Network


Latest external collaboration on country level. Dive into details by clicking on the dots.

Hotspot


Dive into the research topics where Charles A. Lewis is active.

Publication


Featured researches published by Charles A. Lewis.


Proceedings of the National Academy of Sciences of the United States of America | 2010

Impact of temperature on the time required for the establishment of primordial biochemistry, and for the evolution of enzymes

Randy B. Stockbridge; Charles A. Lewis; Yang Yuan; Richard Wolfenden

All reactions are accelerated by an increase in temperature, but the magnitude of that effect on very slow reactions does not seem to have been fully appreciated. The hydrolysis of polysaccharides, for example, is accelerated 190,000-fold when the temperature is raised from 25 to 100 °C, while the rate of hydrolysis of phosphate monoester dianions increases 10,300,000-fold. Moreover, the slowest reactions tend to be the most heat-sensitive. These tendencies collapse, by as many as five orders of magnitude, the time that would have been required for early chemical evolution in a warm environment. We propose, further, that if the catalytic effect of a “proto-enzyme”—like that of modern enzymes—were mainly enthalpic, then the resulting rate enhancement would have increased automatically as the environment became cooler. Several powerful nonenzymatic catalysts of very slow biological reactions, notably pyridoxal phosphate and the ceric ion, are shown to meet that criterion. Taken together, these findings greatly reduce the time that would have been required for early chemical evolution, countering the view that not enough time has passed for life to have evolved to its present level of complexity.


Proceedings of the National Academy of Sciences of the United States of America | 2008

Uroporphyrinogen decarboxylation as a benchmark for the catalytic proficiency of enzymes

Charles A. Lewis; Richard Wolfenden

The magnitude of an enzymes affinity for the altered substrate in the transition state exceeds its affinity for the substrate in the ground state by a factor matching the rate enhancement that the enzyme produces. Particularly remarkable are those enzymes that act as simple protein catalysts, without the assistance of metals or other cofactors. To determine the extent to which one such enzyme, human uroporphyrinogen decarboxylase, enhances the rate of substrate decarboxylation, we examined the rate of spontaneous decarboxylation of pyrrolyl-3-acetate. Extrapolation of first-order rate constants measured at elevated temperatures indicates that this reaction proceeds with a half-life of 2.3 × 109 years at 25 °C in the absence of enzyme. This enzyme shows no significant homology with orotidine 5′-monophosphate decarboxylase (ODCase), another cofactorless enzyme that catalyzes a very slow reaction. It is proposed that, in both cases, a protonated basic residue (Arg-37 in the case of human UroD; Lys-93 in the case of yeast ODCase) furnishes a counterion that helps the scissile carboxylate group of the substrate leave water and enter a relatively nonpolar environment, stabilizes the incipient carbanion generated by the departure of CO2, and supplies the proton that takes its place.


Journal of the American Chemical Society | 2011

Kinetic Challenges Facing Oxalate, Malonate, Acetoacetate, and Oxaloacetate Decarboxylases

Richard Wolfenden; Charles A. Lewis; Yang Yuan

To compare the powers of the corresponding enzymes as catalysts, the rates of uncatalyzed decarboxylation of several aliphatic acids (oxalate, malonate, acetoacetate, and oxaloacetate) were determined at elevated temperatures and extrapolated to 25 °C. In the extreme case of oxalate, the rate of the uncatalyzed reaction at pH 4.2 was 1.1 × 10(-12) s(-1), implying a 2.5 × 10(13)-fold rate enhancement by oxalate decarboxylase. Whereas the enzymatic decarboxylation of oxalate requires O(2) and Mn(II), the uncatalyzed reaction is unaffected by the presence of these cofactors and appears to proceed by heterolytic elimination of CO(2).


Proceedings of the National Academy of Sciences of the United States of America | 2015

Temperature dependence of amino acid hydrophobicities

Richard Wolfenden; Charles A. Lewis; Yang Yuan; Charles W. Carter

Significance Systematic relationships have long been recognized between the hydrophobicities of amino acids and (i) their tendencies to be located at the exposed surfaces of globular and membrane proteins and (ii) the composition of their triplets in the genetic code. Here, we show that the same coding relationships are compatible with the high temperatures at which life is widely believed to have originated. An accompanying paper reports that these two properties appear to be encoded separately by bases in the acceptor stem and the anticodon of tRNA. The hydrophobicities of the 20 common amino acids are reflected in their tendencies to appear in interior positions in globular proteins and in deeply buried positions of membrane proteins. To determine whether these relationships might also have been valid in the warm surroundings where life may have originated, we examined the effect of temperature on the hydrophobicities of the amino acids as measured by the equilibrium constants for transfer of their side-chains from neutral solution to cyclohexane (Kw>c). The hydrophobicities of most amino acids were found to increase with increasing temperature. Because that effect is more pronounced for the more polar amino acids, the numerical range of Kw>c values decreases with increasing temperature. There are also modest changes in the ordering of the more polar amino acids. However, those changes are such that they would have tended to minimize the otherwise disruptive effects of a changing thermal environment on the evolution of protein structure. Earlier, the genetic code was found to be organized in such a way that—with a single exception (threonine)—the side-chain dichotomy polar/nonpolar matches the nucleic acid base dichotomy purine/pyrimidine at the second position of each coding triplet at 25 °C. That dichotomy is preserved at 100 °C. The accessible surface areas of amino acid side-chains in folded proteins are moderately correlated with hydrophobicity, but when free energies of vapor-to-cyclohexane transfer (corresponding to size) are taken into consideration, a closer relationship becomes apparent.


Biochemistry | 2009

Orotic Acid Decarboxylation in Water and Nonpolar Solvents: a Potential Role For Desolvation in the Action Of OMP Decarboxylase

Charles A. Lewis; Richard Wolfenden

OMP decarboxylase (ODCase) generates a very large rate enhancement without the assistance of metals or other cofactors. The uncatalyzed decarboxylation of 1-methylorotate in water is shown to involve the monoanion, although uncharged 1-methylorotic acid is decarboxylated at a similar rate. To measure the extent to which the rate of the nonenzymatic decarboxylation of orotate derivatives might be enhanced by their removal from solvent water, the 1-phosphoribosyl moiety of OMP was replaced with 1-substituents that would allow it to enter less polar solvents. When the tetrabutylammonium salt of 1-cyclohexylorotate was transferred from water to a series of dipolar aprotic solvents, its rate of decarboxylation increased markedly, varying with the relative ability of each solvent to release the substrate in the ground state from stabilization by solvent water acting as a proton donor. These findings are consistent with the view that separation of the substrate from solvent water may contribute, at least to a limited extent, to the rate enhancement produced by ODCase. This enzymes active site, like that of another cofactorless enzyme recently shown to produce a rate enhancement similar in magnitude (uroporphyrinogen decarboxylase), is equipped with an ammonium group positioned in such a way as to balance the electrostatic charge of the carboxylate group of the substrate and later supply a proton to the incipient carbanion in a relatively waterless environment.


Journal of the American Chemical Society | 2014

The Nonenzymatic Decomposition of Guanidines and Amidines

Charles A. Lewis; Richard Wolfenden

To establish the rates and mechanisms of decomposition of guanidine and amidine derivatives in aqueous solution and the rate enhancements produced by the corresponding enzymes, we examined their rates of reaction at elevated temperatures and used the Arrhenius equation to extrapolate the results to room temperature. The similar reactivities of methylguanidine and 1,1,3,3-tetramethylguanidine and their negative entropies of activation imply that their decomposition proceeds by hydrolysis rather than elimination. The influence of changing pH on the rate of decomposition is consistent with attack by hydroxide ion on the methylguanidinium ion (k2 = 5 × 10(-6) M(-1) s(-1) at 25 °C) or with the kinetically equivalent attack by water on uncharged methylguanidine. At 25 °C and pH 7, N-methylguanidine is several orders of magnitude more stable than acetamidine, urea, or acetamide. Under the same conditions, the enzymes arginase and agmatinase accelerate substrate hydrolysis 4 × 10(14)-fold and 6 × 10(12)-fold, respectively, by mechanisms that appear to involve metal-mediated water attack. Arginine deiminase accelerates substrate hydrolysis 6 × 10(12)-fold by a mechanism that (in contrast to the mechanisms employed by arginase and agmatinase) is believed to involve attack by an active-site cysteine residue.


Proceedings of the National Academy of Sciences of the United States of America | 2016

Cytosine deamination and the precipitous decline of spontaneous mutation during Earth's history

Charles A. Lewis; Jesse Crayle; Shuntai Zhou; Ronald Swanstrom; Richard Wolfenden

Significance Cytosine deamination appears to be largely responsible for spontaneous mutations in the modern world. Because of its sensitivity to temperature (Q10 = 4), that reaction would have furnished a mechanism for rapid evolution on a warm earth. As the temperature fell from 100° to 25 °C, the rate of cytosine-based mutation would have fallen by a factor of more than 4,000, with a corresponding increase in the stability of genetic information. Other potentially mutagenic events are known to be even more sensitive to temperature, and would presumably have led to an even steeper decline in the rate of spontaneous mutation as the earth cooled. The hydrolytic deamination of cytosine and 5-methylcytosine residues in DNA appears to contribute significantly to the appearance of spontaneous mutations in microorganisms and in human disease. In the present work, we examined the mechanism of cytosine deamination and the response of the uncatalyzed reaction to changing temperature. The positively charged 1,3-dimethylcytosinium ion was hydrolyzed at a rate similar to the rate of acid-catalyzed hydrolysis of 1-methylcytosine, for which it furnishes a satisfactory kinetic model and a probable mechanism. In agreement with earlier reports, uncatalyzed deamination was found to proceed at very similar rates for cytosine, 1-methylcytosine, cytidine, and cytidine 5′-phosphate, and also for cytosine residues in single-stranded DNA generated from a phagemid, in which we sequenced an insert representing the gene of the HIV-1 protease. Arrhenius plots for the uncatalyzed deamination of cytosine were linear over the temperature range from 90 °C to 200 °C and indicated a heat of activation (ΔH‡) of 23.4 ± 0.5 kcal/mol at pH 7. Recent evidence indicates that the surface of the earth has been cool enough to support life for more than 4 billion years and that life has been present for almost as long. If the temperature at Earths surface is assumed to have followed Newtons law of cooling, declining exponentially from 100 °C to 25 °C during that period, then half of the cytosine-deaminating events per unit biomass would have taken place during the first 0.2 billion years, and <99.4% would have occurred during the first 2 billion years.


Biochemistry | 2011

Amide bonds to the nitrogen atoms of cysteine and serine as "weak points" in the backbones of proteins.

Charles A. Lewis; Richard Wolfenden

During the initial event in protein self-splicing, a peptide bond to the nitrogen atom of an internal cysteine or serine residue is usually cleaved by the side chain -SH or -OH group to yield a thioester or oxyester intermediate that undergoes further reactions. Self-splicing reactions also accompany the maturation of hedgehog signaling proteins, plant-type asparaginases, and pyruvoyl enzymes. It would be of interest to know whether peptide bonds that involve the nitrogen atoms of cysteine or serine are more susceptible to cleavage than peptide bonds to amino acids that lack reactive side chains. Extrapolations of the results of model reactions conducted at elevated temperatures indicate that the -SH group of N-acetylcysteine enhances the rate of its hydrolysis by a factor of 70, while the OH group of N-acetylserine enhances the rate of its hydrolysis 12-fold, compared with the rate of hydrolysis of N-acetylalanine in neutral solution at 25 °C. Several lines of evidence suggest that the rate-enhancing effects of these -SH and -OH side chains arise from their ability to act as intramolecular general acid-base catalysts for hydrolysis, rather than as nucleophilic catalysts. The protein environment within self-splicing proteins appears to redirect the actions of these side chains to nucleophilic attack, generating rate enhancements that approach the rate enhancements generated by conventional enzymes.


Biochemistry | 2017

Three Pyrimidine Decarboxylations in the Absence of a Catalyst

Charles A. Lewis; Lin Shen; Weitao Yang; Richard Wolfenden

The epigenetic modification of DNA by 5-methylation of cytosine residues can be reversed by the action of the TET family of dioxygenases that oxidize the methyl group to produce 5-carboxycytosine (5caC), which can be converted to cytosine in a final decarboxylation step. Likewise, 5-carboxyuracil (5caU) is decarboxylated to uracil in the last step in pyrimidine salvage. In view of the extreme difficulty of decarboxylating derivatives of orotic acid (6caU), it seemed desirable to establish the rates of decarboxylation of 5caC and 5caU in the absence of a catalyst. Arrhenius analysis of experiments performed at elevated temperatures indicates that 5caU decomposes with a rate constant of 1.1 × 10-9 s-1 (ΔH⧧ = 25 kcal/mol) in a neutral solution at 25 °C. The decomposition of 5caC is somewhat slower (k25 = 5.0 × 10-11 s-1; ΔH⧧ = 27 kcal/mol) and leads to the initial accumulation of cytosine as an intermediate, followed by the relatively rapid deamination of cytosine (k25 = 1.9 × 10-10 s-1; ΔH⧧ = 23.4 kcal/mol). Both 5caC and 5caU are decarboxylated many orders of magnitude more rapidly than 6caU is (k25 = 1.3 × 10-17 s-1). Ab initio simulations indicate that in all three cases, the favored route of spontaneous decarboxylation in water involves direct elimination of CO2 with the assistance of an explicit water molecule.


Biochemistry | 1977

Antiproteolytic aldehydes and ketones: substituent and secondary deuterium isotope effects on equilibrium addition of water and other nucleophiles.

Charles A. Lewis; Richard Wolfenden

Collaboration


Dive into the Charles A. Lewis's collaboration.

Top Co-Authors

Avatar

Richard Wolfenden

University of North Carolina at Chapel Hill

View shared research outputs
Top Co-Authors

Avatar

Yang Yuan

University of North Carolina at Chapel Hill

View shared research outputs
Top Co-Authors

Avatar

R. Bruce Dunlap

University of South Carolina

View shared research outputs
Top Co-Authors

Avatar
Top Co-Authors

Avatar

Charles W. Carter

University of North Carolina at Chapel Hill

View shared research outputs
Top Co-Authors

Avatar

Jesse Crayle

University of North Carolina at Chapel Hill

View shared research outputs
Top Co-Authors

Avatar
Top Co-Authors

Avatar

Paul D. Ellis

University of South Carolina

View shared research outputs
Top Co-Authors

Avatar

Randy B. Stockbridge

University of North Carolina at Chapel Hill

View shared research outputs
Researchain Logo
Decentralizing Knowledge