Network


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

Hotspot


Dive into the research topics where David R. Edwards is active.

Publication


Featured researches published by David R. Edwards.


Biochemistry | 2012

Catalysis by a De Novo Zinc-Mediated Protein Interface: Implications for Natural Enzyme Evolution and Rational Enzyme Engineering

Bryan S. Der; David R. Edwards; Brian Kuhlman

Here we show that a recent computationally designed zinc-mediated protein interface is serendipitously capable of catalyzing carboxyester and phosphoester hydrolysis. Although the original motivation was to design a de novo zinc-mediated protein-protein interaction (called MID1-zinc), we observed in the homodimer crystal structure a small cleft and open zinc coordination site. We investigated if the cleft and zinc site at the designed interface were sufficient for formation of a primitive active site that can perform hydrolysis. MID1-zinc hydrolyzes 4-nitrophenyl acetate with a rate acceleration of 10(5) and a k(cat)/K(M) of 630 M(-1) s(-1) and 4-nitrophenyl phosphate with a rate acceleration of 10(4) and a k(cat)/K(M) of 14 M(-1) s(-1). These rate accelerations by an unoptimized active site highlight the catalytic power of zinc and suggest that the clefts formed by protein-protein interactions are well-suited for creating enzyme active sites. This discovery has implications for protein evolution and engineering: from an evolutionary perspective, three-coordinated zinc at a homodimer interface cleft represents a simple evolutionary path to nascent enzymatic activity; from a protein engineering perspective, future efforts in de novo design of enzyme active sites may benefit from exploring clefts at protein interfaces for active site placement.


Journal of the American Chemical Society | 2012

Catalytic Proficiency: The Extreme Case of S–O Cleaving Sulfatases

David R. Edwards; Danielle C. Lohman; Richard Wolfenden

As benchmarks for judging the catalytic power of sulfate monoesterases, we sought to determine the rates of spontaneous hydrolysis of unactivated alkyl sulfate monoesters by S-O bond cleavage. Neopentyl sulfate proved to be unsuitable for this purpose, since it was found to undergo hydrolysis by a C-O bond cleaving mechanism with rearrangement of its carbon skeleton. Instead, we examined the temperature dependence of the spontaneous hydrolyses of aryl sulfate monoesters, which proceed by S-O cleavage. Extrapolation of a Bronsted plot [log(k(25)(N)) = (-1.81 ± 0.09) pK(a)(LG) + (3.6 ± 0.7)] based on the rate constants at 25 °C for hydrolysis of a series of sulfate monoesters to a pK(a)(LG) value of 16.1, typical of an aliphatic alcohol, yields k(25)(N) = 3 × 10(-26) s(-1). Comparison of that value with established k(cat) values of bacterial sulfatases indicates that these enzymes produce rate enhancements (k(cat)/k(uncat)) of up to 2 × 10(26)-fold for the hydrolysis of sulfate monoesters. These rate enhancements surpass by several orders of magnitude the ~10(21)-fold rate enhancements that are generated by phosphohydrolases, the most powerful biological catalysts previously known. The hydrolytic rates of phosphate and sulfate monoesters are compared directly, and the misleading impression that the two classes of ester are of similar reactivity is dispelled.


Journal of the American Chemical Society | 2009

Leaving Group Assistance in the La3+-Catalyzed Cleavage of Dimethyl (o-Methoxycarbonyl)aryl Phosphate Triesters in Methanol

David R. Edwards; C. Tony Liu; Graham E. Garrett; Alexei A. Neverov; R. Stan Brown

The catalytic methanolysis of a series of dimethyl aryl phosphate triesters where the aryl groups contain an o-methoxycarbonyl (o-CO2Me) substituent (4a-i) was studied at 25 degrees C in methanol containing La3+ at various concentrations and (s)(s)pH. Determination of the second-order rate constant for La3+(2)-catalyzed cleavage of substrate 4a (dimethyl (o-methoxycarbonyl)phenyl phosphate) as a function of (s)(s)pH was assessed in terms of a speciation diagram that showed that the process was catalyzed by La3+(2)(-OCH3)x dimers, where x = 1-5, that exhibit only a 5-fold difference in activity between all the species. The second-order catalytic rate constants (k2(La)) for the catalyzed methanolysis of 4a-i at (s)(s)pH 8.7 fit a Brønsted relationship of log k2(La) = (-0.82 +/- 0.11)(s)(s)pKa(lg) + (11.61 +/- 1.48), where the gradient is shallower than that determined for a series of dimethyl aryl phosphates that do not contain the o-CO2Me substituent, log k2(La) = (-1.25 +/- 0.06)(s)(s)pKa(lg) + (16.23 +/- 0.75). Two main observations are that (1) the o-CO2Me group preferentially accelerates the cleavage of the phosphate triesters with poor leaving groups relative to those with good leaving groups and (2) it provides an increase in cleavage rate relative to those of comparable substrates that do not have that functional group, e.g., k2(La)(dimethyl o-(methoxycarbonyl)phenyl phosphate)/k2(La)(dimethyl phenyl phosphate) = 60. Activation parameters for the La3+(2)-catalyzed methanolysis of 4a and dimethyl 4-nitrophenyl phosphate show respective DeltaH(double dagger) (DeltaS(double dagger)) values of 3.3 kcal/mol (-47 cal/mol x K) and 0.7 kcal/mol (-46.5 cal/mol x K). The data are analyzed in terms of a concerted reaction where the catalytic complex (La3+(2)(-OCH3)(x-1)) binds to the three components of a rather loose transition state composed of a nucleophile CH3O-, a nucleofuge -OAr, and a central (RO)2P(2+)-O(-) in a way that provides leaving group assistance to the departing aryloxy group.


Inorganic Chemistry | 2011

A Study of the Kinetics of La3+-Promoted Methanolysis of S-Aryl Methylphosphonothioates: Possible Methodology for Decontamination of EA 2192, the Toxic Byproduct of VX Hydrolysis

Basab B. Dhar; David R. Edwards; R. Stan Brown

The kinetics of the La3+-catalyzed methanolysis of a series of S-aryl methylphosphonothioates (4a-e, phenyl substituents = 3,5-dichloro, 4-chloro, 4-fluoro, 4-H, 4-methoxy) were studied at 25 °C with s(s)pH control. The reaction involves saturation binding of the anionic substrates to dimeric La3+/methoxide catalysts formulated as La2(3+)(-OCH3)x, where x = 2-5 depending on the solution s(s)pH. Cleavage of the La3+-bound methylphosphonothioates is fast, ranging from 5 × 10(-3) s(-1) to 5.5 × 10-(5) s(-1) for substrates 4a-e at a s(s)pH of 8.4 and 1.6 × 10(-1) s(-1) to 4 × 10(-3) s(-1) at a s(s)pH of 11.7. The rate accelerations for the methanolysis of substrates 4a-e, relative to their background methoxide-promoted reactions, average 7 × 10(10) and 1.5 × 10(9), respectively, at s(s)pHs of 8.4 and 11.7. The catalytic system is predicted to cleave EA 2192 (S-2(N,N-di-iso-propylaminoethyl)methylphosphonothioate), a toxic byproduct of the hydrolysis of VX, with a t1/2 between 4 and 8 min at a s(s)pH of 8.4, and 27 min at a s(s)pH of 11.7.


Journal of the American Chemical Society | 2010

Mechanistic and computational study of a palladacycle-catalyzed decomposition of a series of neutral phosphorothioate triesters in methanol.

C. Tony Liu; Christopher I. Maxwell; David R. Edwards; Alexei A. Neverov; Nicholas J. Mosey; R. Stan Brown

The methanolytic cleavage of a series of O,O-dimethyl O-aryl phosphorothioates (1a−g) catalyzed by a C,N-palladacycle, (2-[N,N-dimethylamino(methyl)phenyl]-C1,N)(pyridine) palladium(II) triflate (3), at 25 °C and sspH 11.7 in methanol is reported, along with data for the methanolytic cleavage of 1a−g. The methoxide reaction gives a linear log k2−OMe vs sspKa (phenol leaving group) Brønsted plot having a gradient of βlg = −0.47 ± 0.03, suggesting about 34% cleavage of the P−OAr bond in the transition state. On the other hand, the 3-catalyzed cleavage of 1 gives a Brønsted plot with a downward break at sspKa (phenol) 13, signifying a change in the rate-limiting step in the catalyzed reaction, with the two wings having βlg values of 0.0 ± 0.03 and −1.93 ± 0.06. The rate-limiting step for good substrates with low leaving group sspKa values is proposed to be substrate/pyridine exchange on the palladacycle, while for substrates with poor leaving groups, the rate-limiting step is a chemical one with extensive cleavage of the P−OAr bond. DFT calculations support this process and also identify two intermediates, namely, one where substrate/pyridine interchange has occurred to give the palladacycle coordinated to substrate through the S═P linkage and to methoxide (6) and another where intramolecular methoxide attack has occurred on the P═S unit to give a five-coordinate phosphorane (7) doubly coordinated to Pd via the S− and through a bridging methoxide linked to P and Pd. Attempts to identify the existence of the phosphorane by 31P NMR in a d4-methanol solution containing 10 mM each of 3, trimethyl phosphorothioate (a very slow cleaving substrate), and methoxide proved unsuccessful, instead showing that the phosphorothioate was slowly converted to trimethyl phosphate, with the palladacycle decomposing to Pd0 and free pyridine. These results provide the first reported example where a palladacycle-promoted solvolysis reaction exhibits a break in the Brønsted plot signifying at least one intermediate, while the DFT calculations provide further insight into a more complex mechanism involving two intermediates.


Journal of the American Chemical Society | 2009

Dissociative Solvolytic Cleavage of Methyl (ortho-Carboxymethyl)Aryl Phosphate Diesters Mediated by Yb3+ in Methanol Gives a 1012-Fold Rate Acceleration Attributable to Leaving Group Assistance

David R. Edwards; Alexei A. Neverov; R. Stan Brown

The Yb3+-catalyzed cleavage of a series of eight methyl aryl phosphates (2a−h) where the aryl groups all contain an ortho-methoxycarbonyl group was studied in acidic methanol from 1.34 ≤ sspH ≤ 3.34 at 25 °C. All substrates show saturation binding of the metal ion that is analyzed to provide a conditional binding constant (Kb) for a 1:1 substrate/Yb3+ complex and catalytic rate constant (kcat) that varies between about 2 × 10−3 and 50 × 10−3 s−1 over the range of substrates. Detailed analysis indicates that at very low concentration of Yb3+, 3 equiv of substrate are bound, and with increasing [Yb3+], the binding changes to a 1:1 complex which decomposes by a pathway independent of sspH over the range investigated. Control studies show that substrates without the o-methoxycarbonyl group still bind to the Yb3+ with approximately the same strength as do the o-methoxycarbonyl containing substrates but have no observable reaction when bound. A Jaffe plot of the kcat vs substituent σ-values indicates that, duri...


Journal of the American Chemical Society | 2013

Catalysis by desolvation: the catalytic prowess of SAM-dependent halide-alkylating enzymes.

Danielle C. Lohman; David R. Edwards; Richard Wolfenden

In the biological fixation of halide ions, several enzymes have been found to catalyze alkyl transfer from S-adenosylmethionine to halide ions. It proves possible to measure the rates of reaction of the trimethylsulfonium ion with I(-), Br(-), Cl(-), F(-), HO(-), and H2O in water at elevated temperatures. Comparison of the resulting second-order rate constants, extrapolated to 25 °C, with the values of k(cat)/K(m) reported for fluorinase and chlorinase indicates that these enzymes enhance the rates of alkyl halide formation by factors of 2 × 10(15)- and 1 × 10(17)-fold, respectively. These rate enhancements, achieved without the assistance of cofactors, metal ions, or general acid-base catalysis, are the largest that have been reported for an enzyme that acts on two substrates.


Journal of Organic Chemistry | 2012

Proton-in-Flight Mechanism for the Spontaneous Hydrolysis of N-Methyl O-Phenyl Sulfamate: Implications for the Design of Steroid Sulfatase Inhibitors

David R. Edwards; Richard Wolfenden

The hydrolysis of N-methyl O-phenyl sulfamate (1) has been studied as a model for steroid sulfatase inhibitors such as Coumate, 667 Coumate, and EMATE. At neutral pH, simulating physiological conditions, hydrolysis of 1 involves an intramolecular proton transfer from nitrogen to the bridging oxygen atom of the leaving group. Remarkably, this proton transfer is estimated to accelerate the decomposition of 1 by a factor of 10(11). Examination of existing kinetic data reveals that the sulfatase PaAstA catalyzes the hydrolysis of sulfamate esters with catalytic rate accelerations of ~10(4), whereas the catalytic rate acceleration generated by the enzyme for its cognate substrate is on the order of ~10(15). Rate constants for hydrolysis of a wide range of sulfuryl esters, ArOSO(2)X(-), are shown to be correlated by a two-parameter equation based on pK(a)(ArOH) and pK(a)(ArOSO2XH).


Journal of Organic Chemistry | 2012

Hydrolysis of N-alkyl sulfamates and the catalytic efficiency of an S-N cleaving sulfamidase.

Danielle C. Lohman; Richard Wolfenden; David R. Edwards

The final step in the degradation of heparin sulfate involves the enzymatic hydrolysis of its 2-sulfamido groups. To evaluate the power of the corresponding sulfamidases as catalysts, we examined the reaction of N-neopentyl sulfamate at elevated temperatures and found it to undergo specific acid catalyzed hydrolysis even at alkaline pH. A rate constant of 10(-16) s(-1) was calculated using the Eyring equation for water attack on the N-protonated species at pH 7, 25 °C. As a model for the pH neutral reaction, a rate constant for hydroxide attack on (CH(3))(3)CCH(2)N(+)H(2)SO(3)(-) at pH 7, 25 °C was calculated to be 10(-19) s(-1). The corresponding rate enhancement (k(cat)/k(non)) produced by the N-sulfamidase of F. heparinum is approximately 10(16)-fold, which is somewhat larger than those generated by most hydrolytic enzymes but considerably smaller than those generated by S-O cleaving sulfatases.


Journal of Organic Chemistry | 2006

Theoretical and Experimental Studies on the Baeyer−Villiger Oxidation of Ketones and the Effect of α-Halo Substituents

Friedrich Grein; Austin C. Chen; David R. Edwards; Cathleen M. Crudden

Collaboration


Dive into the David R. Edwards's collaboration.

Top Co-Authors

Avatar
Top Co-Authors

Avatar
Top Co-Authors

Avatar

Richard Wolfenden

University of North Carolina at Chapel Hill

View shared research outputs
Top Co-Authors

Avatar
Top Co-Authors

Avatar

Danielle C. Lohman

University of North Carolina at Chapel Hill

View shared research outputs
Top Co-Authors

Avatar
Top Co-Authors

Avatar
Top Co-Authors

Avatar

Brian Kuhlman

University of North Carolina at Chapel Hill

View shared research outputs
Top Co-Authors

Avatar

Bryan S. Der

University of North Carolina at Chapel Hill

View shared research outputs
Top Co-Authors

Avatar

Wing Tsang

National Institute of Standards and Technology

View shared research outputs
Researchain Logo
Decentralizing Knowledge