Charge Dispersion and Its Effects on Reactivity of Thiamin-Derived

Chemistry, University of Toronto, 80 St. George Street,. Toronto, Ontario M5S 3H6 Canada. E-mail: [email protected]. Notes .... [30] DePuy, C. H., ...
0 downloads 0 Views 514KB Size
Subscriber access provided by University of Winnipeg Library

Article

Charge Dispersion and Its Effects on Reactivity of Thiamin-Derived Breslow Intermediates Michael Bielecki, Graeme W. Howe, and Ronald Kluger Biochemistry, Just Accepted Manuscript • DOI: 10.1021/acs.biochem.8b00463 • Publication Date (Web): 01 Jun 2018 Downloaded from http://pubs.acs.org on June 1, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 9 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biochemistry

Charge Dispersion and Its Effects on the Reactivity of ThiaminDerived Breslow Intermediates Michael Bielecki, Graeme W. Howe and Ronald Kluger* Davenport Chemistry Laboratories, Department of Chemistry, University of Toronto, Toronto, Ontario M5S 3H6 Canada Supporting Information Placeholder

ABSTRACT:

The enzymic decarboxylation of 2-ketoacids proceed via their C2-thiazolium adducts of thiamin diphosphate (ThDP). Loss of CO2 from these adducts leads to reactive species that are known as Breslow intermediates. The protein-bound adducts of the 2-ketoacids and ThDP are several orders of magnitude more reactive than the synthetic analogues in solution. Studies of enzymes are consistent with formulation of protein-bound Breslow intermediates with localized carbanionic character at the reactive C2α position, reflecting the charge-stabilized transition state that leads to this form. Our present study reveals that nonenzymic decarboxylation of the related thiamin adducts proceeds to the alternative charge-dispersed enol form of the Breslow intermediate. These differences suggest that the greatly enhanced rate of decarboxylation of the precursors to Breslow intermediates in enzymes arise from maintenance of carbanionic character at the position from which the carboxyl group departs, avoiding charge dispersion by stabilizing electrostatic interactions with the protein as formulated by Warshel. Applying Guthrie’s “no-barrier” addition to Marcus theory also leads to the conclusion that maintaining the tetrahedral carbanion at C2α of the resulting adduct minimizes associated kinetic barriers by avoiding re-hybridization as part of steps to and from the intermediate. Finally, maintenance of the reactive energetic carbanion agrees with the concepts of Albery and Knowles as the outcome of evolved enzymic processes.

INTRODUCTION Enzymes that catalyze the decarboxylation of 2ketoacids utilize the cofactor thiamin diphosphate (ThDP) to form covalent adducts of their substrates.1 This induces the necessary reversal of polarity to allow accumulation of electron density from loss of CO2 at what had been the carbonyl carbon. However, while synthesized adducts of 2-ketoacids and thiamin undergo decarboxylation in neutral solutions, those rates are considerably slower than for similar enzyme-

1 ACS Paragon Plus Environment

Biochemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

bound adducts of ThDP.1-3 The immediate product from loss of CO2 is a reactive species known as a Breslow intermediate and can be represented either as an enol or zwitterionic carbanion (Scheme 1). Breslow used chemical models to deduce the nature of the intermediates in the decarboxylation of 2-ketoacids in enzymes that arise upon loss of CO2. The term has been generalized to include related adducts derived from Nheterocyclic carbenes in which the heterocycle is derived from a diazole or triazole rather than from the thiazole (“enol” shown in Scheme 1.4-6

decarboxylation of mandelylthiamin (MTh, Scheme 2: Z = H), a benzoylformate-thiamin adduct, will reflect the degree of charge build-up at the transition state and correlate with the resulting energetic Breslow intermediate. This forms the basis for comparison of transitions state character with the previously reported ρ value of 4.4 for an analogous enzymic decarboxylation.18 Thus, we prepared a series of phenyl-substituted analogues of MTh, accurate comparators for the corresponding enzymic predecarboxylation intermediates derived from ThDP (Scheme 1, R=Ph). Our improved synthetic route to MTh serves as the basis for producing phenyl-substituted derivatives of MTh (Scheme 2: Z = p-OCH3; p-CH3; p-F; p-Cl; m-Br).19, 20 The results of kinetic studies on these compounds allow us to make a confident assessment of the charge distribution in the Breslow intermediate and TS that leads to it. We find that the results are consistent with an induced electrostatic environment on the enzyme that maintains a reactive intermediate that minimizes surrounding intrinsic barriers. The results are consistent with proposals for evolved catalytic efficiency.21, 22

Scheme 1. General scheme for formation of Breslow intermediates from ThDP and 2-ketoacids.

Structures formulated as enols derived from adducts of thiamin diphosphate (as in Scheme 1) have been assigned from X-ray crystallographic and spectroscopic analyses.7-11 However, other studies of the structure and reactivity of enzymic Breslow intermediates bound to ThDP-dependent enzymes conclude that enzymestabilized intermediates are present as chargeseparated carbanions (“carbanion “ in Scheme 1).12-16 Formation of charged intermediates implies a polar transition state (TS) for the decarboxylation step. The rate constant for the enzymic decarboxylation step from these intermediates is at least six orders of magnitude larger than that for the corresponding step in the nonenzymic decarboxylation of adducts of 2-ketoacids and thiamin.1 These results led us to consider to what extent an enzyme maintains the carbanionic form of the Breslow intermediate and what catalytic advantage this provides. Reports from Berkessel reveal that Breslow intermediates (derived from 1,3-diazoliums) are enols in the crystalline state.5, 6 A current report of NMR analysis of thiazolium-derived Breslow intermediates is consistent with enolic character.17 Finally, the critical question as to extent to which charges remain localized in decarboxylation transition states are difficult to assess from structural studies of products. In order to understand the basis of the substantial acceleration provided by association of ThDP adducts with proteins, we need to know the charge distribution in the transition state for a comparable species in solution. The magnitude of the Hammett ρ for the

Scheme 2. MTh spontaneously undergoes decarboxylation in neutral solutions to produce the corresponding Breslow intermediate and CO2.

Materials/Experimental Details Synthesis Methyl benzoylformate was from commercial sources. Substituted derivatives of methyl benzoylformate were synthesized from commercially available substituted acetophenones by oxidation with selenium dioxide in pyridine,23 followed by esterification with dimethyl sulfate. Liquid isobutylene was prepared from acidic dehydration of t-butanol. This was used to produce tbutyl esters with sulfuric acid as a catalyst.24 Esters of the phenyl-substituted MTh derivatives were synthesized by condensing the substituted benzoylformate esters with O-t-butyl-diphenylsilylthiamin chloride.20 The generalized synthetic route is in Scheme 2 (see SI for details). Kinetic and product studies The rates of decarboxylation of MTh derivatives were monitored for reactions in pH 7.0 HEPES buffers (0.10, 0.20, 0.30 and 0.40 M) at 25.0 (±0.1) o C with ionic strength maintained at 1.00 M with potassium chloride. The reaction progress was followed by monitoring changes in UV-VIS spectra. Data were obtained in

2 ACS Paragon Plus Environment

Page 2 of 9

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biochemistry

triplicate. The observed first order rate coefficients were calculated from non-linear regression analysis of the absorbance at 330 nm from PTK. The changes in absorbance with time were fit to the integrated first order rate law to obtain first order rate constants. Repeated wavelength scans gave stable isosbestic points over the course of the reactions. Rate constants at zero buffer concentration are from extrapolation of plots of observed rate coefficients vs. buffer concentration. 1H NMR, MS in larger scale runs and UVVis spectra provided a basis for assigning structures.

Results and Discussion We extended our previously reported method for the efficient preparation of the adduct of thiamin and benzoylformate, MTh, to produce phenyl-substituted derivatives as shown in Scheme 3.

25° ) of the resulting Breslow intermediate.28 Thus, the rate of appearance of the fragmentation products provides data for the rate of the earlier decarboxylation step.28 A plot of log (kX/kH) as a function of σ gives ρ = 0.6(±0.1) with R2 = 0.89 (Figure 2). 1 0.8 0.6 Log(kx /kH)

Page 3 of 9

0.4 0.2 0 -0.2 -0.4 -0.4

-0.2

0

0.2

0.4

σ

Figure 2. Hammett plot of first order rate constants for decarboxylation of MTh and derivatives at 25 oC.

Scheme 3. Preparation of phenyl-substituted derivatives of mandelylthiamin.

MTh and its phenyl-substituted derivatives undergo rate-limiting decarboxylation followed by rapid protonation at C2α to give derivatives of 2-(1hydroxylbenzyl)thiamin (HBnTh, Figure 1). Unlike the enzymic process, the reactions in solution form additional products upon loss of CO2, principally the result of the Oka fragmentation to substituted phenyl thiazole ketones (PTKs) and 2,5-dimethyl-4-amino-1,3pyrimidine (DMAP) (Figure 1).25.26, 27

Figure 1. Products from decarboxylation of MTh and derivatives in neutral buffers include those from fragmentation.

The absorbance near 330 nm arises from phenylsubstituted derivatives of PTK, the common product of fragmentation of the Breslow intermediates.25 The rate constant for decarboxylation of MTh (k~10-4 s-1) is considerably smaller than the rate constant for subsequent fragmentation (k~105 art 40°C and 104 at

The effects of substituents on rates of decarboxylation reflect the extent of carbanion character that develops at C2α in the transition state as the C-C bond cleaves. Since the enzymic results are consistent with the formation of a localized carbanion, the substituent effects that we have determined provide a comparative basis for the corresponding reactions in solution. Applying the Hammond postulate enables us to deduce the extent to which charge remains localized at C2α in the energetic Breslow intermediates, despite their transient existence. Decarboxylation leading to the immediate carbanion in the Breslow intermediate rather than to the subsequent enol would result in a much larger sensitivity to substituents. Values of ρ > 3 are typical for E1cb reactions with rate-determining formation of carbanions.29, 30 Transfers of benzylic protons give ρ = 4.031 and the anionic polymerization of styrene gives ρ = 5.32 In contrast, the deprotonation of 1phenylnitromethane, in which the developing charge is internally neutralized, has ρ = 1.2.33 In that case, there is a lag in charge delocalization relative to the initial charge development (imperfect synchronization34). Still, the value of ρ is small in comparison to that for production of ultimately localized carbanions. Thus, we conclude that the ρ value of 0.6 for the decarboxylation of MTh is clearly significantly smaller than expected for reactions that produce carbanions. Thus, we conclude that the extent of charge delocalization at the transition state for the departure of CO2 is very extensive in solution and that this leads to the uncharged enol-like character of the reactive Breslow intermediate. Studies of the enzymic decarboxylations of analogous adducts of ThDP and phenyl-substituted benzoylformates provide evidence for a dramatically

3 ACS Paragon Plus Environment

Biochemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

different form of the Breslow intermediate from that in solution. Schutz et al. reported results that they present as a Hammett plot for the decarboxylation of parasubstituted phenyl derivatives of benzoylformate by IPDC (indole pyruvate decarboxylase).18 That enzyme efficiently promotes the decarboxylation of benzoylformate and its para-substituted analogues. The Hammett plot of kcat vs. σ is non-linear with concave downward curvature at its midpoint. The authors reasonably ascribe this curvature to a change in ratelimiting step due to the effects of substituents on a reactive intermediate. The section with a positive slope (ρ = 4.4) occurs with substrates with electron-donating substituents (methyl, ethyl, methoxy). The portion with a negative slope (ρ = -2.5) occurs with substrates with electron-withdrawing substituents (fluoro, bromo, chloro, nitro). The ρ value of 4.4 is substantial and is consistent with a transition state leading to the localized carbanion of the charge-separated intermediate. Of the various steps in the catalytic pathway of ThDP-dependent decarboxylases, only the decarboxylation leading to a localized carbanion could account for such a large value. The loss of CO2 from electron-rich para-substituted benzoylformates on benzoylformate decarboxylase (BFDC) is partially ratelimiting as deduced from carbon-isotope effects.35 Therefore, a rate-limiting decarboxylation step in IPDC with electron-rich substrates is reasonable. In IPDC, kcat for the bulkier and less electron-donating p-CH3CH2benzoylformate is larger than kcat for the smaller but stronger electron-donating p-CH3-benzoylformate. This indicates that the observed substituent effects are primarily the result of electronic interactions. The value of ρ = 4.4 for the enzymic process contrasts with the non-enzymic ρ = 0.6 for decarboxylation of derivative of MTh that we have now determined. Since the carbanion is formed on the enzyme, it is reasonable that electrostatic interactions contribute to stabilizing the enzyme-bound state.36 This is consistent with direct observations of localized carbanion-like intermediates on enzymes.12-16 In solution, localized carbanions would be considerably higher in energy than the readily accessed enol that forms upon relaxation. Furthermore, in the enzymic cases derived from ThDP, the presence of the diphosphate facilitates binding without interfering with catalysis. In solution, the diphosphate provides no role for a catalytic effect.37, 38 As noted above, the protein-bound form favors a path to an otherwise more energetic form of the Breslow intermediate. Warshel has elucidated the importance of electrostatic stabilization of transition states and reactive intermediates that occur through relief of electrostatic stress within the associated protein.39-42 This would allow the localized carbanion to be maintained upon cleavage of the C-C bond that leads to formation of CO2. Warshel’s formulation establishes the overall importance of significant electrostatic effects for

catalytic processes within proteins that utilize release of locally induced stress.43, 44 Why would an enzyme evolve to stabilize the intrinsically more energetic form of the Breslow intermediate? An important insight comes from reports by Kresge and coworkers who observed that transfer of a proton to a localized carbanion has no intrinsic kinetic barrier (following Marcus theory), while transfer to sites with dispersed electron density overcome substantial intrinsic barriers.45 The effects of localization of charge on reactivity are not limited to proton transfers and can be understood from the basis of Guthrie’s No Barrier theory: processes that alter bond lengths and bond angles while other bonds are being formed or broken are the source of intrinsic barriers.46, 47 Interestingly, Guthrie specifically applied no-barrier theory to decarboxylation reactions in terms of the effects of the conversion of the carboxyl group to carbon dioxide.48 The change in hybridization and the requisite concerted atomic movements that avoid impossible intermediates in single-movement steps produce a substantial portion of the intrinsic barrier for formation of CO2 in a decarboxylation process. We extend the application No Barrier theory to the residual structure in decarboxylation. This leads to a conclusion that parallels Kresge’s analysis of the source of barriers to proton transfers from carbon. That is, loss of a carboxyl that leads to a localized carbanion will have a lower intrinsic barrier than those where electron density is delocalized to an enolic species. According to Albery and Knowles, catalysis of an elementary step, as we see here, is the most sophisticated adaptation in an enzyme's evolution.49 Warshel’s formulation of stressed charges in proteins provides efficient electrostatic stabilization of carbanions (and associated transition states)50 that would be dispersed in solution through rehybridization and general solvation (Figure 3).

Figure 3. More O’Ferrall-Guthrie plot for consideration of pathways for decarboxylation following No-Barrier Theory:” C-C cleavage without changes in hybridization of residual species (lower left, “enzyme”), producing a minimal intrinsic barrier. Direct formation of the neutral product (lower right “no enzyme”) requires both rehybridization and C-C cleavage with a high intrinsic barrier. The alternative stepwise route (via

4 ACS Paragon Plus Environment

Page 4 of 9

Page 5 of 9 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biochemistry

upper right, re-hybridization to sp2, then C-C cleavage) would produce a thermally inaccessible intermediate With the reactive carbanion bound to the enzyme in a region with positive electrostatic character, the thermodynamic barrier in the transition state that produces such an intermediate can be overcome by attractive interactions that overcome electrostatic stress within the protein following the general principles of enzymic catalysis that have been developed by Warshel (Figure 4).36, 39, 51, 52

intermediates form via transition states that resemble charge-dispersed enols. These structures contrast with the corresponding charge-localized carbanionic Breslow intermediates that bind to enzymes. Proteins can provide environments that favor the zwitterionic carbanion over the enol according to the Warshel formulation.40 Importantly, the pathway via a zwitterionic carbanion would lead to acceleration of the decarboxylation process by minimizing the associated intrinsic Marcus barrier associated with rehybridization at C2α as proposed in general by Guthrie.55 Maintaining the zwitterionic form also retains the aromatic state of its thiazolium component and prevents the otherwise unavoidable fragmentation that occurs in solution. Finally, the retention of separated charges exemplifies how enzymes evolve to maximize efficiency by maintaining energetic intermediates through structural control in their bound state that would be lost in the unbound state in solution as summarized in Figure 4.21

■ AUTHOR INFORMATION Corresponding Author

Figure 4. A hypothetical energy profile for the enzymic decarboxylation of 2-ketoacids via formation of the enol (dashed line) and via the localized carbanion (solid line).

In addition, examination of the charged form of the Breslow intermediate might illustrate why the Oka fragmentation of the ThDP derivatives of benzoylformates does not occur on enzymes. If the fragmentation occurs via the conjugate base of the enolic C2α-OH group as we have proposed,20 the required ionization of the C2α-OH for fragmentation would not occur from the localized carbanion due to internal electrostatic repulsion.53, 54 The clear contrast in the nature and resulting reactivity of enzymic and nonenzymic Breslow intermediates that arise from differences in charge dispersion would have provided an evolutionary advantage for intermediates with protein-stabilized localized electrostatic character.

*Davenport Chemistry Laboratories, Department of Chemistry, University of Toronto, 80 St. George Street, Toronto, Ontario M5S 3H6 Canada. E-mail: [email protected]. Notes The authors declare no competing financial interest

FUNDING The authors are grateful for funding from NSERC Canada through a Discovery Grant (RGPIN-201604993) to R.K. and a Postgraduate Scholarship-Doctoral to G.W.H.

■ DEDICATION We dedicate this in memory of Professors Ronald Breslow and Peter Guthrie. They led by example.

■ Conclusions The small value of ρ for the spontaneous decarboxylation of phenyl-substituted derivatives of MTh in solution indicates that the resulting Breslow

5 ACS Paragon Plus Environment

Biochemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 9

■ REFERENCES [1] Kluger, R., and Tittmann, K. (2008) Thiamin diphosphate catalysis: Enzymic and nonenzymic covalent intermediates, Chem. Rev. 108, 1797-1833. [2] Breslow, R. (1958) Mechanism of thiamine action. Iv. Evidence from studies on model systems, J. Am. Chem. Soc. 80, 3719-3726. [3] Hu, Q., and Kluger, R. (2002) Reactivity of intermediates in benzoylformate decarboxylase: Avoiding the path to destruction, J. Am. Chem. Soc. 124, 14858-14859. [4] Berkessel, A., Elfert, S., Etzenbach-Effers, K., and Teles, J. H. (2010) Aldehyde umpolung by Nheterocyclic carbenes: NMR characterization of the breslow intermediate in its keto form, and a spiro-dioxolane as the resting state of the catalytic system, Angew. Chem. Int. Ed. Engl. 49, 7120-7124. [5] Berkessel, A., Elfert, S., Yatham, V. R., Neudoerfl, J.-M., Schloerer, N. E., and Teles, J. H. (2012) Umpolung by N-heterocyclic carbenes: Generation and reactivity of the elusive 2,2-diamino enols (Breslow intermediates), Angew. Chem., Int. Ed. 51, 12370-12374. [6] Berkessel, A., Yatham, V. R., Elfert, S., and Neudorfl, J.-M. (2013) Characterization of the key intermediates of carbene-catalyzed umpolung by NMR spectroscopy and X-ray diffraction: Breslow intermediates, homoenolates, and azolium enolates, Angew. Chem., Int. Ed. Engl. 52, 11158-11162. [7] Sergienko, E. A., Wang, J., Polovnikova, L., Hasson, M. S., McLeish, M. J., Kenyon, G. L., and Jordan, F. (2000) Spectroscopic detection of transient thiamin diphosphate-bound intermediates on benzoylformate decarboxylase, Biochemistry. 39, 13862-13869. [8] Nakai, T., Nakagawa, N., Maoka, N., Masui, R., Kuramitsu, S., and Kamiya, N. (2004) Ligandinduced conformational changes and a reaction intermediate in branched-chain 2-oxo acid dehydrogenase (E1) from thermus thermophilus HB8, as revealed by X-ray crystallography, J. Mol. Biol. 337, 1011-1033. [9] Zhang, S., Liu, M., Yan, Y., Zhang, Z., and Jordan, F. (2004) C2-α-lactylthiamin diphosphate is an intermediate on the pathway of thiamin diphosphate-dependent pyruvate decarboxylation: Evidence on enzymes and models, J. Biol. Chem. 279, 54312-54318. [10] Arjunan, P., Sax, M., Brunskill, A., Chandrasekhar, K., Nemeria, N., Zhang, S., Jordan, F., and Furey, W. (2006) A thiamin-bound, pre-decarboxylation reaction intermediate analogue in the pyruvate dehydrogenase e1 subunit induces large scale disorder-to-order transformations in the enzyme and reveals novel structural features in the covalently bound adduct, J. Biol. Chem. 281, 15296-15303. [11] Chakraborty, S., Nemeria, N. S., Balakrishnan, A., Brandt, G. S., Kneen, M. M., Yep, A., McLeish, M. J., Kenyon, G. L., Petsko, G. A., Ringe, D., and Jordan, F. (2009) Detection and time course of formation of major thiamin diphosphate-bound covalent intermediates derived from a chromophoric substrate analogue on benzoylformate decarboxylase, Biochemistry 48, 981-994. [12] Machius, M., Wynn, R. M., Chuang, J. L., Li, J., Kluger, R., Yu, D., Tomchick, D. R., Brautigam, C. A., and Chuang, D. T. (2006) A versatile conformational switch regulates reactivity in human branched-chain α-ketoacid dehydrogenase, Structure (Cambridge, MA, U. S.) 14, 287-298. [13] Berthold, C. L., Toyota, C. G., Moussatche, P., Wood, M. D., Leeper, F., Richards, N. G. J., and Lindqvist, Y. (2007) Crystallographic snapshots of oxalyl-CoA decarboxylase give insights into catalysis by nonoxidative ThDP-dependent decarboxylases, Structure (Cambridge, MA, U. S.) 15, 853-861. [14] Wagner, T., Barilone, N., Alzari, P. M., and Bellinzoni, M. (2014) A dual conformation of the postdecarboxylation intermediate is associated with distinct enzyme states in mycobacterial KGD (αketoglutarate decarboxylase), Biochem. J. 457, 425-434. [15] Song, H., Dong, C., Qin, M., Chen, Y., Sun, Y., Liu, J., Wan, C., and Guo, Z. (2016) A thiaminedependent enzyme utilizes an active tetrahedral intermediate in vitamin K biosynthesis, J. Am. Chem. Soc. 138, 7244-7247. ACS Paragon Plus Environment

6

Page 7 of 9 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biochemistry

[16] Meyer, D., Neumann, P., Koers, E., Sjuts, H., Luedtke, S., Sheldrick, G. M., Ficner, R., and Tittmann, K. (2012) Unexpected tautomeric equilibria of the carbanion-enamine intermediate in pyruvate oxidase highlight unrecognized chemical versatility of thiamin, Proc. Natl. Acad. Sci. U. S. A. 109, 10867-10872, S10867/10861-S10867/10827. [17] Berkessel, A., Paul, M., Sudkaow, P., Wessels, A., Schlorer, N. E., and Neudorfl, J. M. (2018) Breslow intermediates from aromatic N-heterocyclic carbenes (benzimidazolin-2-ylidenes, thiazolin-2-ylidenes), Angew. Chem., Int. Ed. Engl. [18] Schutz, A., Golbik, R., Tittmann, K., Svergun, D. I., Koch, M. H. J., Hubner, G., and Konig, S. (2003) Studies on structure-function relationships of indolepyruvate decarboxylase from enterobacter cloacae, a key enzyme of the indole acetic acid pathway, Eur. J. Biochem. 270, 2322-2331. [19] Bielecki, M., Howe, G. W., and Kluger, R. (2015) Lithium-stabilized nucleophilic addition of thiamin to a ketone provides an efficient route to mandelylthiamin, a critical pre-decarboxylation intermediate, Bioorg. Chem. 62, 124-129. [20] Bielecki, M., and Kluger, R. (2017) The need for an alternative to radicals as the cause of fragmentation of a thiamin-derived breslow intermediate, Angew. Chem. Int. Ed. Engl. 56, 63216323. [21] Albery, W. J., and Knowles, J. R. (1976) Evolution of enzyme function and development of catalytic efficiency, Biochemistry, 15, 5631-5640. [22] Burbaum, J. J., Raines, R. T., Albery, W. J., and Knowles, J. R. (1989) Evolutionary optimization of the catalytic effectiveness of an enzyme, Biochemistry 28, 9293-9305. [23] Zhuang, J., Wang, C., Xie, F., and Zhang, W. (2009) One-pot efficient synthesis of aryl α-keto esters from aryl-ketones, Tetrahedron. 65, 9797-9800. [24] Hurd, C. D., and Blunck, F. H. (1929) The pyrolysis of hydrocarbons: Isobutylene, J. Am. Chem. Soc. 51, 3561-3572. [25] Kluger, R., Lam, J. F., Pezacki, J. P., and Yang, C.-M. (1995) Diverting thiamin from catalysis to destruction. Mechanism of fragmentation of N(1')-methyl-2-(1-hydroxybenzyl)thiamin, J. Am. Chem.Soc. [26] Oka, Y., Kishimoto, S., and Hirano, H. (1970) Vitamin B1 and related compounds. CIX. A novel cleavage of thiamin and its homologs by the reaction with aromatic aldehydes, Chem. Pharm. Bull. 18, 527-533. [27] Oka, Y., Kishimoto, S., and Hirano, H. (1970) Vitamin B1 and related compounds. CX. Rearrangements of α-hydroxybenzylthiamin and its homologs, Chem. Pharm. Bull 18, 534-541. [28] Kluger, R., and Moore, I. F. (2000) Destruction of vitamin B1 by benzaldehyde. Reactivity of intermediates in the fragmentation of N1'-benzyl-2-(1-hydroxybenzyl)thiamin, J. Am. Chem. Soc. 122, 6145-6150. [29] Baker, R., and Spillett, M. J. (1969) Elcb nature of the transition state in the elimination reaction of phenethyl sulfoxides in basic media, J. Chem. Soc. B, 481-484. [30] DePuy, C. H., Morris, G. F., Smith, J. S., and Smat, R. J. (1965) Electronic effects of elimination reactions. V. Bimolecular cis eliminations. 2-arylcyclopentyl tosylates, J. Am. Chem. Soc. 87, 2421-2428. [31] Streitwieser, A., Jr., and Koch, H. F. (1964) Acidity of hydrocarbons. X. Exchange rates of ringsubstituted toluene-α-t's with lithium cyclohexylamide in cyclohexylamine, J. Am. Chem. Soc. 86, 404-409. [32] Shima, M., Bhattacharyya, D. N., Smid, J., and Szwarc, M. (1963) Hammett's relations in anionic copolymerizations, J. Am. Chem. Soc. 85, 1306-1310. [33] Bordwell, F. G., and Boyle, W. J., Jr. (1972) Acidities, Broensted coefficients, and transition state structures for 1-arylnitroalkanes, J. Am. Chem. Soc. 94, 3907-3911. [34] Bernasconi, C. F. (1985) The principle of imperfect synchronization .1. Ionization of carbon acids, Tetrahedron 41, 3219-3234. ACS Paragon Plus Environment

7

Biochemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 9

[35] Weiss, P. M., Garcia, G. A., Kenyon, G. L., Cleland, W. W., and Cook, P. F. (1988) Kinetics and mechanism of benzoylformate decarboxylase using carbon-13 and solvent deuterium isotope effects on benzoylformate and benzoylformate analogs. 27, 2197-2205. [36] Warshel, A., Sharma, P. K., Kato, M., Xiang, Y., Liu, H., and Olsson, M. H. M. (2006) Electrostatic basis for enzyme catalysis, Chem Rev 106, 3210-3235. [37] Kluger, R., Chin, J., and Smyth, T. (1981) Thiamin-catalyzed decarboxylation of pyruvate. Synthesis and reactivity analysis of the central, elusive intermediate, α-lactylthiamin, J. Am. Chem. Soc. 103, 884-888. [38] Kluger, R., and Smyth, T. (1981) Interaction of pyruvate-thiamin diphosphate adducts with pyruvate deacrboxylase. Catalysis through "closed" transition states, J.Am. Chem. Soc. 103, 1214-1216. [39] Warshel, A., Florián, J., Štrajbl, M., and Villà, J. (2001) Circe effect versus enzyme preorganization: What can be learned from the structure of the most proficient enzyme?, ChemBioChem 2, 109-111. [40] Warshel, A., Sharma, P. K., Kato, M., Xiang, Y., Liu, H., and Olsson, M. H. (2006) Electrostatic basis for enzyme catalysis, Chem. Rev. 106, 3210-3235. [41] Warshel, A., Štrajbl, M., Villà, J., and Florián, J. (2000) Remarkable rate enhancement of orotidine 5‘-monophosphate decarboxylase is due to transition-state stabilization rather than to groundstate destabilization, Biochemistry 39, 14728-14738. [42] Fersht, A. R. (2013) Profile of Martin Karplus, Michael Levitt, and Arieh Warshel, 2013 Nobel laureates in chemistry, Proc. Natl. Acad. Sci. USA 110, 19656-19657. [43] Richard, J. P., Amyes, T. L., and Reyes, A. C. (2018) Orotidine 5′-monophosphate decarboxylase: Probing the limits of the possible for enzyme catalysis, Acc. Chem. Res. 51, 960-969. [44] Reyes, A. C., Amyes, T. L., and Richard, J. P. (2017) Enzyme architecture: Erection of active orotidine 5′-monophosphate decarboxylase by substrate-induced conformational changes, J. Am. Chem. Soc. 139, 16048-16051. [45] Kresge, A. J. (1975) What makes proton transfer fast, Acc. Chem. Res. 8, 354-360. [46] Guthrie, J. P. (1996) Multidimensional Marcus theory: An analysis of concerted reactions, J. Am. Chem. Soc. 118, 12878-12885. [47] DiRocco, D. A., Oberg, K. M., and Rovis, T. (2012) Isolable analogues of the Breslow intermediate derived from chiral triazolylidene carbenes, J. Am. Chem. Soc. 134, 6143-6145. [48] Guthrie, J. P. (2002) Uncatalyzed and amine catalyzed decarboxylation of acetoacetic acid: An examination in terms of no barrier theory, Bioorg. Chem. 30, 32-52. [49] Albery, W. J., and Knowles, J. R. (1976) Evolution of enzyme function and the development of catalytic efficiency, Biochemistry. 15, 5631-5640. [50] Guthrie, J. P., and Kluger, R. (1993) Electrostatic stabilization can explain the unexpected acidity of carbon acids in enzyme-catalyzed reactions, J. Am. Chem. Soc. 115, 11569-11572. [51] Warshel, A. (1978) Energetics of enzyme catalysis, Proc. Natl. Acad. Sci. U. S. A. 75, 5250-5254. [52] Warshel, A., and Levitt, M. (1976) Theoretical studies of enzymic reactions: Dielectric, electrostatic and steric stabilization of the carbonium ion in the reaction of lysozyme, J. Mol. Biol. 103, 227-249. [53] Kirkwood, J. G., and Westheimer, F. H. (1938) The electrostatic influence of substituents on the dissociation constants of organic acids. III, J. Chem. Phys. 6, 506-512. [54] Simonson, T., and Roux, B. (2016) Concepts and protocols for electrostatic free energies, Mol. Simul. 42, 1090-1101. [55] Guthrie, J. P. (2011) No barrier theory and the origins of the intrinsic barrier, Adv. Phys. Org. Chem. 45, 171-220.

ACS Paragon Plus Environment

8

Page 9 of 9 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biochemistry

the greatly enhanced rate of decarboxylation of the precursors to Breslow intermediates in enzymes arise from maintenance of carbanionic character at the position from which the carboxyl group departs

ACS Paragon Plus Environment