scispace - formally typeset
Search or ask a question

Showing papers in "Biopolymers in 1974"


Journal ArticleDOI
TL;DR: In this article, a method for determining chemical kinetic constants and diffusion coefficients by measuring the rates of decay of spontaneous concentration fluctuations was proposed, which is based on the principle that the rate of spontaneous microscopic fluctuations are determined by the same phenomenological rate coefficients as those of macroscopic departures from equilibrium which result from external perturbations.
Abstract: We describe a method for determining chemical kinetic constants and diffusion coefficients by measuring the rates of decay of spontaneous concentration fluctuations. The equilibrium of the system is not disturbed during the measurement. We measure the number of molecules of a specified type in a defined open volume as a function of time and compute the time course of the deviations from the thermodynamic mean concentration. The method is based on the principle that the rates of decay of spontaneous microscopic fluctuations are determined by the same phenomenological rate coefficients as those of macroscopic departures from equilibrium which result from external perturbations. Hence, an analysis of fluctuations yields the same chemical rate constants and diffusion coefficients as are measured by conventional procedures. In practice the number of the specified molecules is measured by a property such as absorbance or fluorescence which is specific and sensitive to chemical change. The sample volume is defined by a light beam which traverses the cell. As the molecules appear in or disappear from the light beam, either due to diffusion or chemical reaction, their concentration fluctuations give rise to corresponding fluctuations of the intensity of absorbed or emitted light. This paper presents the theory needed to derive chemical rate constants and diffusion coefficients from these fluctuations in light intensity. The theory is applied to three examples of general interest: pure diffusion in the absence of chemical reaction; the binding of a small rapidly diffusing ligand to a larger slowly diffusing macromolecule; and a unimolecular isomerization. The method should be especially useful in studying highly cooperative systems, relatively noncooperative systems with intermediate states closely spaced in free energy, small systems, and systems not readily subject to perturbations of state.

1,373 citations


Journal ArticleDOI
TL;DR: This paper describes the first experimental application of fluorescence correlation spectroscopy, a new method for determining chemical kinetic constants and diffusion coefficients by observing the time behaviour of the tiny concentration fluctuations which occur spontaneously in the reaction system even when it is in equilibrium.
Abstract: synopsis This paper describes the first experimental application of fluorescence correlation spectroscopy, a new method for determining chemical kinetic constants and diffusion coefficients. These quantities are measured by observing the time behavior of the tiny concentration fluctuations which occur spontaneously in the reaction system even when it is in equilibrium. The equilibrium of the system is not disturbed during the experiment. The diffusion coefficients and chemical rate constants which determine the average time behavior of these spontaneous fluctuations are the same as those sought by more conventional methods including temperature-jump or other perturbation tecliniques. The experiment consists essentially in measuring the variation with time of the number of molecules of specified reactants in a defined open volume of solution. The concentration of a reactant is measured by its fluorescence; the sample volume is defined by a focused laser beam which excites the fluorescence. The fluorescent emission fluctuates in proportion with the changes in the number of fluorescent molecules as they diffuse into and out of the sample volume and as they are created or eliminated by the chemical reactions. The number of these reactant molecules must be small to permit detection of the concentration fluctuations. Hence the sample volume is small (10-8 rnl) and the concentration of the solutes is low (-10-9 M). We have applied this technique to the study of two prototype systems: the simple example of pure diffusion of a single fluorescent species, rhodamine 6G, and the more interesting but more challenging example of the reaction of macromolecular DNA with the drug ethidium bromide to form a fluorescent complex. The increase of the fluorescence of the ethidium bromide upon formation of the complex permits the observation of the decay of concentration fluctuations via the chemical reaction and consequently the determination of chemical rate constants.

1,289 citations


Journal ArticleDOI
TL;DR: In this paper, the X-ray diffraction data for specimens from the sea alga Valonia ventricosa was used to show a preference for a parallel chain structure with specific chain polarity with respect to the c axis.
Abstract: Native cellulose has been shown to consist of a crystalline array of parallel chains, based on the X-ray diffraction data for specimens from the sea alga Valonia ventricosa. The unit cell is monoclinic with dimensions a = 16.34 A, b = 15.72 A, c = 10.38 A (fiber axis), and β = 97.0°. The space group is P21 and the cell contains disaccharide segments of eight chains. Models containing chains with the same sense (parallel) or alternating sense (antiparallel) were refined against the intensity data using rigidbody least squares procedures. The results show a preference for a parallel chain structure with specific chain polarity with respect to the c axis. The refinement places the CH2OH side chains approximately 20′ from the so-called tg conformation, with a result that an 02′H…06 intramolecular bond is formed. The structure also contains an 03H…05′ intramolecular bond and an 06H…03 intermolecular bond along the a axis. All these bonds lie in the 020 planes, and the structure is an array of hydrogen-bonded sheets. A major consequence of this work is that regular chain folding can be ruled out and cellulose is seen as extended chain polymer single crystals.

598 citations


Journal ArticleDOI
TL;DR: A model for the flexibility of DNA is proposed that is based on discrete variations in the direction of propagation in going from one subunit to the next, which gives a Gaussian distribution function.
Abstract: A model for the flexibility of DNA is proposed that is based on discrete variations in the direction of propagation in going from one subunit to the next. Expansion of the local free energy in terms of the local bending gives a Gaussian distribution function. The assumption of the independence of local bends on neighbors lead to very simple formulae for the persistence length and the characteristic ratio. Emphasis, however, is placed on the application of the formulae for molecules of finite size where the persistence length and C∞ are not defined. The formulae are worked out for two models, which should serve as limits for the real physical situation.

347 citations


Journal ArticleDOI
TL;DR: The elastic properties of elastin are not explicable in terms of a two-phase model consisting of discrete globules of compact elastic molecules fused one to another by cross-linkages, with diluent (water) filling the interstices.
Abstract: The thermoelastic properties of elastin immersed in water or in aqueous solutions of alcohols closely resemble those of typical polymers in the rubber elastic state. The evolution of heat much in excess of the work performed on elastin when it is stretched while immersed in water at ca. 25°C is attributable to the exothermic heat of dilution by water absorbed into the polymer during elongation. The negative sign of the temperature coefficient of swelling is confirmatory of this explanation. A network of random chains within the elastin fibers, like that in a typical rubber, is clearly indicated. The elastic properties of elastin are not explicable in terms of a two-phase model consisting of discrete globules of compact elastin molecules fused one to another by cross-linkages, with diluent (water) filling the interstices.

276 citations


Journal ArticleDOI
TL;DR: This work generalizes the approach of Lacombe and Simha for nearest‐neighbor correlations in specific sequence macromolecules for chains of thousands of units to be calculated in minutes making no approximation in the basic model.
Abstract: The problem of calculating detailed probability profiles giving the probability of each unit in the chain to be in the ordered state (and all other average quantities as well including the fraction of strand association) for specific-sequence macromolecules requiring statistical weights that correlate up to the total number of units in the chain (e.g., DNA, collagen) is formulated in terms of recursion relations for appropriate a priori and conditional probabilities, thus generalizing the approach of Lacombe and Simha for nearest-neighbor correlations in specific sequence macromolecules. The technique allows the probability profiles for chains of thousands of units to be calculated in minutes making no approximation in the basic model.

224 citations


Journal ArticleDOI
TL;DR: In this paper, the dependence of Forster long-range resonance energy transfer efficiency on the orientational freedom of donor D and acceptor A molecules attached to a macromolecular substrate is examined.
Abstract: The dependence of Forster long-range resonance energy transfer efficiency on the orientational freedom of donor D and acceptor A molecules attached to a macromolecular substrate is examined. The usefulness of polarized emission measurements in determining the mutual orientation as well as the degree of orientational freedom of D and A and thereby deriving maximum and minimum values for the D–A separation from the transfer efficieny is demonstrated.

222 citations


Journal ArticleDOI
TL;DR: The endotherm enthalpy changes ΔHD and temperatures TD of thermal denaturation of tropocollagen fibers were measured by DSC calorimetry as functions of water content and the effect of water uptake on theEnthalpy term is explained by water bridge formation within the collagen triple helix.
Abstract: The endotherm enthalpy changes ΔHD and temperatures TD of thermal denaturation of tropocollagen fibers were measured by DSC calorimetry as functions of water content. The denaturation temperatures decrease with increasing water content. The enthalpy change values increase sharply in the range 0–28% of water content, where a maximum of 14.3 cal g−1 is reached. The effect of water uptake on the enthalpy term is explained by water bridge formation within the collagen triple helix. Evidence is given for the existence of approximately three intercatenary water bridges per triplet at the enthalpy maximum, their H-bond energy amounting to approximately 4000 kcal/mol of protein. In the 30–60% range of water content, ΔHD decreases by 2 cal−1 probably due to interactions between secondary water structures and the stabilizing intrahelical water bonds. The influence of two neutral potassium salts, with a structure-stabilizing and a structure-breaking anion (F− and I−), on the hydration dependence of ΔHD and TD was also studied. It was shown that the primary hydration is not influenced by these ions, but that TD and ΔHD are altered in an ion specific way in the presence of interface and bulk water. Hydrophobic interactions do not explain the experimental results. A reaction mechanism of the effects of ions upon the structural stability of collagen is proposed and discussed in terms of interactions of the medium water molecules with the intrahelical water bonds, and in terms of proton-donor/proton-acceptor equilibria between peptide groups, hydrated ions, and intrahelical water molecules.

121 citations


Journal ArticleDOI
TL;DR: Intensities and other spectral parameters of infrared amide I and II bands of α‐helical polypeptides in solutions have been determined for poly(γ‐benzylglutamate), poly (γ‐ethylglutamine), and polymethionine in chloroform, polylysine, poly(glutamic acid), and fibrillar protein tropomyosin from rabbit muscles in heavy water.
Abstract: Intensities and other spectral parameters of infrared amide I and II bands of α-helical polypeptides in solutions have been determined for poly(γ-benzylglutamate), poly(γ-ethylglutamate), and polymethionine in chloroform, polylysine, poly(glutamic acid), and fibrillar protein tropomyosin from rabbit muscles in heavy water. The majority of spectral parameters are characteristic. The half-width of the amide I band was found to vary in the range of 15–40 cm−1 for different polypeptides in the different solutions. The correlation between this parameter of the amide I band and the stability of the α-helix was estimated. A new weak band near 1537 cm−1 of unknown origin was observed for the hydrogen form of polypeptides in the α-helical state.

105 citations


Journal ArticleDOI
TL;DR: The effects of metal ions of the first‐row transition and of alkaline earth metals on the DNA helix conformation have been studied by uv difference spectra, circular dichroism, and sedimentation measurements.
Abstract: The effects of metal ions of the first-row transition and of alkaline earth metals on the DNA helix conformation have been studied by uv difference spectra, circular dichroism, and sedimentation measurements. At low ionic strength (10−3M NaClO4) DNA shows a maximum in the difference absorption spectra in the presence of Zn2+, Mn2+, Co2+, Cd2+, and Ni2+ but not with Mg2+ or Ca2+. The amplitude of this maximum is dependent on GC content as revealed by detailed studies of the DNA-Zn2+ complex of eight different DNA's. Pronounced changes also occur in the CD spectra of DNA transition metal complexes. A transition appears up to a total ratio of approximately 1 Zn2+ per DNA phosphate at 10−3M NaClO4; then no further change was observed up to high concentrations. The characteristic CD changes are strongly dependent on the double-helical structure of DNA and on the GC content of DNA. Differences were also observed in hydrodynamic properties of DNA metal complexes as revealed by the greater increase of the sedimentation coefficient of native DNA in the presence of transition metal ions. Spectrophotometric acid titration experiments and CD measurements at acidic pH clearly indicate the suppression of protonation of GC base-pair regions on the addition of transition metal ions to DNA. Similar effects were not observed with DNA complexes with alkaline earth metal ions such as Mg2+ or Ca2+. The data are interpreted in terms of a preferential interaction of Zn2+ and of other transition metal ions with GC sites by chelation to the N-7 of guanine and to the phosphate residue. The binding of Zn2+ to DNA disappears between 0.5 M and 1 M NaClO4, but complex formation with DNA is observable again in the presence of highly concentrated solutions of NaClO4 (3−7.2 M NaClO4) or at 0.5 to 2 M Mn2+. At relatively high cation concentration Mg2+ is also effective in changing the DNA comformation. These structural alterations probably result from both the shielding of negatively charged phosphate groups and the breakdown of the water structure along the DNA helix. Differential effects in CD are also observed between Mn2+, Zn2+ on one hand and Mg2+ on the other hand under these conditions. The greater sensitivity of the double-helical conformation of DNA to the action of transition metal ions is due to the affinity of the latter to electron donating sites of the bases resulting from the d electronic configuration of the metal ions. An order of the relative phosphate binding ability to base-site binding ability in native DNA is obtained as follows: Mg2+, Ba2+, < Ca2+ < Fe2+, Ni2+, Co2+ < Mn2+, Zn2+ < Cd2+ < Cu2+. The metal-ion induced conformational changes of the DNA are explained by alternation of the winding angle between base pairs as occurs in the transition from B to C conformation. These findings are used for a tentative molecular interpretation of some effects of Zn2+ and Mn2+ in DNA synthesis reported in the literature.

100 citations


Journal ArticleDOI
TL;DR: In this paper, the conformation and crystalline packing of V-anhydrous amylose has been investigated by a combination of linked atom model building and X-ray diffraction analysis.
Abstract: The conformation and crystalline packing of V-anhydrous amylose has been investigated by a combination of linked atom model building and X-ray diffraction analysis. The unit cell, the P212121 space group, the left-handed sixfold helical conformation with all O(6) in gt rotational positions, and the intrahelical O(2)---O(3) and O(2)---O(6) hydrogen bonds are substantially in agreement with previous studies. A new model for packing of the chains in the unit cell and the presence of crystallographic water is proposed. Packing appears to be stabilized by corner-to-center chain O(2)---O(2) hydrogen bonds. The nature of the transition from the amylose–DMSO complex to Va-amylose was considered and it is shown that the transition involves translation of the amylose chains parallel to the a and b unit cell axes with only slight changes in the orientation of the helix. No significant conformational changes result from the transition.

Journal ArticleDOI
TL;DR: The results demonstrate the importance of incorporating far uv transitions in the polarizability approximation in order to achieve an adequate theoretical explanation of the CD spectra of polypeptides.
Abstract: The circular dichroism (CD) spectrum of polyproline II (PPII) has heretofore been moderately well calculated from exciton theory only at the expense of assuming unreasonable chain conformations and accepting a conservative spectrum in the 180–250-nm region (which is not observed). We have incorporated far uv transitions in the polarizability approximation and, together with the π2π* transition, have calculated the resulting correction to the exciton model. This has been accompanied by a modified assignment of the ππ* transition in PPII, and a simultaneous calculation of the absorption and CD spectra of the α-helix, β structure, PPI, and PPII. We obtain good agreement with the observed CD spectrum of PPII in the 180–250-nm region for acceptable chain conformations. In addition, we predict a negative CD into the far uv, in agreement with recent experimental observations. Our calculations also reproduce features of the far uv CD spectrum of the α-helix, and are in agreement with the CD spectra of the β chain and PPI. The calculated CD of the unordered polypeptide chain is not significantly influenced by far uv contributions, indicating that our previous calculation is valid for such a system. These results demonstrate the importance of incorporating far uv transitions in order to achieve an adequate theoretical explanation of the CD spectra of polypeptides.

Journal ArticleDOI
TL;DR: The amide III mode in the Raman spectrum of poly-L-alanine (PLA) is more sensitive than the amide I and II modes to changes in secondary structure of the polypeptide chain this article.
Abstract: Raman spectroscopy has been used in investigating the conformational transitions of poly-L-alanine (PLA) induced by mechanical deformation. We see evidence of the alpha-helical, antiparallel beta-sheet, and a disordered conformation in PLA. The disordered conformation has not been discussed in previous infrared and X-ray diffraction investigations and may have local order similar to the left-handed 31 poly glycine helix. The amide III mode in the Raman spectrum of PLA is more sensitive than the amide I and II modes to changes in secondary structure of the polypeptide chain. Several lines below 1200 cm−1 are conformationally sensitive and may generally be useful in the analysis of Raman spectra of proteins. A line at 909 cm−1 decreases in intensity after deformation of PLA. In general only weak scattering is observed around 900 cm−1 in the Raman spectra of antiparallel beta-sheet polypeptides. The Raman spectra of the amide N–H deuterated PLA and poly-L-leucine (PLL) in the alpha-helical conformation and poly-L-valine (PLV) in the beta-sheet conformation are presented. Splitting is observed in the amide III mode of PLV and the components of this mode are assigned. The Raman spectrum of an alpha-helical random copolymer of L-leucine and L-glutamic acid is shown to be consistent with the spectra of other alphahelical polypeptides.

Journal ArticleDOI
TL;DR: Effective rotational correlation times (τeff), obtained from measurements of spin‐lattice relaxation times (T1) of individual backbone and side‐chain carbons, demonstrated that backbone reorientation is approximately isotropic for the five polypeptides and is characterized by correlation times of 0.3–0.6 nanoseconds as a result of rapid segmental motion.
Abstract: The molecular conformations and dynamics of poly(L-prolyl), poly(hydroxyl-L-prolyl), poly(L-prolyl-glycyl), poly(hydroxyl-L-prolyl), and poly(glycyl-glycyl-L-prolyl-glycyl), in aqueous solution, have been studied using 13C pulse Fourier transform nmr spectroscopy. From a measurement of the intensities of major and minor resonances in the spectra of the copolypeptides, it was determined that 15–20% of the glycyl-prolyl and glycyl-hydroxyprolyl peptide bonds are cis. Effective rotational correlation times (τeff), obtained from measurements of spin-lattice relaxation times (T1) of individual backbone and side-chain carbons, demonstrated that backbone reorientation is approximately isotropic for the five polypeptides and is characterized by correlation times of ca. 0.3–0.6 nanoseconds as a result of rapid segmental motion. In a given polypeptide glycyl and pyrrolidine residues were found to have the same backbone correlation times, but backbone carbon τeff values did decrease as the glycyl content of the peptides increased. A semi-quantitative analysis of Cβ, Cγ, and Cδ correlation times suggests that rapid ring motion in both prolyl and hydroxyprolyl involves primarily Cγ and Cβ, with the prolyl ring being more mobile than the hydroxyprolyl ring.

Journal ArticleDOI
TL;DR: In this paper, the ultraviolet circular dichroism spectra of oligo d(pG)5, poly dN AcG, poly I, poly DG, and double-stranded poly dG:dC were measured.
Abstract: We have measured the ultraviolet circular dichroism spectra of oligo d(pG)5, poly dN AcG, poly dI, poly dC, two samples of poly dG, and four samples containing double-stranded poly dG:dC. We find that oligo d(pG)5 and poly dG exist in self-complexed forms as well as in single-stranded forms. Unlike the self-complexed form of poly dG, the single-stranded form of poly dG can hydrogen-bond with single-stranded poly dC. We present spectral data for double-stranded poly dG:dC, which can be used to help characterize poly dG:dC preparations and which provide a basis for resolving discrepancies among other reported poly dG:dC spectra.

Journal ArticleDOI
TL;DR: A two‐dimensional representation, accurate to about 25 Å, of the modified quarter‐stagger arrangement of molecules in the collagen fibril is prepared and it is possible to demonstrate the relationship of loci on individual collagen molecules within the fibrils.
Abstract: The location of transverse bands within the major repeating period of positively stained collagen fibrils was determined from electron micrographs by an optical averaging procedure. From these data and the published location of bands in SLS crystallites, we have prepared a two-dimensional representation, accurate to about 25 A, of the modified quarter-stagger arrangement of molecules in the collagen fibril. With this information it is possible to demonstrate the relationship of loci on individual collagen molecules within the fibril. For example, the site where the collagen molecule is cleaved by tadpole collagenase, the site where a disaccharide unit is covalently bound to the α1-CB5 peptide, and the site of carboxyl-terminal intermolecular cross-linking all occur in the fibril near the amino-terminal edge of the “hole zone;” and the site of amino-terminal cross-linking occurs near the carboxyl-terminal edge of the “hole zone.”

Journal ArticleDOI
TL;DR: Assignments of the absorption bands of the dye are proposed in connection with wave‐mechanical calculations recently reported on the binding of ethidium bromide to DNA.
Abstract: The binding of ethidium bromide to DNA has been studied by various optical methods. From fluorescence polarization studies, and film, electric linear dichroism, and circular dichroism spectra, we propose assignments of the absorption bands of the dye, which are discussed in connection with wave-mechanical calculations recently reported. The optical activity induced in the dye absorption bands upon binding to DNA was attributed to various origins depending on the electronic transition considered. The visible absorption band displayed a circular dichroism due to the asymmetry of the binding site and independent of the amount of binding. The transition identified at 378 nm from the circular dichroism and electric dichroism observations was thought to be due to a magnetic-dipole transition. It remained constant with increasing amounts of dye bound. The main ultraviolet band showed circular dichroism characteristics corresponding to exciton interactions between dye molecules bound to neighboring sites. The electric dichroism observed for the strongly bound dye molecules indicated that the phenanthridinium ring of ethidium bromide was probably not perfectly parallel to the DNA base planes. When the amount of dye bound to DNA exceeded the maximum amount compatible with the exclusion of adjacent binding sites, the electric dichroism decreased owing to the appearance of externally bound dye molecules with no contribution to the dichroism. Sonicated DNA was used to study the lengthening of the DNA molecule upon complexation. Although the viscosity of the complexes increased with the amount of binding, the rotational diffusion coefficient measured by the electric birefringence relaxation was not detectably affected. The absence of variation in the electric birefringence with the binding indicated that the DNA base stacking remained unaltered.

Journal ArticleDOI
TL;DR: Double‐stranded DNA can be readily adsorbed on mica or directly on carbon coated grids from the surface of solutions containing ethidium bromide, actinomine, or propidium diiodide.
Abstract: Double-stranded DNA can be readily adsorbed on mica or directly on carbon coated grids from the surface of solutions containing ethidium bromide, actinomine, or propidium diiodide. The DNA molecules are unfolded, well separated, and show a length distribution similar to molecules prepared by protein monolayer techniques. Since the intercalating dyes tested do not lead to an increased apparent diameter of the nucleic acid the method is useful for the study of nucleic acid–protein complexes. As a model, the binding of E. coli RNA polymerase to phage T7 and T3 DNA was examined under different conditions. The enzyme can easily be identified and its position along the DNA molecule can be mapped.

Journal ArticleDOI
TL;DR: The crystal and molecular structure of amylose with dimethyl sulfoxide has been studied by a combination of stereochemical analysis, potential energy, and X-ray diffraction methods as mentioned in this paper.
Abstract: The crystal and molecular structure of the complex of amylose with dimethyl sulfoxide has been studied by a combination of stereochemical analysis, potential energy, and X-ray diffraction methods. The complex crystallizes in a pseudotetragonal unit cell with a = b = 19.17 A and c (fiber axis) = 24.39 A, with two antiparallel chains per unit cell and space group P212121. The amylose chain is a left-handed 61(1.355) helix with three turns per crystallographic repeat. The O(6) rotational position is approximately gt. Dimethyl sulfoxide is located inside the helix with one DMSO molecule for every three glucose residues. An additional four DMSO molecules and eight water molecules each are located in the large interstices between chains, and it is the interaction of these molecules with the helix that results in the pseudotetragonal chain packing. The interstitial DMSO is the source of the previously reported additional layer lines, which are not consistent with the 8.13-A amylose repeat distance. The final R factor for the layers with amylose contribution to the structure factors was 0.29, while the overall R factor was 0.35. The stereochemical packing analysis provided suitable phasing models for the subsequent X-ray refinement.

Journal ArticleDOI
TL;DR: The Raman spectra of the pH denaturation of tropomyosin are presented and are consistent with the helix‐to‐coil studies of the polypeptides.
Abstract: Raman spectra of the pH denaturation of tropomyosin are presented. In the native state tropomyosin has an alpha-helical content of nearly 90%, but this value drops rapidly as the pH is raised above 9.5. The Raman spectrum of the native state is characterized by a strong amide I line appearing at 1655 cm−1, very weak scattering in the amide III region around 1250 cm−1, and a medium-intensity line at 940 cm−1. When the protein is pH-denatured, a strong amide III line appears at 1254 cm−1 and the 940 cm−1 line becomes weak. The intensities of the latter two lines are a sensitive measure of the alpha-helical and disordered chain content. These results are consistent with the helix-to-coil studies of the polypeptides. The Raman spectra of α-casein and prothrombin, proteins thought to have little or no ordered secondary structure, are investigated. The amide III regions of both spectra display strong lines at 1254 cm−1 and only weak scattering is observed at 940 cm−1, features characteristic of the denatured tropomyosin spectrum. The amide I mode of α-casein appears at 1668 cm−1, in agreement with the previously reported spectra of disordered polypeptides, poly-L-glutamic acid and poly-L-lysine at pH 7.0 and mechanically deformed poly-L-alanine.

Journal ArticleDOI
TL;DR: All of the general features of polyribonucleotide spectra have been reproduced and, in many cases, good quantitative agreement between observed and calculated frequencies is obtained.
Abstract: A normal coordinate analysis has been performed for the polyribonucleotides poly(rU) and poly(rA). The polymers are assumed to be 11-fold infinite helical structures in the A conformation. The hydrogen atoms have been rigidly attached to the appropriate atoms in order to reduce the dimension of the problem. The potential energy is defined in terms of a valence force field initially and a model for the inclusion of nonbonded interactions has been presented. A method for factoring the secular equation for infinite helical polymers in Cartesian coordinates is presented. All of the general features of polyribonucleotide spectra have been reproduced and, in many cases, good quantitative agreement between observed and calculated frequencies is obtained. More specific assignments are offered for some of the observed lines.

Journal ArticleDOI
TL;DR: Circular dichroism spectroscopy has been used to study the interactions of hyaluronic acid, heparitin sulfate, and keratan sulfate with cationic polypeptides, and the results indicate that the presence of these mucopolysaccharides has an effect in the conformation of poly(L‐lysine) and poly( L‐arginine), such that the former adopts the “random” form and the latter takes up
Abstract: Circular dichroism spectroscopy has been used to study the interactions of hyaluronic acid, heparitin sulfate, and keratan sulfate with cationic polypeptides. The results indicate that the presence of these mucopolysaccharides has an effect in the conformation of poly(L-lysine) and poly(L-arginine), such that the former adopts the “random” form and the latter takes up the α-helical conformation, rather than the “charged coil” form expected at neutral pH. The relative strengths of the interactions can be judged from the melting temperatures above which they are disrupted. Both the stoichiometry and the strength of the interactions depend on the position, number, and type of anionic groups attached to the polysaccharide backbone. Such considerations place the six common mucopolysaccharides in order of increasing strength of interaction: hyaluronic acid < chondroitin 4-sulfate < heparitin sulfate < chondroitin 6-sulfate < keratan sulfate ⩽ dermatan sulfate. These differences should be paralleled by differences in the interaction of the mucopolysaccharides with collagen and fibrous proteins.

Journal ArticleDOI
TL;DR: The optical properties of the diketopiperazine chromophore of the cyclic dipeptides have been investigated as a function of molecular conformation and it has been observed that the degree of conservatism of the π → π* couplet of L‐propyl–L‐proline dketopipersazine depends markedly upon the nature of the solvent that is used.
Abstract: The optical properties of the diketopiperazine chromophore of the cyclic dipeptides have been investigated as a function of molecular conformation. The rotatory strengths of L-alanyl–L-alanine diketopiperazine and L-prolyl–L-proline diketopiperazine have been calculated as a function of the angle of fold of the diketopiperazine ring. The results of these theoretical calculations have been compared with experimental circular dichroism and optical rotatory dispersion data. It is shown that the observed optical properties of these molecules can be explained only if their diketopiperazine rings are folded in opposite directions. The direction of fold is established for each molecule. In solution, the diketopiperazine ring of L-alanyl-L-alanine diketopiperazine is folded in the direction opposite to that found by X-ray diffraction analysis of crystals. It has been observed that the degree of conservatism of the π π* couplet of L-propyl–L-proline diketopiperazine depends markedly upon the nature of the solvent that is used. In addition, a shoulder has been discovered in the CD spectrum of L-alanyl–L-alanine diketopiperazine, which may not be directly attributable to the n π* and π π* transitions of the peptide chromophores.

Journal ArticleDOI
TL;DR: Use of the regression of specific fluidity (1 − 1/ηrel) with concentration was found superior to other methods and its only deficiency was a requirement that solution density be corrected for at each concentration studied rather than applying a single correction for density after using kinematic viscosity data.
Abstract: Intrinsic viscosity [η] is a characteristic of proteins and other molecules related directly to their ability to disturb flow and indirectly to their size and shape. It is usually determined by extrapolating reduced viscosity to zero concentration. Four other methods for deriving [η] have been utilized by previous investigators. Studies of the intrinsic viscosity of bovine serum albumin had been carried out two years apart as a test of viscometry technique; the data obtained were used to compare the five methods. Four of the five produced [η] values ranging from 3.92 to 4.21 ml/g. Agreement was good between the two studies. The five methods were compared to each other using linearity of regression, statistical error of determination, effect of varying solvent time, and result obtained in different concentration ranges. By these four criteria, use of the regression of specific fluidity (1 − 1/ηrel) with concentration was found superior to other methods. Its only deficiency was a requirement that solution density be corrected for at each concentration studied rather than applying a single correction for density after using kinematic viscosity data. All methods for deriving intrinsic viscosity are based on one of three equations; flow is expressed either in terms of reduced viscosity (Huggins), inherent viscosity (Kraemer), or specific fluidity. Of these three equations, specific fluidity is the most closely related both to theoretical analyses and to experimental studies of rigid spheres. There is abundant evidence in past reports that in contrast to rigid spheres, flexible polymers do not produce a linear rise in specific fluidity as their concentration increases, strongly suggesting that their molecular conformation is changing with concentration. A linear relation between fluidity and concentration has been observed for almost all proteins and protein mixtures studied. When this linear relation is present it indicates both that molecular conformation during flow is independent of concentration in the range studied and that the specific fluidity method for deriving intrinsic viscosity is the most appropriate.

Journal ArticleDOI
TL;DR: Increasing the macroscopic viscosity with dissolved neutral or anionic polymers increases, rather than decreases, renaturation rates due to the excluded volume of the dissolved polymers.
Abstract: The effect of solvent viscosity on DNA renaturation rates has been investigated as a function of temperature for a number of solvent systems. The results are all consistent with a microscopic viscosity limitation of the rate determining step. Rates of renaturation in perchlorate and quaternary ammonium salt solutions are also discussed. Increasing the macroscopic viscosity with dissolved neutral or anionic polymers increases, rather than decreases, renaturation rates due to the excluded volume of the dissolved polymers.

Journal ArticleDOI
TL;DR: Two configurations of the glycan strands and the peptide side chains are described, which by their constancy of form reflect the fundamental constancies of the covalent structures, which will accommodate any chemical modification that has been observed in bacteria without change in the configuration of the peptides backbone.
Abstract: Bacterial cell wall peptidoglycans are built from unbranched β-(1 → 4)-linked glycan chains composed of alternately repeating units of N-acetylglucosamine and N-acetylmuramic acid residues, with peptide side chains attached to the muramic acid residues. The glycan chains are interconnected by peptide bonds formed between the peptide side chains. Through the use of three-dimensional molecular models, two configurations of the glycan strands and the peptide side chains are described, which by their constancy of form reflect the fundamental constancies of the covalent structures. Each of these two models will accommodate any chemical modification that has been observed in bacteria without change in the configuration of the peptide backbone. Some alterations in the chemical structure, which have been sought in bacteria, but not found, would not be tolerated by the models. In these models, glycan strands are parallel, with their lengths and widths predominantly in the plane of the cell wall. The cross-bridging portions of the peptide side chains are at right angles to the glycan strand, in a separate, parallel plane. A compact model is presented in which the peptide side chain is closely appressed to the glycan strand and is stabilized by three hydrogen bonds per disaccharide–peptide subunit. In a second model, the peptide side chain is raised away from the glycan strand in an entirely extended configuration. The compact and extended forms are interconvertible. The thickness of a sheet of peptidoglycan would be from 10.6 to 11.1 A for the compact model, and 19.1 A for the extended model.

Journal ArticleDOI
TL;DR: The alternating copolypeptide poly(Lys-Phe) was obtained by synthesis and polymerization of the respective tetrapeptide Conformation and conformational transitions in aqueous solutions and in water-organic solvent mixtures were determined and compared with those of the random copolymer poly (Lys51,Phe49) by circular dichroism measurements.
Abstract: The alternating copolypeptide poly(Lys-Phe) was obtained by synthesis and polymerization of the respective tetrapeptide Conformation and conformational transitions in aqueous solutions and in water–organic solvent mixtures were determined and compared with those of the random copolymer poly(Lys51,Phe49) by circular dichroism measurements The alternating copolymer reveals a very strong tendency to adopt the β structure, which is accomplished by raising the pH or by adding sodium perchlorate or methanol Partial α-helical conformation, however, is obtained in the presence of 10–20% of the strong helix-forming solvent hexafluoroisopropanol The random copolymer reveals a much weaker preference for the β structure It assumes the α helix in mixtures of water with hexafluoroisopropanol and also in those with methanol As was already reported by Peggion et al and confirmed by the present work, the addition of sodium perchlorate also leads to the helix conformation [(1972) Biopolymers11, 633], whereas β structure is obtained by raising the pH [(1970) Macromolecules3, 194] These studies reveal the significance of affinities determined by near neighbor interactions in polypeptides

Journal ArticleDOI
TL;DR: The present results demonstrate the sensitivity of the Raman technique to sequence‐related structural differences in oligonucleotides and provide additional spectra–structure correlations for future conformational studies of RNA by laser Raman spectroscopy.
Abstract: Laser Raman spectra of the trinucleoside diphoshate ApApA and dinucleoside phosphates ApU, UpA, GpC, CpG, and GpU are reported and discussed. Assignments of conformationally sensitive frequencies are-facilitated by comparison with spectra reported here of poly(rA), poly(rC), and poly(rU) in deuterium oxide solutions. The significant spectral differences between ApU and UpA, and between GpC and CpG, reveal that the sequence isomers have nonidentical conformations in aqueous solution. In UpA at low temperature the bases are stacked and the backbone conformation is similar to that found in ordered polynucleotide structures and RNA. In ApU no base stacking can be detected and the backbone conformation differs from that found in UpA, both in the orientation of phosphodiester linkages and in the internal conformation of ribose. At the conditions employed neither ApU nor UpA exhibits base pairing in aqueous solutions. In both GpC and CpG the bases are stacked and the phosphodiester conformations are similar to those encountered for UpA and RNA. However, major differences between spectra of GpC and CpG indicate that the geometries of stacking and ribosyl conformations are different. In GpC the Raman data favor the formation of hydrogen bonded dimers containing GC pairs. Protonation of C in GpC is sufficient to eliminate the ordered conformation detected by Raman spectroscopy. Despite the ordered backbone conformation evident in GpU, this dinucleoside apparently contains neither stacked nor hydrogen bonded bases at the conditions employed here. The Raman data also confirm the stacking interactions in ApApA, poly(rA), and poly(rC) but suggest that the backbone conformation in poly(rC) differs qualitatively from that found in most ordered polynucleotide structures and is thermally more stable. The present results demonstrate the sensitivity of the Raman technique to sequence-related structural differences in oligonucleotides and provide additional spectra–structure correlations for future conformational studies of RNA by laser Raman spectroscopy.

Journal ArticleDOI
TL;DR: It is concluded that 5′AMP at pH 8 forms multistacked aggregates at high concentration levels and that a preferred orientation is such that the bases are aligned face to back with considerable base overlap; and that the ribose moieties of adjacent molecules are near one another with the phosphate groups well separated.
Abstract: Proton magnetic resonance spectra of 5′AMP were taken in the concentration range of 0.001–2.2M. The concentration profiles of all the nonexchangeable protons were determined. The data for 5′AMP was compared to those of adenine, adenosine, and poly(A). Theoretically computed isoshielding lines of the adenine moiety were used to qualitatively predict a preferred stacking geometry of 5′AMP in aqueous solution. It is concluded that 5′AMP at pH 8 forms multistacked aggregates at high concentration levels and that a preferred orientation is such that the bases are aligned face to back with considerable, though less than 100%, base overlap; and that the ribose moieties of adjacent molecules are near one another with the phosphate groups well separated. Mn(II) ion binding studies show that the stacks are not restricted to one unique orientation type. Specific evidence is given showing that base-stacking orientations in the solid state may in some cases be considerably different from that in aqueous solution, due in part to numerous hydrogen bonding differences, and this is shown to be the case for base-stacked adenosine. In the case of 5′AMP the stacking orientations between the solid and liquid states are also different, except in this comparison the solid-state structure carries a positive charge.

Journal ArticleDOI
TL;DR: The 300‐MHz proton nmr spectra (between 11 and 14 ppm) of a series of double‐stranded deoxy oligonucleotides of known sequence have been recorded in H2O solution and perturbations on these up‐field ring‐current contributions that arise from twisting and tilting a base pair adjacent to the ring NH under study have been evaluated.
Abstract: The 300-MHz proton nmr spectra (between 11 and 14 ppm) of a series of double-stranded deoxy oligonucleotides of known sequence have been recorded in H2O solution. These resonances have been assigned to the GN1H and TN3H protons of specific base pairs from an evaluation of the temperature dependence of the ring NH linewidths and from the selective ring NH chemical shift changes on actinomycin-D binding. The deoxy oligonucleotides exist predominantly in the DNA-B conformation as evaluated from antibiotic binding studies. Ring-current calculations have been utilized to evaluate the up-field shifts of the GN1H and TN3H protons in Watson-Crick base pairs due to the ring currents from the pyrimidine and purine rings of nearest neighbor base pairs in regular DNA-B- and RNA-A-type helices. The perturbations on these up-field ring-current contributions that arise from twisting and tilting a base pair adjacent to the ring NH under study have been evaluated and found to change the calculated chemical shift by ±0.6 ppm for twist and tilt distortions of <30°C in a single adjacent base pair. A knowledge of the experimentally assigned ring NH chemical shifts of specific base pairs in known sequences of double-stranded deoxy oligonucleotides coupled with the ring-current tables for the DNA-B helical structure permit the assignment of 13.6 ± 0.1 ppm and 14.6 ± 0.2 ppm for the GN1H proton of an isolated GC base pair and the TN3H proton of an isolated AT base pair, respectively.