Multidimensional Reaction Dynamics Reveal How the Enzyme TcTS

Aug 15, 2016 - We use multidimensional QM/MM free energy computations to reveal a competition between a minor elimination reaction and the dominant ...
0 downloads 5 Views 2MB Size
Subscriber access provided by Northern Illinois University

Article

Multidimensional Reaction Dynamics Reveal How the Enzyme TcTS Suppresses Competing Side Reactions and their Side Products Ian L. Rogers, and Kevin J. Naidoo ACS Catal., Just Accepted Manuscript • DOI: 10.1021/acscatal.6b01522 • Publication Date (Web): 15 Aug 2016 Downloaded from http://pubs.acs.org on August 23, 2016

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 free 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 accessible to all readers and 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.

ACS Catalysis 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 27

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

ACS Catalysis

Multidimensional Reaction Dynamics Reveal How the Enzyme TcTS Suppresses Competing Side Reactions and their Side Products

Ian L. Rogers1,2 and Kevin J. Naidoo1,2* 1.Scientific Computing Research Unit, University of Cape Town, Rondebosch 7701, South Africa 2. Department of Chemistry, University of Cape Town, Rondebosch 7701, South Africa

Correspondence: Kevin J. Naidoo ([email protected])

1 ACS Paragon Plus Environment

ACS Catalysis

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

Abstract The suppression of competing reactions that lead to side products is one of the key mechanistic actions defining enzyme catalysis. The transfer of sialic acid (SA) from host sialyl glycoconjugates to galactosyl glycoconjugates in a water solution is susceptible to two competing side reactions, but a single product is the outcome in the glycosylation and deglycosylation steps of the trans-sialidase (TS) in the Trypanosoma cruzi (T. cruzi) parasite, generally known as TcTS. We use multidimensional QM/MM free energy computations to reveal a competition between a minor elimination reaction and the dominant displacement reaction present in both steps. The simultaneous monitoring of the progression of the competing reactions reveal lower barriers in the free energy profiles, greater sampling of favorable reactant stereoelectronic alignments and a greater number of possible transition paths leading to successful crossing reaction trajectories for the dominant displacement reactions compared with that of the elimination reactions. KEYWORDS: Transition State Theory, Reaction Dynamics, Free Energy, FEARCF, glycosylation, Sialidase mechanisms, Glycosyltransferase, Enzyme Selectivity

Introduction The activation free energy is of central importance to the phenomenological understanding of rate-acceleration and selectivity observed for catalyzed reactions compared with reference reactions in solution. In selectivity, kinetic considerations distinguish between pathways restricting the possible reaction paths that result in singular products. Consider Trypanosoma cruzi trans-Sialidase (TcTS) that catalyzes the transfer of α(2-3)-linked sialic acid (SA) from host sialyl glycoconjugates to T. cruzi β-galactosyl glycoconjugates. In this enzyme, the complete two-step transferase reaction is susceptible to two side reactions that are

2 ACS Paragon Plus Environment

Page 2 of 27

Page 3 of 27

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

ACS Catalysis

rooted in the native uncatalyzed reaction scheme in solution (Scheme 1). In the first side reaction, hydrolysis of the covalent intermediate (CI) by an active site water molecule competes with the primary transfer of SA to an acceptor glycoconjugate.1 The second side reaction is the elimination of a SA H-3 proton to form 2-deoxy-2,3-didehydro-N-acetylneuraminic acid (DANA) and has been detected for prolonged incubation of α(2–3)-sialyllactose (SL) with TcTS.2 More generally, this competition between elimination side reactions and displacement main reactions has been observed in other sialidases such as those expressed in Influenza B virus3, Vibrio cholera4 and Streptococcus pneumonia 5 Elimination ostensibly proceeds via H-3 abstraction by a mediating active site water molecule,6 or alternatively by an appropriately located catalytic acid.5 In contrast to the above enzymatic reactions, where the elimination reaction is observed to be in minor competition to the main displacement reaction, DANA formation readily reduces the synthetic yield of SA glycosylation in solution.7 Here, the anomeric carboxylate group destabilizes the oxocarbenium ion through an inductive electron withdrawing effect and approach of the nucleophile to SA C-2 is hindered by steric constraints. As a result, competing DANA formation via elimination is promoted. Therefore, the detection of DANA in only trace amounts for excess TcTS incubated with SL points to the ability of the trans-sialidase to mitigate the unfavorable side reaction. Characterization of the elimination reaction at an atomistic level allows a unique opportunity to examine how enzymes diminish the success of competing reactions. TcTS is a critical enzyme in T. cruzi’s survival8 upon which a wide range of experimental and computational methods have been used to investigate the

3 ACS Paragon Plus Environment

ACS Catalysis

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

Scheme 1 The putative catalytic mechanism for TcTS follows a catalytic Ping-Pong mechanism with the formation of a CI. The primary transferase activity is susceptible to two side reactions: hydrolysis of the CI to SA, and elimination to DANA. The atom numbering of the SA ring is given for reference (bottom left).

glycosyltransferase action. Although some debate surrounding the precise details of the TcTS catalyzed transfer mechanism remains,9,10 the prevailing opinion is in favor of a classic Ping-Pong mechanism in which the protein is glycosylated to form the CI and Asp59 serves as the catalytic acid/base (Scheme 1).8 A classic Ping-Pong mechanism is consistent with the absence of a distinct acceptor binding site in crystal structures,11 and has been supported by kinetic and chemical rescue experiments.12 These spectroscopic and assay experiments have been used to characterize binding, identify enzymatic side products, determine the rate limiting step and calculate reaction rates for the overall TcTS-catalyzed reaction. Nonetheless, a complete characterization of the reaction itinerary using only kinetic data is not possible. For instance, KIEs measured in a competitive fashion characterize only up to the first irreversible step13 and NMR spectroscopy 4 ACS Paragon Plus Environment

Page 4 of 27

Page 5 of 27

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

ACS Catalysis

is limited to time averages and requires deconvolution of relevant peaks from the complex protein environment. Furthermore, deciphering the mechanistic details of reactions from these measurements is complicated by the relatively gross nature of the experimental spatial and temporal resolution. Equally, analyses of X-Ray diffraction patterns are cumbersome involving multistep synthesis of carbohydrate-related substrates that will inhibit the enzyme, and in so doing provide representative snapshots along the native reaction path. Computational models, for example Molecular Dynamics (MD) simulations, have been used to show that SA -binding, and induced conformational changes, both promotes the formation of a catalytically competent Tyr342 – Glu230 conformation,14and mitigates the hydrolysis side reaction by reducing the solvent exposure of the catalytic site.15 QM/MM free energy calculations revealed that selectivity is further explained by the ~7 kcal/mol difference in energy barrier between these chemical steps.14 Despite experimental observations of the DANA side product in a number of sialidases, there is an absence of a detailed molecular understanding of this competition. Here we interrogate the molecular details of TcTS’ suppression of viable competing reactions under physiological conditions to assert the precision and selectivity synonymous with enzyme catalysis. In the presence of lactose acceptor, the β-Galactose will be hydrogen bonded to Asp59, and abstraction of the bottom-facing SA H-3 in the deglycosylation step will be by the acceptor molecule’s O-3 oxygen, analogous to the water-mediated mechanism recorded by Jongkees and Withers.6

5 ACS Paragon Plus Environment

ACS Catalysis

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

Results and Discussion TcTS’ catalytic itinerary was explored using free energies calculated from probability histograms along an appropriate progress coordinate, ξ. The progress coordinate, used synonymously with the reaction coordinate, is commonly defined as a linear combination of geometric coordinates that gives a measure of the reaction progress. In this work the two- or three-dimensional ξs comprise discretized bins that are populated by points taken from multiple reaction dynamics trajectories that are run in parallel in each FEARCF iteration (see Methods for details). We have mapped each breaking or forming bond to an element of ξi, where a 2-D (ξ1, ξ2) results in a free energy surface (FES) while a 3-D (ξ1, ξ2, ξ3) gives a free energy volume (FEV). An initial FES describing CI formation, resolved along the ξ1 = (Tyr342 O)…(SA C-2) attack and ξ2 = (SA C-2)…(Gal O-3) leaving distances, displayed a local free energy minimum located at a long glycoside bond length. The sampling of the oxocarbenium ion species at this location of the FES as the result of a higher energy reaction path is expected. Surprisingly, a number of structures from trajectories visiting this region showed the formation of DANA, in so providing evidence for the presence of a competing side reaction. Supporting this observation is the previous detection of trace amounts of DANA (the 2-deoxy-2,3-didehydro species) in kinetic studies on TcTS, albeit under hydrolyzing conditions. Although DANA is a weak inhibitor of TcTS,8c further evidence demonstrating that the trans-sialidase is capable of catalyzing SA H-3 elimination is provided by a TcTS:DANA crystal structure that corresponds to monoclinic crystals soaked with SA.11 To uncover the TcTS catalytic action resulting in the exclusive transfer of SA from host donors, it is best to monitor the elimination and displacement reactions 6 ACS Paragon Plus Environment

Page 6 of 27

Page 7 of 27

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

ACS Catalysis

simultaneously in both steps. Consequently, two FEVs that define free energy as a function of three bonds were computed. The FEVs reveal the mechanistic drivers of the elimination reactions that are competing with the dominant displacement reactions within a catalytic itinerary under physiological (non-hydrolyzing) conditions. The structural and thermodynamic results are summarized in Table 1.

Table 1. Summary of structural and thermodynamic results.

Reaction

ΔG* (kcal/mol)

MC→CI

19.5

(2.2,2.3)

MC→DANA

27.9

(1.3, 2.1)

CI →MC

20.6

CI→DANA

27.4

FEV TS (Å)

a,b

FEV TS Volume (Å3)

PES TS Structure a,b SCC-DFTB (Å)

M06-2x (Å)

0.021

(2.18, 2.22)

(2.57,2.10)

0.001

(1.26, 2.15)

(1.34, 2.44)

(2.1,2.4)

0.024

(2.27, 2.18)

(2.08, 2.57)

(1.3,2.3)

0.006

(1.30, 2.12)

(1.31, 2.34)

a Parameters for SCC-DFTB were the mio-1-1 set, and a 6-31G(d) basis set was used for DFT calculations. b Reaction coordinates are (SA-Nu, SA-LG) for substitution reactions and (SA-B, SA-LG) for elimination reactions.

Step 1: Glycosylation The competing reactions explored for the glycosylation step, Displacement 1A and Elimination 1B, are shown Scheme 2.

7 ACS Paragon Plus Environment

ACS Catalysis

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

Scheme 2 The primary reaction, Displacement 1A, that leads to TcTS glycosylation is shown

alongside the competing Elimination 1B reaction to DANA.

CI formation is characterized by an isosurface of 19.5 kcal/mol that is carved out of the FEV (Figure 1) resolved for ξ1= (Tyr342 O)…( SA C-2), ξ2 = (Tyr342 O)…(SA H31) and ξ3 = (SA C-2)…(Gal O-3). The isosurface extends from the Michaelis complex (MC) and CI regions to connect at ξ1 = 2.2 Å and ξ3 = 2.3 Å. This barrier compares directly with Umbrella sampling results (ΔGTS = 20.8 kcal/mol) where average values of 2.52 Å for ξ1 and 2.19 Å for ξ3 can be computed from the TS bin of the conflated reaction coordinates.10 In comparison, an isosurface of 27.9 kcal/mol connects the MC and DANA stationary points at ξ2 = 1.3 Å and ξ3 = 2.1 Å (Figure 1). In both the displacement and elimination reactions, dividing TS volumes separate the reactant and product stationary points suggestive of bimolecular mechanisms. 8 ACS Paragon Plus Environment

Page 8 of 27

Page 9 of 27

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

ACS Catalysis

Figure 1 FEV for glycosylation step with ξ illustrated in the molecular drawing. Isosurfaces are

shown at 21.0 kcal/mol (green) and 29.4 kcal/mol (red), and corresponding Displacement 1A (blue) and Elimination 1B (yellow) free energy paths (FEPs) are traced. SCC-DFTB free energy TSs are indicated by black dots, while the M06-2x/6-31G(d) PE TS structures shown in gray. TS volumes are shown flanking the FEV. Probability contour plots for the geometrical analysis of the 500 ps MC equilibrium trajectory are shown below at confidence intervals of 30%.

The difference of 8.4 kcal/mol in activation energies shows that CI formation via Displacement 1A is kinetically more favored. The free energy paths include entropic contributions that can be qualitatively analyzed by comparing the free energy landscape local to the TS and reactant. Isoenergetic surfaces connecting points 1.5 kcal/mol greater than the TS depict the entry channel to the TS (Figure 1). The volume encompassed by the surfaces are defined by the entropic degrees of freedom orthogonal to ߦ.16 Relative narrowing of the channels indicates an unfavorable entropy change and diminished flux of

9 ACS Paragon Plus Environment

ACS Catalysis

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

activated complexes for both reactions. However, the TS volume for the primary displacement reaction is some 20 times larger than the elimination reaction in so registering a much faster rate of reaction for CI formation. A Near Attack Conformation (NAC) type analysis was performed to probe the mechanistic detail underlying the free energy results. The geometric relationship between Tyr342 and SA was analyzed from the data taken from a 500 ps SCCDFTB/CHARMM equilibrium trajectory. For the competing reactions, the probability contours were constructed as a function of the distance of the forming bond and a proxy for the angle of attack (Figure 1). In the case of CI formation this angle is θ1 = (Tyr342 O)…(SA C-2)…(Gal O-3) and in the case of DANA formation it is θ2 = (Tyr342 O)…(SA H-31)…(SA C-3). The ξ1 bond length and θ1 angle in the MC remain tightly clustered over the 500 ps, within van der Waals contact distance and occupying a linear angle of nucleophilic attack with favorable stereoelectronic alignment. Conversely, the ξ2 bond and θ2 angle show a more diffuse distribution. Mechanistic Details of Displacement 1A The FEV is built from the total of all reaction dynamics trajectories making up each of the FEARCF iterations (see Methods). The reactant well evolves over FEARCF iterations to the product well, which allows electronic and conformational information to be extracted from the reaction trajectories. The crossing trajectory that most closely follows the Displacement 1A FEP (see Methods) shows synchronous lengthening of the glycoside bond and shortening of the (Tyr342 O)…(SA C-2) distance on a timescale shorter than 1 ps (Figure 2A). This profile and the location of the free energy TS at ξ1 = 2.2 Å and ξ3 = 2.3 Å is consistent with a mechanism positioned between an SN2 (ANDN) and a dissociative SN2 (DNAN).

10 ACS Paragon Plus Environment

Page 10 of 27

Page 11 of 27

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

ACS Catalysis

Figure 2 Mechanistic analysis of representative (A) Displacement 1A and (B) Elimination 1B

crossings. Geometric analysis of the breaking and forming bonds comprising ξ is plotted on the left along with pucker. The forming bond is colored to match the ξ definition in Figure 1, and grayed sections illustrate the TS region from which the TS structure was optimized. The NBOs for this structure are shown on the right at an electron density of -0.02 and 0.02.

The M06-2x/6-31G(d) TS structure was gained from optimizing a dynamic snapshot selected from the reactive trajectory (ξ1=2.52 Å, ξ3=2.03 Å; Table 1). The 13C

primary KIE associated with this structure was 1.025 which compares

favorably with the experimental value of 1.021 ± 0.014.17 On the other hand, the βdideuterium secondary KIE was calculated at 1.107 and shows some variation from the experimental value of 1.053 ± 0.010. The DFT refinement implies an earlier TS with an estimated bond order of 0.21 for the breaking glycoside bond. Nucleophilic 11 ACS Paragon Plus Environment

ACS Catalysis

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

participation is seen through the donation of electron density from the Tyr342 OH lone pair into the (SA C-2)…(SA O-6) antibonding NBO that has a second-order interaction energy of 5.4 kcal/mol. An oxocarbenium ion like TS is evident from the anomeric carbon that has a valence electron configuration approximating an idealized sp2 hybridization, 2s0.942p2.39. Further supporting evidence for an oxocarbenium ion comes from the increased (SA C-2)…(SA O-6) bond order (1.27 cf. reactant 0.95) and reduced charge on O-6 (-0.47 cf. reactant -0.59). The ring conformation starts at the Equator (B2,5 and 4S2) of the Crè mer−Pople sphere and travels longitudinally through E5 and 4H5 (TS) to 2C5 (the CI pucker conformer) at the pole. To achieve a flattened TS the SA C-5 is held below the plane of the pyranose ring by a hydrogen bond between its acetylamido group and Asp96, and C2 is drawn down onto the plane. Mechanistic Details of Elimination 1B The representative Elimination 1B crossing trajectory is less synchronous and shows a greater degree of fluctuation in the forming bond between Tyr342 OH base and SA H-31 (Figure 2B). The glycoside bond stretches to 2.0 Å early in the trajectory with the SA ring adopting a flattened E5 pucker due to dissociation of the leaving group. In the absence of nucleophilic attack, Tyr342 OH forms contact with SA H-31. There is significant fluctuation in this contact distance with a variance of 0.18 Å over the first 1 ps of the trajectory, compared to 0.03 Å for (Tyr342 O)…(SA C-2) in the displacement reactive trajectory. The positional instability of the proton hinders the system from meeting the selective stereoelectronic requirement for proton abstraction. The effect of the proton’s positional instability on the failure of the elimination reaction to progress is quantifiably evident from an examination of

12 ACS Paragon Plus Environment

Page 12 of 27

Page 13 of 27

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

ACS Catalysis

the diffuse equilibrium distribution of proton abstraction distance and angle (Figure 1). The probability of Tyr342 OH successfully approaching SA H-31, to form a NAC, is far less than Tyr342 OH approaching the SA C-2. However, when the former does happen and Tyr342 OH comes within 2 Å of SA H-31, the pair fluctuates about this configuration for a further ~0.4 ps before abstracting the proton, thereby implying a dissociative E2-type mechanism (DNAxhDH). On closer inspection, when the tyrosine oxygen approaches SA H-31, the ring puckers through a 6S2 conformation displaying a pseudo-axial glycoside bond. This approximate anti-periplanar geometry of the glycoside bond relative to H-31 is favorable for an E2 elimination reaction (AxhDHDN). An additional impediment to the successful completion of the elimination reaction is the narrowing of the free energy channel into a well-defined TS volume (Figure 1) therefore supporting a bimolecular reaction mechanism. The combination of the restricted TS volume, the positional instability of the SA H-31 and the associated ring puckering suggests that the mechanism derived from the SCC-DFTB level of theory is somewhere between DNAxhDH and AxhDHDN. The DFT TS structure (ξ2 = 1.34 Å, ξ3 = 2.44 Å) supports a DNAxhDH mechanism with an estimated bond order of only 0.08 for ξ3 (cf. displacement TS structure 0.23). Support for a DNAxhDH mechanism is provided by the DFT TS structure optimized from the representative trajectory. Animation of the single negative frequency shows major contribution from the proton transfer, but little movement along the glycoside bond stretching vibration. The electron deficiency on the anomeric carbon is quenched by donation of electron density into the (SA C2)…(SA O-6) bond, giving the structure some oxocarbenium character (O-6 charge -0.48, (SA C-2)…(SA O-6) bond order 0.97), as well as donation of electron density 13 ACS Paragon Plus Environment

ACS Catalysis

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

into the (SA C-3)…(SA C-2) bond. The bond order for the latter increases to 1.13 compared to 0.98 in the displacement TS structure. The DFT TS structure reveals that the Tyr342 OH lone pair donates into the C-3 – H-31 antibonding orbital with an interaction energy of 89.2 kcal/mol. The protonation of both Glu230 and Tyr342 residues leads to a new interaction mode in which Tyr342 forms a hydrogen bond with the SA C-2 carboxylate group. Step 2: Deglycosylation A second possible source of DANA formation in the TcTS-catalyzed transferase reaction is the deglycosylation step (Scheme 3).

Scheme 3 The primary reaction, Displacement 2A, that leads to TcTS deglycosylation is shown

alongside the competing Elimination 2B reaction to DANA.

14 ACS Paragon Plus Environment

Page 14 of 27

Page 15 of 27

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

ACS Catalysis

The FEV for the proposed competing reactions was computed and resolved as a function of ξ1 = (Gal O-3)…(SA C-2), ξ1 = (Gal O-3)…(SA H-32) and ξ3 = (SA C2)…(Tyr342 O) in order to describe the transfer of covalently bound SA to the lactose acceptor (Figure 3). The results show that the primary displacement reaction is favored by both barrier height and larger TS volume. The displacement TS is located at ξ1 = 2.1 Å and ξ3 = 2.4 Å with an associated free energy of 20.6 kcal/mol. There is a degree of asymmetry in the FEARCF free-energy barrier height and TS location for the forward and reverse displacement reactions. This may arise from the greater degree of translational freedom of β-Gal O-3 compared to the Tyr342 OH nucleophile of the reverse reaction, Displacement 1A. The elimination reaction crosses a TS barrier that is 6.8 kcal/mol higher. This isoenergy contour extends along ξ3 through a bottleneck TS volume that is smaller than the displacement reaction TS volume by a factor of 4 (Table 1). The TS volume (top left frame in Figure 3) is flat in appearance and implies that elimination proceeds at a range of leaving group distances, although the FEV grid point with the smallest gradient was found at ξ2 = [(Gal O-3)…(SA H-32)] = 1.3 Å and ξ3 = 2.3 Å. While the MC equilibrium sampling about the Displacement 1A NAC is significantly more concentrated than that about the Elimination 1B NAC (Figure 3), this difference in sampling area is not as pronounced in the second step. In Displacement 2A, the nucleophilic attack distance, ξ1 , shifts to longer bond lengths, with the probability contours centered just outside the van der Waals contact distance for carbon and oxygen, 3.77 Å. On the other hand, the postures adopted by the reacting substrates become more favorable for elimination. The ξ2 distribution narrows and the θ2 = (Gal O-3)…(SA H-32)…(SA C-3) angle becomes more linear, which enhances

15 ACS Paragon Plus Environment

ACS Catalysis

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

stereoelectronic alignment of the lone pair on β-Gal O-3 and the (SA C-3)…(SA H32) antibonding orbital.

Figure 3 FEV for deglycosylation step with ξ illustrated in the molecular drawing. Isosurfaces are

shown at 22.1 kcal/mol (green) and 28.9 kcal/mol (red), and corresponding Displacement 2A (blue) and Elimination 2B (yellow) FEPs are traced. SCC-DFTB free energy TSs are indicated by black dots, while the M06-2x/6-31G(d) PE TS structures are shown in gray. TS volumes are shown flanking the FEV. Probability contour plots for the geometrical analysis of the 500 ps MC equilibrium trajectory are shown below at confidence intervals of 30%.

Mechanistic Details of Displacement 2A Representative crossing trajectories for the deglycosylation step show that the dominant Displacement 2A reaction is more associative (Figure 4A) than competing Elimination 2B (Figure 4B). In the former reaction, the C-O bond to Tyr342 fluctuates between 1.5 and 2.0 Å with the SA ring moving from 2C5 to adopt a 4H 5

pucker. The SA-Tyr342 bond then starts breaking at 3.6 to 4.0 ps triggering

16 ACS Paragon Plus Environment

Page 16 of 27

Page 17 of 27

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

ACS Catalysis

Figure 4 Mechanistic analysis of representative (A) Displacement 2A and (B) Elimination 2B

crossings. Geometric analysis of the breaking and forming bonds comprising ξ is plotted on the left along with pucker. The forming bond is colored to match the ξ definition in Figure 3, and grayed sections illustrate the TS region from which the TS structure was optimized. The NBOs for this structure are shown on the right at an electron density of -0.02 and 0.02.

nucleophilic attack by the β-Gal O-3 nucleophile and passage through a planar TS (E5 pucker). This profile and the location of the free energy TS at ξ1 = 2.1 Å and ξ3 = 2.4 Å is in line with a dissociative SN2-type mechanism (DNAN). Analyzing the electronic nature of the DFT TS structure, ξ1 = 2.08 Å and ξ3 = 2.57 Å, reveals that the β-Gal O-3 lone pair donates into the (SA C-2)…(SA O-6) antibonding NBO, localized on C-2, with a stabilizing energy of 50.0 kcal/mol. Since the TS structure is

17 ACS Paragon Plus Environment

ACS Catalysis

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

the same as Displacement 1A but approached from the opposite end, the oxocarbenium chemical character is the same as described above i.e., SA C-2 electron configuration 2s0.952p2.39, SA O-6 charge -0.47 and the (SA C-2)…(SA O-6) estimated bond order 1.27. Mechanistic Details of Elimination 2B On the other hand, the elimination reaction comprises a dissociating SA-Tyr342 bond fluctuating around 2.0 Å for significant periods while the SA ring adopts a 4H5 pucker. Here the lactose base approaches 1.5 Å from SA H-32 but does not immediately abstract the proton. The (Gal O-3)…(SA H-32) bond shows a variance of 0.12 Å over the first 1 ps of the shown trajectory - cf. 0.04 Å for (Gal O-3)…(SA C2) in the displacement trajectory. Elimination eventually proceeds through 4H5 that has a favorable anti-periplanar geometry. The M06-2x/6-31G(d) TS structure (ξ2 = 1.31 Å, ξ3 = 2.34 Å), shows overlap of the β-Gal O-3 lone pair with the (SA C-3)…(SA H-32) antibonding NBO with an interaction energy of 105.3 kcal/mol. Electron deficiency on C-2 is quenched by increased density in the (SA C-2)…(SA O-6) and (SA C-2)…(SA C-3) bonds giving a chemical profile for the SA ring very similar to the elimination TS in the first step: (SA C-2)…(SA C-3) estimated bond order 1.30, SA O-6 charge -0.47 and (SA C2)…(SA O-6) estimated bond order 1.20. The free energy TS, the DFT TS structure, and the positional instability of the SA H-32 all point to a DNAxhDH mechanism. Concluding Remarks In closing, multidimensional, computations that included the distance between the nucleophile (of the primary displacement reaction) and SA C-3 hydrogen in the ξ set for both steps allowed us to simultaneously monitor the elimination and 18 ACS Paragon Plus Environment

Page 18 of 27

Page 19 of 27

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

ACS Catalysis

displacement pathways. This approach has the advantage of directly comparing the fundamental thermodynamics defining the elimination side reactions producing DANA with that of the displacement reactions. While the elimination reaction paths require crossing comparatively higher energy barriers (exceeded by ΔGTS ∼7 kcal/mol) the importance of controlling biased conformational and configurational sampling is a key observation that assists in our understanding of enzyme catalytic mechanisms. Deconvolution of the statistics underpinning the free energy for the equilibrium states provide the spatial-temporal details of the TcTS catalyzed transformation. In both the glycosylation and deglycosylation steps the probability of the nucleophile (alternatively Tyr342 OH or β-Gal O-3) successfully approaching SA C-2 to form a NAC, through which the ground state must pass to become the TS, is significantly greater than the same group approaching the SA C-3 hydrogen to form the elimination NAC. The nucleophile is held in a position with favorable stereoelectronic alignment for attack on the nucleophilic site, while the positional instability of the SA C-3 proton hinders the system from meeting the selective stereoelectronic requirement for proton abstraction.

Methods Here we use multidimensional reaction dynamics to map out the Ping-Pong reaction as well as the competing reactions leading to DANA formation in TcTS (Scheme 1). In reaction modelling computations the free energy is plotted as a function of the order parameter (i.e., reaction coordinate, ξ), where this parameter is a measure of reaction progress that includes atomic motions fundamental to its success. It is intuitive to compose ξ from a number of variables drawn from an internal coordinate system, such as bonds or angles. Typically, the free energy is

19 ACS Paragon Plus Environment

ACS Catalysis

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

sampled along a 1-D (reaction path) or 2-D (reaction surface) ξ. However, the simplified picture derived from one or two-dimensional reaction dynamics models results in the loss of physically and chemically important information.18 To eliminate these shortcomings that prevent the simultaneous monitoring of multiple and diverging reaction paths possible in an enzyme we use 4-D free energy volumes (FEV) to explore the TcTS-catalyzed trans-sialidase activity on SL. Protein Preparation The TcTS used had been crystallized, at a resolution of 1.6 Å, with SL in the active site by mutation of the putative catalytic acid/base residue Asp59 to Ala (PDB accession code 1S0I)19. This structure was prepared and equilibrated under periodic boundary conditions using a molecular mechanics force field as previously described.20 The equilibrated TcTS systems derived from the experimental MC crystal structure were then used to commence all computational reaction dynamics simulations. FEVs Following the previously published protocol, the system was further equilibrated for 500 ps using SCCDFTB/MM molecular dynamics as implemented in CHARMM21 (see SI for details), and used as the starting point for simulating the glycosylation reaction. Here we use the free energies from adaptive reaction coordinate forces (FEARCF) method, which we have previously used and described in detail.18, 22 The method employs a flat histogram strategy and applies forces to atoms comprising ξ in order to drive the system to equally visit each state defined by ξ. The driving forces are obtained directly from the gradient of an iteratively improved estimate of the PMF. In each iteration i (comprising an ensemble of

20 ACS Paragon Plus Environment

Page 20 of 27

Page 21 of 27

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

ACS Catalysis

simulations) driving forces derived from the PMF gained from iterations 1 through i-1, were used in every (30 ps) simulation. The ξ-space for the first and second steps of TcTS-catalyzed hydrolysis is defined by three bonds: SA C-2 – Nu, SA H3 – B and SA C-2 – LG. The addition of H-3 abstraction among the geometric variables allows us to follow the free energy of reaction along the elimination reaction path, and to independently assess the reactions leading to product (CI or SL) and side product (DANA) in each step. Protonation of the glycosidic linkage and proton transfer between Tyr342 and Glu230 is possible using only thermal energy and so requires no biasing force to drive this reaction coordinate. Consequently, these bonds were excluded from the reaction coordinate definitions. The PMF dimensions are 0.5-6.5 Å × 0.5-6.5 Å × 0.5-6.5 Å and all elements of the 3-D ξs were partitioned into discrete bins comprising a spacing of 0.1 Å. The individual bond and angle sampling statistics describing the postures adopted by reacting substrates are conflated within the 3-D ξs that compose the FEV. Unpacking the mechanistic details required the deconvolution of the population distributions of selected internal coordinates from the equilibrium distributions that are derived from representative reactive trajectories. Consequently, the geometric relationship between Tyr342 and SA was analyzed from the data taken from a 500 ps SCC-DFTB/CHARMM equilibrium trajectory. The Transition State (TS) is derived from points on the FEV and is defined as the point that has the smallest numerical gradient. The minimum pathway is then computed by following the steepest gradient down to the stationary points. TS volumes are calculated by a voronoi tessellation of free energy contour about the TS (see SI for details). 21 ACS Paragon Plus Environment

ACS Catalysis

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

Reaction trajectory analysis Representative crossing trajectories were selected from the first crossing trajectories and then compared with the MEP gained from the gradient of the FEV. Discrete Frè chet distances were calculated between the trajectories and MEP path. The trajectory with the smallest distance was selected whereupon SA ring puckers were calculated for each step using in-house code. In addition, TS structures were extracted from the representative FEARCF trajectories and optimized using SCC-DFTB/CHARMM with POLYRATE23 or M062x/6-31G(d)/CHARMM in NWChem 6.5.24 This is done by relaxing a structure taken from a dynamic snapshot onto the PES while constraining the atoms comprising the reaction coordinate, before following the single negative frequency to the TS structure (see SI for details). Normal modes were computed for the converged structures to confirm the existence of a single negative frequency corresponding to reaction progress. A Natural bond orbital (NBO) analysis was run on the DFT structures using NBO as implemented in Gaussian 0925 including all background charges. Bond orders reported are the Wiberg bond index. Primary 13C and β- 2H KIEs were also calculated for Displacement 1A (Scheme 2) for comparison against experimentally calculated values. Ratios were calculated from the numerical Hessians of the optimized TS and MC structures using the BigeleisenMayer formalism.26

AUTHOR INFORMATION Corresponding Author * Kevin J. Naidoo (e-mail): [email protected] Author Contributions 22 ACS Paragon Plus Environment

Page 22 of 27

Page 23 of 27

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

ACS Catalysis

All authors have given approval to the final version of the manuscript. Funding Sources This work is based in part upon research supported by the South African Research Chairs Initiative (SARChI) of the Department of Science and Technology (DST) and National Research Foundation (NRF) grant 48103 (KJN).

ACKNOWLEDGMENT IR thanks SARChI for doctoral fellowship support. We thank the Centre for HighPerformance Computing (CHPC) for use of the compute clusters.

ABBREVIATIONS CI covalent intermediate, SA sialic acid, MC Michaelis complex, TS transition state, PES potential energy surface.

ASSOCIATED CONTENT Supporting Information Available: Details of QM/MM simulations as well as QM optimizations. This material is available free of charge via the Internet at http://pubs.acs.org.

References 1.

Scudder, P.; Doom, J. P.; Chuenkova, M.; Manger, I. D.; Pereira, M. E. A., J. Biol. Chem. 1993, 268, 9886-9891.

2.

Todeschini, A. R.; Mendonça-Previato, L.; Previato, J. O.; Varki, A.; van Halbeek, H., Glycobiology 2000, 10, 213-221.

3.

Burmeister, W. P.; Henrissat, B.; Bosso, C.; Cusack, S.; Ruigrok, R. W. H., Structure 1993, 1, 19-26.

23 ACS Paragon Plus Environment

ACS Catalysis

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

4.

Moustafa, I.; Connaris, H.; Taylor, M.; Zaitsev, V.; Wilson, J. C.; Kiefel, M. J.; von Itzstein, M.; Taylor, G., J. Biol. Chem. 2004, 279, 40819-40826.

5.

Xu, G.; Kiefel, M. J.; Wilson, J. C.; Andrew, P. W.; Oggioni, M. R.; Taylor, G. L., J. Am. Chem. Soc. 2011, 133, 1718-1721.

6.

Jongkees, S. A. K.; Withers, S. G., Acc. Chem. Res. 2014, 47, 226-235.

7.

Ress, D.; Linhardt, R., Curr. Org. Synth. 2004, 1, 31-46.

8.

(a) Schenkman, S.; Jiang, M. S.; Hart, G. W.; Nussenzweig, V., Cell 1991, 65, 1117−1125; (b) Paris, G.; Cremona, M. L.; Amaya, M. F.; Buschiazzo, A.; Giambiagi, S.; Frasch, A. C.; Alzari, P. M., Glycobiology 2001, 11, 305-311; (c) Freire-de-Lima, L.; Oliveira, I. A.; Neves, J. L.; Penha, L. L.; Alisson-Silva, F.; Dias, W. B.; Todeschini, A. R., Front. Immunol. 2012, 3, 356.

9.

Todeschini, A. R.; Dias, W. B.; Girard, M. F.; Wieruszeski, J.; MendonçaPreviato, L.; Previato, J. O., J. Biol. Chem. 2004, 279, 5323-5328.

10.

Oliveira, I. A.; Gonçalves, A. S.; Neves, J. L.; von Itzstein, M.; Todeschini, A. R., J. Biol. Chem. 2014, 289, 423-436.

11.

Buschiazzo, A.; Amaya, M. F.; Cremona, M. L.; Frasch, A. C.; Alzari, P. M., Mol. Cell 2002, 10, 757-68.

12.

Damager, I.; Buchini, S.; Amaya, M. F.; Buschiazzo, A.; Alzari, P.; Frasch, A. C.; Watts, A.; Withers, S. G., Biochemistry 2008, 47, 3507-12.

13.

Bennet, A. J., Curr. Opin. Chem. Biol. 2012, 16, 472-8.

14.

Bueren-Calabuig, J. A.; Pierdominici-Sottile, G.; Roitberg, A. E., J. Phys. Chem. B 2014, 118, 5807-16.

15.

Demir, O.; Roitberg, A. E., Biochemistry 2009, 48, 3398-406.

16.

Anslyn, E. V.; Dougherty, D. A., Modern Physical Organic Chemistry. University Science Books: Sausalito, California, 2005.

24 ACS Paragon Plus Environment

Page 24 of 27

Page 25 of 27

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

ACS Catalysis

17.

Yang, J.; Schenkman, S.; Horenstein, B. a., Biochemistry 2000, 39, 5902-10.

18.

Naidoo, K. J., Phys. Chem. Chem. Phys. 2012, 14, 9026-9036.

19.

Amaya, M. F.; Watts, A. G.; Damager, I.; Wehenkel, A.; Nguyen, T.; Buschiazzo, A.; Paris, G.; Frasch, A. C.; Withers, S. G.; Alzari, P. M., Structure 2004, 12, 775-784.

20.

Rogers, I. L.; Naidoo, K. J., J. Phys. Chem. B 2015, 119, 1192-1201.

21.

(a) Cui, Q.; Elstner, M.; Kaxiras, E.; Frauenheim, T.; Karplus, M., J. Phys. Chem. B 2001, 105, 569-585; (b) Brooks, B. R.; Bruccoleri, R. E.; Olafson, B. D.; States, D. J.; Swaminathan, S.; Karplus, M., J. Comput. Chem. 1983, 4, 187217; (c) Brooks, B. R.; Brooks, C. L.; Mackerell, A. D.; Nilsson, L.; Petrella, R. J.; Roux, B.; Won, Y.; Archontis, G.; Bartels, C.; Boresch, S.; Caflisch, A.; Caves, L.; Cui, Q.; Dinner, A. R.; Feig, M.; Fischer, S.; Gao, J.; Hodoscek, M.; Im, W.; Kuczera, K.; Lazaridis, T.; Ma, J.; Ovchinnikov, V.; Paci, E.; Pastor, R. W.; Post, C. B.; Pu, J. Z.; Schaefer, M.; Tidor, B.; Venable, R. M.; Woodcock, H. L.; Wu, X.; Yang, W.; York, D. M.; Karplus, M., J. Comput. Chem. 2009, 30, 1545-1614.

22.

Naidoo, K. J., Sci. China Chem. 2011, 54, 1962-1973.

23.

J. C. Corchado, Y.-Y. C., P. L. Fast, J. Villa, W.-P. Hu, Y.-P. Liu, G. C. Lynch, K. A. Nguyen, C. F. Jackels, V. S. Melissas, B.J. Lynch, I. Rossi, E. L. Coitino, A. Fernandez-Ramos, J. Pu, T. V. Albu, R. Steckler, B. C. Garrett, A. D. Isaacson, and D. G. Truhlar POLYRATE, Version 9.0; University of Minnesota: Minneapolis.

24.

Aprà, E.; Bylaska, E. J.; Dean, D. J.; Fortunelli, A.; Gao, F.; Krstić, P. S.; Wells, J. C.; Windus, T. L., Comput. Mater. Sci. 2003, 28, 209-221.

25 ACS Paragon Plus Environment

ACS Catalysis

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

25.

(a) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.; Nakatsuji, H.; Caricato, M.; Li, X.; Hratchian, H. P.; Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Montgomery Jr., J. A.; Peralta, J. E.; Ogliaro, F.; Bearpark, M. J.; Heyd, J.; Brothers, E. N.; Kudin, K. N.; Staroverov, V. N.; Kobayashi, R.; Normand, J.; Raghavachari, K.; Rendell, A. P.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Rega, N.; Millam, N. J.; Klene, M.; Knox, J. E.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Martin, R. L.; Morokuma, K.; Zakrzewski, V. G.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Dapprich, S.; Daniels, A. D.; Farkas, Ö.; Foresman, J. B.; Ortiz, J. V.; Cioslowski, J.; Fox, D. J. Gaussian 09, Gaussian, Inc.: Wallingford, CT, USA, 2009; (b) Glendening, E. D.; Reed, A. E.; Carpenter, J. E.; Weinhold, F. NBO version 3.1.

26.

Bigeleisen, J.; Mayer, M. G., J. Chem. Phys. 1947, 15, 261-267.

26 ACS Paragon Plus Environment

Page 26 of 27

Page 27 of 27

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

ACS Catalysis

TOC

27 ACS Paragon Plus Environment