6. Examples of tertiary structure of proteins. Top The enzyme triose phosphate isomerase. Note the elegant and symmetrical arrangement of alternating p sheets and a helices. (Courtesy of J Richardson.) Bottom Two-domain structure of the subunit of a homodimeric enzyme, a bacterial class II HMG-CoA reductase. As indicated by the numbered residues, the single polypeptide begins in the large domain, enters the small domain, and ends in the large domain. (Courtesy ofC Lawrence, V Rod well, and C Stauffacher, Purdue University.)... Figure 5-6. Examples of tertiary structure of proteins. Top The enzyme triose phosphate isomerase. Note the elegant and symmetrical arrangement of alternating p sheets and a helices. (Courtesy of J Richardson.) Bottom Two-domain structure of the subunit of a homodimeric enzyme, a bacterial class II HMG-CoA reductase. As indicated by the numbered residues, the single polypeptide begins in the large domain, enters the small domain, and ends in the large domain. (Courtesy ofC Lawrence, V Rod well, and C Stauffacher, Purdue University.)...
The extrusion process frequently results in realignment of disulfide bonds and breakage of intramolecular bonds. Disulfide bonds stabilize the tertiary structure of protein and may limit protein imfolding during extrusion (Taylor et al., 2006). Flow and melt characteristics were improved when other proteins were extruded with disulfide reducing agents (Areas, 1992), which indicates that disulfide bonds adversely affect... [Pg.181]

However, 2D NOE studies are invaluable in structure determination, in particular of peptides and proteins here the NOEs give invaluable information for conformational analysis and the determination of the tertiary structures of proteins. [Pg.42]

The denaturation of proteins generally involves at least partial unfolding, with the loss of secondary and tertiary structure. In the present context, we are interested in the end point of this process — proteins that are unfolded to the maximal extent by various agents heat, cold, acid, urea, Gdm-HCl.1 Three major questions concerning unfolded proteins are of interest in the present chapter. Do different unfolding agents... [Pg.221]

From the atomic to the macroscopic level chirality is a characteristic feature of biological systems and plays an important role in the interplay of structure and function. Originating from small chiral precursors complex macromolecules such as proteins or DNA have developed during evolution. On a supramolecular level chirality is expressed in molecular organization, e.g. in the secondary and tertiary structure of proteins, in membranes, cells or tissues. On a macroscopic level, it appears in the chirality of our hands or in the asymmetric arrangement of our organs, or in the helicity of snail shells. Nature usually displays a preference for one sense of chirality over the other. This leads to specific interactions called chiral recognition. [Pg.135]

Fournier and DePristo96 calculated bond energies in several small compounds containing disulfide bonds which are known to stabilize the tertiary structure of proteins. Bond dissociation energies are generally overestimated when LDA(SVWN) is used whereas the PW86/P86 functional brings them to within 5 kcal/mol of experimental values. [Pg.97]

Evolutionary processes driven by environmental changes and varying conditions have an impact on all components in a living cell. Thus, the primary, secondary and tertiary structure of proteins determines their function and location, giving different properties in different compartments, such as outer membrane, periplasmic space, cytoplasmic membrane or cytoplasm. Proteins can function as monomers or oligomers and can occur in a soluble form, as integral constituents embedded within the membrane, or can be found associated with the lipid bilayer itself or components therein. [Pg.278]

The normal cellular form of prion protein (PrPc) can exist as a Cu-metalloprotein in vivo (492). This PrPc is a precursor of the pathogenic protease-resistant form PrPsc, which is thought to cause scrapie, bovine spongiform encephalopathy (BSE), and Creutzfeldt—Jakob disease. Two octa-repeats of PHGGGWGQ have been proposed as Cu(II) binding sites centered on histidine (493). They lack secondary and tertiary structure in the absence of Cu(II). Neurons may therefore have special mechanisms to regulate the distribution of copper. [Pg.264]

The dependence of the residual dipolar coupling on the angle that the vector forms with a reference axis explains why the use of dipolar couplings makes possible the determination of the relative orientation of different domains in a multidomain protein and facilitates nucleic acid structure determination. Dipolar couplings can constitute up to 50% of the total structural data available for nucleic acids, while this number drops to 10-15% in proteins. Thus, the impact of the use of dipolar couplings on the structure determination of nucleic acids is generally more substantial than in the case of proteins. Furthermore, the presence or absence of tertiary structure in a protein or nucleic acid does not have a major influence on the number of dipolar couplings that can be measured, in contrast to the case of the NOE. [Pg.181]

Gilmanshin R., Williams S., Callender R. H., Woodruff W. H. and Dyer R. B. Fast events in protein folding relaxation dynamics of secondary and tertiary structure in native apomyoglobin. Proc. Natl. Acad. Sci., USA (1997) 94(8) 3709-3713. [Pg.99]

The influence of secondary structure on reactions of deamidation has been confirmed in a number of studies. Thus, deamidation was inversely proportional to the extent of a-helicity in model peptides [120], Similarly, a-hel-ices and /3-turns were found to stabilize asparagine residues against deamidation, whereas the effect of /3-sheets was unclear [114], The tertiary structure of proteins is also a major determinant of chemical stability, in particular against deamidation [121], on the basis of several factors such as the stabilization of elements of secondary structure and restrictions to local flexibility, as also discussed for the reactivity of aspartic acid residues (Sect. 6.3.3). Furthermore, deamidation is markedly decreased in regions of low polarity in the interior of proteins because the formation of cyclic imides (Fig. 6.29, Pathway e) is favored by deprotonation of the nucleophilic backbone N-atom, which is markedly reduced in solvents of low polarity [100][112],... [Pg.324]

The PMR spectrum of protein SI suggests that the protein has considerable tertiary structure in physiological buffer and is more flexible than normal globular proteins of its molecular weight (Moore and Laughrea, 1979). No difference was observed when the protein was prepared in the presence or absence of urea at neutral pH. The spectra obtained in this study resemble those previously obtained with salt-extracted SI by Little-child and Malcolm (1978). [Pg.13]

Calorimetric studies have been made on proteins S4, S7, S8, S15, S16, S18, Lll, and L7 (Khechinashvili et al., 1978 Gudkov and Behike, 1978). Most of these proteins displayed a cooperative tertiary structure in solution. Proteins S4, S7, SI5, and SI8 were extracted from the ribosome by a urea-LiCl technique followed by renaturation, whereas proteins S8, S16, and Lll were prepared by the mild isolation method. A calorimetric study on protein SI showed a noncooperative transition around 70-80 C, suggesting a flexible tertiary structure (L. Giri, unpublished). [Pg.14]


See other pages where Tertiary structure in protein is mentioned: [Pg.14]    [Pg.289]    [Pg.599]    [Pg.91]    [Pg.658]    [Pg.224]    [Pg.392]    [Pg.96]    [Pg.552]    [Pg.14]    [Pg.289]    [Pg.599]    [Pg.91]    [Pg.658]    [Pg.224]    [Pg.392]    [Pg.96]    [Pg.552]    [Pg.184]    [Pg.161]    [Pg.6]    [Pg.103]    [Pg.128]    [Pg.36]    [Pg.231]    [Pg.91]    [Pg.568]    [Pg.25]    [Pg.28]    [Pg.31]    [Pg.49]    [Pg.164]    [Pg.41]    [Pg.472]    [Pg.292]    [Pg.197]   
See also in sourсe #XX -- [ Pg.392 , Pg.393 , Pg.394 , Pg.395 , Pg.396 , Pg.397 , Pg.398 , Pg.399 , Pg.400 , Pg.401 , Pg.402 ]




SEARCH



Protein tertiary

Protein tertiary structure

Secondary and tertiary restraints in assembly of protein structures

Structures Tertiary structure

Tertiary structure

© 2024 chempedia.info