Your browser doesn't support javascript.
loading
Show: 20 | 50 | 100
Results 1 - 20 de 61
Filter
1.
J Am Chem Soc ; 123(27): 6649-53, 2001 Jul 11.
Article in English | MEDLINE | ID: mdl-11439052

ABSTRACT

We have developed an instrumental setup that uses transient absorption to monitor protein folding/unfolding processes following a laser-induced, ultrafast release of protons from o-nitrobenzaldehyde. The resulting increase in [H(+)], which can be more than 100 microM, is complete within a few nanoseconds. The increase in [H(+)] lowers the pH of the solution from neutrality to approximately 4 at the highest laser pulse energy used. Protein structural rearrangements can be followed by transient absorption, with kinetic monitoring over a broad time range (approximately 10 ns to 500 ms). Using this pH-jump/transient absorption technique, we have examined the dissociation kinetics of non-native axial heme ligands (either histidine His26 or His33) in GuHCl-unfolded Fe(III) cytochrome c (cyt c). Deligation of the non-native ligands following the acidic pH-jump occurs as a biexponential process with different pre-exponential factors. The pre-exponential factors markedly depend on the extent of the pH-jump, as expected from differences in the pK(a) values of His26 and His33. The two lifetimes were found to depend on temperature but were not functions of either the magnitude of the pH-jump or the pre-pulse pH of the solution. The activation energies of the deligation processes support the suggestion that GuHCl-unfolded cyt c structures with non-native histidine axial ligands represent kinetic traps in unfolding.


Subject(s)
Bacterial Proteins , Cytochrome c Group/chemistry , Guanidine/chemistry , Heme/chemistry , Histidine/chemistry , Lasers , Animals , Horses , Hydrogen-Ion Concentration , Kinetics , Ligands , Protein Folding
2.
Biophys J ; 79(5): 2714-21, 2000 Nov.
Article in English | MEDLINE | ID: mdl-11053143

ABSTRACT

Photoactivatable caged protons have been used to trigger proton transfer reactions in aqueous solutions of acetate, glutamate, and poly-L-glutamic acid, and the volumetric and enthalpic changes have been detected and characterized by means of time-resolved photoacoustics. Neutralization of carboxylates in aqueous solutions invariably results in an expansion of the solution due to the disappearance of two charges and is accompanied by little enthalpic change. The reactions occur with thermally activated, apparent bimolecular rates on the order of 10(10) M(-1)s(-1). In the case of aqueous solutions of poly-L-glutamic acid at pH around the pK(a) of the coil-to-helix transition, diffusional binding of a proton by carboxylates is followed by a sequential reaction with rate 1.06 (+/- 0.05) x 10(7)s(-1). This step is not thermally activated in the temperature range we have investigated and is likely related to local formation of hydrogen bonds near the protonation site. This structural event may constitute a rate-limiting step in helix propagation.


Subject(s)
Polyglutamic Acid/chemistry , Acoustics , Biophysical Phenomena , Biophysics , Carboxylic Acids/chemistry , In Vitro Techniques , Kinetics , Photolysis , Protein Folding , Protein Structure, Secondary , Protons , Solutions , Water
3.
Biophys J ; 78(1): 405-15, 2000 Jan.
Article in English | MEDLINE | ID: mdl-10620304

ABSTRACT

Ultrafast, laser-induced pH jump with time-resolved photoacoustic detection has been used to investigate the early protonation steps leading to the formation of the compact acid intermediate (I) of apomyoglobin (ApoMb). When ApoMb is in its native state (N) at pH 7.0, rapid acidification induced by a laser pulse leads to two parallel protonation processes. One reaction can be attributed to the binding of protons to the imidazole rings of His24 and His119. Reaction with imidazole leads to an unusually large contraction of -82 +/- 3 ml/mol, an enthalpy change of 8 +/- 1 kcal/mol, and an apparent bimolecular rate constant of (0.77 +/- 0.03) x 10(10) M(-1) s(-1). Our experiments evidence a rate-limiting step for this process at high ApoMb concentrations, characterized by a value of (0. 60 +/- 0.07) x 10(6) s(-1). The second protonation reaction at pH 7. 0 can be attributed to neutralization of carboxylate groups and is accompanied by an apparent expansion of 3.4 +/- 0.2 ml/mol, occurring with an apparent bimolecular rate constant of (1.25 +/- 0.02) x 10(11) M(-1) s(-1), and a reaction enthalpy of about 2 kcal/mol. The activation energy for the processes associated with the protonation of His24 and His119 is 16.2 +/- 0.9 kcal/mol, whereas that for the neutralization of carboxylates is 9.2 +/- 0.9 kcal/mol. At pH 4.5 ApoMb is in a partially unfolded state (I) and rapid acidification experiments evidence only the process assigned to carboxylate protonation. The unusually large contraction and the high energetic barrier observed at pH 7.0 for the protonation of the His residues suggests that the formation of the compact acid intermediate involves a rate-limiting step after protonation.


Subject(s)
Apoproteins/chemistry , Hydrogen-Ion Concentration , Myoglobin/chemistry , Protein Folding , Animals , Horses , Kinetics , Models, Molecular , Protein Conformation , Protein Structure, Secondary , Spectrometry, Fluorescence
4.
Biophys Chem ; 73(1-2): 13-22, 1998 Jul 13.
Article in English | MEDLINE | ID: mdl-17029715

ABSTRACT

A fast perturbation in proton concentration can be induced in aqueous solution using a pulsed ultraviolet laser and suitable photolabile compounds which, upon photoexcitation, irreversibly release protons. The volume change and the rate constant for the reaction of the photodetached protons with proton-accepting groups in solution can be monitored using time resolved photoacoustics. A typical proton concentration jump of 1 microM can be obtained with a 200-microJ laser pulse at 308 nm. Reaction dynamics from 20 ns to 5 micros can be easily followed. The methodology we establish represents a direct, time-resolved measurement of the reaction volume in proton transfer processes and an extension to the nanosecond-microsecond range of traditional relaxation techniques, such as stopped-flow. We report example applications to reactions involving simple molecules and polypeptides.

6.
Biochemistry ; 32(40): 10527-31, 1993 Oct 12.
Article in English | MEDLINE | ID: mdl-8399198

ABSTRACT

Nucleosome core particles undergo a conformational change at ionic strengths below 0.2 mM; the fluorescence anisotropy decay of bound ethidium indicates that under these conditions the particle adopts a highly extended structure. We have measured the distribution of UV-induced DNA damage (primarily cyclobutane-pyrimidine dimers) through a process termed photofootprinting. As the core particle is exposed to ionic strengths below 0.2 mM, the photofootprint pattern changes from that observed for native cores, with a characteristic 10.3 base repeat pattern presumably derived largely from the bending of DNA around the histone octamer, to a more evenly distributed pattern resembling that of free DNA. These results provide clear evidence that the DNA in the core particle at these very low ionic strengths, although still tightly bound to histones, is no longer bent to a significant degree.


Subject(s)
DNA Damage , DNA/radiation effects , Nucleosomes/ultrastructure , Ultraviolet Rays , Animals , Chickens , DNA/ultrastructure , Erythrocytes/ultrastructure , Models, Structural , Nucleic Acid Conformation/radiation effects , Nucleosomes/radiation effects , Osmolar Concentration
7.
Biophys Chem ; 42(1): 29-48, 1992 Jan.
Article in English | MEDLINE | ID: mdl-1581513

ABSTRACT

Pulsed-laser photoacoustics is a technique which measures photoinduced enthalpic and volumetric changes on the nano- and microsecond timescales. Analysis of photoacoustic data generally requires deconvolution for a sum of exponentials, a procedure which has been developed extensively in the field of time-resolved fluorescence decay. Initial efforts to adapt an iterative nonlinear least squares computer program, utilizing the Marquardt algorithm, from the fluorescence field to photoacoustics indicated that significant modifications were needed. The major problem arises from the wide range of transient decay times which must be addressed by the photoacoustic technique. We describe an alternative approach to numerical convolution with exponential decays, developed to overcome the problems. Instead of using an approximation method (Simpson's rule) for evaluating the convolution integral, we construct a continuous instrumental response function by quadratic fitting of the discrete data and evaluate the convolution integral directly, without approximations. The success and limitations of this quadratic-fit convolution program are then demonstrated using simulated data. Finally, the program is applied to the analysis of experimental data to compare the resolution capabilities of two commercially available transducers. The advantages of a broadband, heavily damped transducer are shown for a standard organic photochemical system, the quenching of the triplet state of benzophenone by 2,5-dimethyl-2,4-hexadiene.


Subject(s)
Benzophenones/chemistry , Lasers , Photochemistry , Software , Spectrum Analysis , Algorithms , Fluorescence , Least-Squares Analysis , Mathematics
8.
Methods Enzymol ; 210: 237-79, 1992.
Article in English | MEDLINE | ID: mdl-1584039

ABSTRACT

If one has a convoluted fluorescence decay and wishes to analyze it for a sum of exponential, then one can begin by asking either of two questions: (1) What sum of exponentials best fits the data? (2) What physical decay parameters gave rise to the data? At first these two questions may sound equivalent; in fact, they represent different philosophical approaches to data analysis. In resolving the first question, one adjusts the decay parameters until a calculated curve agrees within arbitrarily chosen limits to the original data. This is what we did in the fourth section of Table II. The fit obtained was decent, but the resulting parameters were wrong. A more difficult approach is to design a method of data analysis which is intrinsically insensitive to the presence of anticipated errors, aiming directly at recovering the decay parameters without regard to the fit. This is what we have done with the method of moments with MD. If particular errors do not have much effect on the recovered parameters, then such a method of data analysis is said to be robust with respect to those errors. Robust methods are widely used in engineering but have not seen much introduction yet to biophysics. Least-squares, the basis of the commonly used data fitting methods for pulse fluorometry, is nonrobust with respect to underlying noise distributions. Isenberg has shown that least-squares is nonrobust with respect to the nonrandom light scatter, time origin shift, and lamp width errors as well. As shown in Isenberg's paper, as well as here, the method of moments with MD is quite robust with respect to these nonrandom errors. Perhaps question (1) could be modified to include all of the errors that might be present in the data; but then, how would one decide which errors to include and whether an error is present? What fitting criterion would tell one this? Why choose a method which depends so strongly on this information when robust alternatives exist? As a rule, fitting should not be used as a criterion for correct decay parameters, unless all of the significant nonrandom errors have been included in the fit. If one fits the data but has not incorporated an important error, then the best fit will necessarily give the wrong answer. The method of moments provides clear criteria for accepting or rejecting an analysis.(ABSTRACT TRUNCATED AT 400 WORDS)


Subject(s)
Data Interpretation, Statistical , Spectrometry, Fluorescence/methods , Kinetics , Mathematical Computing , Software
9.
Biochemistry ; 30(23): 5644-52, 1991 Jun 11.
Article in English | MEDLINE | ID: mdl-2043608

ABSTRACT

We have previously reported that the binding of ethidium bromide to the nucleosome core particle results in a stepwise dissociation of the structure which involves the initial release of one copy each of H2A and H2B (McMurray & van Holde, 1986). In this report, we have examined the absorbance and fluorescence properties of intercalated and outside bound forms of ethidium bromide. From these properties, we have measured the extent of external, electrostatic binding of the dye versus internal, intercalation binding to the core particle, free from contribution by linker DNA. We have established that dissociation is induced by the intercalation mode of binding to DNA within the core particle DNA, and not by binding to the histones or by nonintercalative binding to DNA. The covalent binding of [3H]-8-azidoethidium to the core particle clearly shows that less than 1.0 adduct is formed per histone octamer over a wide range of input ratios. Simultaneously, analyses of steady-state fluorescence enhancement and fluorescence lifetime data from bound ethidium complexes demonstrate extensive intercalation binding. Combined analyses from steady-state fluorescence intensity with equilibrium dialysis or fluorescence lifetime data revealed that dissociation began when approximately 14 ethidium molecules are bound by intercalation to each core particle and less than 1.0 nonintercalated ion pair was formed per core particle.


Subject(s)
Ethidium/metabolism , Intercalating Agents/metabolism , Nucleosomes/metabolism , Polyvinyls/metabolism , Animals , Binding Sites , Chickens , Erythrocytes , Kinetics , Spectrometry, Fluorescence
10.
Biochemistry ; 30(21): 5293-303, 1991 May 28.
Article in English | MEDLINE | ID: mdl-2036396

ABSTRACT

The fluorescence decay of ethidium intercalated into the DNA of nucleosome core particles increases in average lifetime from about 22 ns in H2O to about 39 ns in D2O. This increase, combined with the acquisition of large amounts of data (on the order of 10(8) counts per decay), allows measurement of anisotropy decays out to more than 350 ns. The overall slow rotational motions of the core particle may thereby be more clearly distinguished from the faster torsional motions of the DNA. In 10 mM NaCl at 20 degrees C, we recover a long correlation time of 198 ns in D2O (159 ns when corrected to a viscosity of 1.002 cP), in agreement with the value of 164 ns obtained in H2O. These values are consistent with hydrodynamic calculations based on the expected size and shape of the hydrated particle. To support our conclusion that this long correlation time derives from Brownian rotational diffusion, we show that the value is directly proportional to the viscosity and inversely proportional to the temperature. No significant changes in the rotational correlation time are observed between 1 and 500 mM ionic strength. Below 1 mM, the particle undergoes the "low-salt transition" as measured by steady-state tyrosine fluorescence anisotropy. However, we observe little change in shape until the ionic strength is decreased below approximately 0.2 mM, where the correlation time increases nearly 2-fold, indicating that the particle has opened up into an extended form. We have previously shown that the transition becomes nonreversible below 0.2 mM salt.


Subject(s)
Nucleosomes/ultrastructure , Animals , Chickens , Diffusion , Ethidium/chemistry , Fluorescence Polarization , Hydrogen-Ion Concentration , In Vitro Techniques , Nucleosomes/chemistry , Osmolar Concentration , Temperature , Viscosity
11.
Biochemistry ; 30(21): 5304-13, 1991 May 28.
Article in English | MEDLINE | ID: mdl-2036397

ABSTRACT

The effects of pH on the torsional flexibility of DNA bound to nucleosome core particles were investigated by using time-resolved fluorescence anisotropy decays of intercalated ethidium. The decays were collected by using time-resolved single-photon counting and were fit to a model developed by J. M. Schurr [(1984) Chem. Phys. 84, 71-96] with a nonlinear least-squares-fitting algorithm developed for this purpose. As the torsional flexibility of DNA is affected by the presence of an intercalating dye, the decays were studied at different ethidium bromide to core particle binding ratios. Because we see large increases in DNA flexibility and in the rotational diffusion coefficient at binding ratios of 0.6 ethidium/core particle and above, we conclude that, under these conditions, the DNA begins to detach from the protein. At lower binding ratios, we observe only small changes in the anisotropy decay. The torsional parameters obtained are a function of N, the number of base pairs of DNA between points of attachment to the histone core. Only if N is greater than 30 base pairs is the torsional rigidity of DNA on a nucleosome core particle higher than that for DNA free in solution. Also, for reasonable values of N (less than 30), the friction felt by the DNA on a core particle is much higher than that felt by free DNA. This indicates that the region of the DNA to which the ethidium binds is highly constrained in its motions. pH changes nearly neutrality at moderate ionic strengths (100 mM) have a substantial effect on the fluorescence anisotropy decays, particularly at early times. These analyses indicated that the observed change on increasing pH can be attributed either to a loosening of the contacts between the DNA and the histone core (increasing N) or to a substantial relaxing of the torsional rigidity of the DNA.


Subject(s)
DNA/chemistry , Nucleosomes/chemistry , Animals , Chickens , Diffusion , Ethidium/chemistry , Fluorescence Polarization , Hydrogen-Ion Concentration , In Vitro Techniques , Motion , Nucleosomes/ultrastructure
12.
Biophys Chem ; 34(3): 269-82, 1989 Nov.
Article in English | MEDLINE | ID: mdl-2611350

ABSTRACT

In fluorescence decay work, distributions of exponential decay lifetimes are anticipated where complex systems are examined. We describe here methods of gaining information on such distributions using the method of moments analysis approach. The information obtained may be as simple as the average and deviation of the lifetime distribution, quantities which we show may be estimated directly from the results of a multiexponential analysis. An approximation to the actual distribution shape may also be obtained using a procedure we call the variable filter analysis (VFA) method without making any assumptions about the shape of the distribution. Tests of VFA using both simulated and experimental data are described. Limitations of this method and of distribution analysis methods in general are discussed. Results of analyses on experimental decays for ethidium intercalated in core particles and in free DNA are reported.


Subject(s)
Spectrometry, Fluorescence/methods , Kinetics , Mathematics , Statistics as Topic , Time Factors
14.
Biophys J ; 53(4): 477-87, 1988 Apr.
Article in English | MEDLINE | ID: mdl-3132988

ABSTRACT

Effects of histone hyperacetylation on transitions of HeLa cell nucleosome core particles were studied. The transitions examined were induced by low salt concentrations, pH, temperature, and nondissociating high salt. Effects of salt dissociation were also examined. The low-salt transition was found to shift to higher ionic strength by approximately three fold for hyperacetylated particles, a change which may be due simply to the increased overall negative charge on the particles caused by acetylation of lysine residues. Some differences were also seen in the way in which core particles refold after exposure to very low salt (which induces a nonreversible change in the particles). Otherwise no significant effects of hyperacetylation were observed.


Subject(s)
Histones/metabolism , Nucleosomes/metabolism , Acetylation , HeLa Cells/metabolism , Humans , Hydrogen-Ion Concentration , Lysine , Nucleosomes/ultrastructure , Osmolar Concentration
15.
Biochemistry ; 27(1): 419-28, 1988 Jan 12.
Article in English | MEDLINE | ID: mdl-3349043

ABSTRACT

We report dynamic fluorescence anisotropy measurements on the purified dityrosine derivative of calmodulin which was generated during UV irradiation of Ca2+-containing solutions of bovine brain calmodulin [Malencik, D. A., & Anderson, S. R. (1987) Biochemistry 26, 695]. Measurements were made by using a high repetition rate picosecond laser source combined with a microchannel plate photomultiplier. This permits the collection of very low noise anisotropy curves with essentially no convolution artifact. Measured anisotropies at high calcium concentrations are monoexponential, and at 20 degrees C, we recover a correlation time of 9.9 ns. When the temperature is varied from 4.8 to 31.8 degrees C, the recovered correlation time is proportional to the viscosity and inversely proportional to the absolute temperature, behavior expected for the rotational diffusion of a macromolecule whose conformation is independent of the temperature. The correlation time is compared to the theory describing the rotational diffusion of a dumbell. At high calcium concentrations, the cross-linked calmodulin is elongated and has a length equal or nearly equal to that predicted by X-ray crystallographic results. In the absence of calcium, the molecule becomes highly compact and exhibits significant segmental motion. Intermediate calcium ion concentrations result in an intermediate degree of elongation and segmental motion. A small increase in the measured rotational correlation time of calmodulin upon the binding of melittin and mastoparan indicates that these peptides cause no major changes in the elongation of the molecule. When the cross-linked calmodulin is bound to troponin I, the complex rotates as a unit with a single rotational correlation time of 22 ns.


Subject(s)
Calcium/pharmacology , Calmodulin/metabolism , Cross-Linking Reagents/pharmacology , Morpholines/pharmacology , Animals , Brain/metabolism , Cattle , Fluorescence Polarization , Kinetics , Mathematics , Models, Molecular , Protein Binding , Protein Conformation , X-Ray Diffraction
16.
Oral Surg Oral Med Oral Pathol ; 65(1): 23-8, 1988 Jan.
Article in English | MEDLINE | ID: mdl-3277108

ABSTRACT

Following dental extractions, prophylactic antibiotic protection of patients at high risk of cardiovalvular infection is usually discontinued before suture removal. To determine whether bacteremia is created upon removal of intraoral sutures, twenty healthy patients who required extractions of at least five erupted teeth and placement of several sutures were selected without regard to sex, age, or race. Blood samples were drawn preoperatively, immediately after the extractions, before suture removal, and immediately following removal of the intraoral silk sutures. The samples were cultured in prereduced and aerobic media suitable for quantitative colony counts. Fourteen of 16 patients yielded positive blood cultures following tooth extractions. One of 20 patients yielded a positive blood culture following suture removal. Even though the incidence of bacteremia following intraoral suture removal is relatively low (5%), this study suggests that intraoral suture removal is not a benign procedure for those persons who are considered high-risk cardiac patients.


Subject(s)
Mouth Mucosa/surgery , Sepsis/etiology , Suture Techniques , Bacteria/isolation & purification , Humans , Risk Factors , Sepsis/microbiology , Streptococcus/isolation & purification , Tooth Extraction
17.
Nucleic Acids Res ; 15(16): 6655-64, 1987 Aug 25.
Article in English | MEDLINE | ID: mdl-3628003

ABSTRACT

The low-salt transition of chromatin core particles is reversible if the monovalent cation concentration is kept above 0.2 mM. Exposure of the particles to salt concentrations below this value results in a nonreversible secondary transition. The nonreversible changes are relatively slow with a half-time of about 15 minutes. Once exposed to such low ionic strength, the particles then begin to refold with increasing salt in at least two steps over a much higher ionic strength range than is required for the usual low-salt transition. The refolding is very fast, with a half-time less than a minute. Small differences between particles which had or had not been exposed to very low salt persist even when the particles are returned to near physiological ionic strengths.


Subject(s)
Chromatin/ultrastructure , Animals , Cations, Monovalent/pharmacology , Chickens , Chromatin/drug effects , Hydrogen-Ion Concentration , Magnesium/pharmacology , Nucleic Acid Conformation/drug effects , Nucleosomes/drug effects , Nucleosomes/ultrastructure , Osmolar Concentration , Protein Conformation/drug effects
18.
Anal Biochem ; 163(2): 500-5, 1987 Jun.
Article in English | MEDLINE | ID: mdl-3661999

ABSTRACT

The choice of laser dyes for exciting tyrosine fluorescence using synchronously pumped cavity-dumped dye laser systems is discussed. Rhodamine 560 was found to be optimal for a system based on an argon-ion pumping laser, whereas rhodamine 575 was preferred using a frequency-doubled Nd:YAG laser. Modifications of our fluorescence decay instrument to permit rejection of multiphoton events using a microchannel plate photomultiplier are described. An example of a four-component resolution of tyrosine decays illustrates the dramatic resolution capabilities attainable.


Subject(s)
Lasers , Rhodamines , Tyrosine , Xanthenes , Spectrometry, Fluorescence
20.
Biophys J ; 47(6): 765-72, 1985 Jun.
Article in English | MEDLINE | ID: mdl-4016197

ABSTRACT

In wavelength-resolved steady state spectra we observe three different kinds of emission from histone H1, a class A protein with only a single tyrosine residue. Unfolded H1 emissions that peak at approximately 300 and 340 nm can both be excited maximally at approximately 280 nm. Another, peaking much further to the red at approximately 400 nm, can be excited maximally at approximately 320 nm. The 300-nm fluorescence can be resolved by lifetime measurements into three components with decay times of approximately 1, 2, and 4 ns. On sodium-chloride-induced refolding of H1, simplification of the emission properties occurs. The 340 and 400-nm components disappear while the two shorter lifetime components of the 300-nm band diminish in amplitude and are replaced by the 4-ns decay. We believe that the 340-nm emission is tyrosinate fluorescence resulting from excited-state proton transfer. The origin of the 400-nm emission remains uncertain. We assign the 1 and 2-ns components of the 300-nm emission to two states of tyrosine in denatured H1 and the 4-ns decay to fluorescence of the single tyrosine residue in the globular region of refolded H1. Our results support the contention that salt induced folding of H1 is a cooperative two state process, and permit us to better understand the previously reported increases in fluorescence intensity and anisotropy on salt-induced folding.


Subject(s)
Histones , Tyrosine/analysis , Animals , Cattle , Chromatin , Kinetics , Spectrometry, Fluorescence/methods , Thymus Gland
SELECTION OF CITATIONS
SEARCH DETAIL
...