Kara E.
Ranaghan
ab,
John E.
Hung
c,
Gail J.
Bartlett
b,
Tiddo J.
Mooibroek
b,
Jeremy N.
Harvey
ab,
Derek N.
Woolfson
bd,
Wilfred A.
van der Donk
*c and
Adrian J.
Mulholland
*ab
aCentre for Computational Chemistry, School of Chemistry, University of Bristol, Bristol, BS8 1TS, UK
bSchool of Chemistry, University of Bristol, Bristol, BS8 1TS, UK. E-mail: Adrian.Mulholland@bristol.ac.uk
cDepartment of Chemistry and The Howard Hughes Medical Institute, University of Illinois at Urbana-Champaign, Urbana, Illinois 61801, USA. E-mail: vddonk@illinois.edu
dSchool of Biochemistry, Medical Sciences, University of Bristol, Bristol, BS8 1TD, UK
First published on 18th February 2014
Combined quantum mechanics/molecular mechanics (QM/MM) simulations of the reaction catalysed by phosphite dehydrogenase (PTDH) identify Met53 as important for catalysis. This catalytic role is verified by experiments (including replacement by norleucine and selenomethionine), which show that mutation of this residue significantly affects kcat, without changing KM for phosphite. QM/MM and ab initio QM calculations show that the catalytic effect is electrostatic in nature. The side chain of Met53 specifically stabilizes the transition state for the hydride transfer step of the reaction catalysed by PTDH, forming a ‘face-on’ interaction with His292. To our knowledge, a defined catalytic role for methionine in an enzyme (as opposed to a steric or binding effect, or interaction with a metal ion) has not previously been identified. Analyses of the Protein Data Bank and Cambridge Structural Database indicate that this type of interaction may be relatively widespread, with implications for enzyme-catalysed reaction mechanisms and protein structure.
Fig. 1 An associative mechanism, as modelled here, for the nucleophilic displacement reaction catalysed by PTDH, with His292 as the base.1 R = reactant state, Int = pentavalent intermediate, P = product. TS1 and TS2 are the transition states for the first and second steps. |
A crystal structure of a thermostable mutant PTDH (TS-PTDH) has been solved,11 which supports the proposed functions of several residues and also revealed another residue in the active site, Met53. This Met is absolutely conserved in PTDHs from different organisms, suggesting an important role (Fig. S1 in the ESI†). In a co-crystal structure of TS-PTDH with sulfite, a competitive inhibitor for phosphite,1 the sidechain of Met53 undergoes a large conformational change from the form with only NAD+ bound, such that the sulfur atom of Met53 forms a close contact with an oxygen of sulfite with a S–O distance of 3.1 Å (Fig. 2a). This observation is interesting, as methionine is more typically associated with hydrophobic interactions and steric effects. A search of the Catalytic Site Atlas12 reveals that methionine is present in 0.8% of enzyme active sites. Previously, it has been suggested that methionine can act as a weak hydrogen bond acceptor for backbone amide NH groups,13 and that it can form stabilizing (including non-hydrogen bond) interactions with other residues.14 Of course, it is also known as a ligand for metal ions in proteins.15 Methionine residues can be oxidized to methionine sulfoxide or methionine sulfone. However, there are very few suggestions of methionine residues playing a direct role in enzyme-catalysed reactions,16,17 and we have found none in which a clear catalytic function of a methionine sidechain has been identified. Identification and analysis of catalytic functions is one area in which the combination of mechanistic modelling and experiments is proving increasingly important in enzymology.18–25 This combined approach was central in the work here: the importance of Met53 in the PTDH reaction was initially identified from QM/MM simulations, and subsequently analysed and verified by experiments.
Fig. 2 (a) A representation of the crystal structures of thermostable TS-PTDH with the NAD+ co-factor and sulfite inhibitor11 in green and the complex without sulfite in blue. NAD+ (shown with yellow carbons) and several key active site residues are shown as sticks and their interactions with the sulfite inhibitor are indicated by dotted lines. Arg237 is important for binding of the substrate and His292 is the putative base for the reaction.7 When sulfite is present, the Met53 sulfur (Sδ) to sulfite oxygen distance is 3.1 Å. In the absence of sulphite, the side chain of Met53 adopts a different orientation. (b) A representative snapshot of the PTDH ternary complex from QM/MM MD simulations showing important interactions in the active site. The black dotted lines indicate the distances involved in the reaction coordinate. |
k cat (s−1) | Relative kcat | K M,Phosphite (μM) | K M,NAD (μM) | k cat/KM,Phosphite (M−1 s−1) | D k cat | D k cat/KM | |
---|---|---|---|---|---|---|---|
a ND: not determined. Errors in parentheses were obtained from fitting data of triplicate experiments to the Michaelis–Menten equation. For experiments varying the phosphite concentration, the NAD+ concentration was held constant at 4 mM, and for experiments in which the NAD+ concentration was varied, the phosphite concentration was held constant at 2 mM. †Dkcat is the substrate deuterium kinetic isotope effect on kcat and Dkcat/KM is the substrate deuterium kinetic isotope effect on kcat/KM,Phosphite. | |||||||
17X-PTDH | 3.1 (0.1) | 1.0 | 28 (7) | 22 (6) | 1.1 (0.3) × 105 | 2.3 (0.1) | 2.1 (0.2) |
M53N | 0.016 (0.001) | 0.005 | 17 (1) | 64 (1) | 920 (70) | ND | ND |
M53A | 0.059 (0.001) | 0.019 | 100 (10) | 17 (1) | 590 (30) | 1.9 (0.1) | 1.9 (0.1) |
Nle–PTDH | 0.11 (0.01) | 0.036 | 28 (1) | 21 (2) | 3900 (200) | 1.8 (0.1) | 1.8 (0.2) |
Nle–PTDH–M53A | 0.072 (0.003) | 0.024 | 47 (7) | 7.8 (0.9) | 1500 (200) | ND | ND |
SeMet–PTDH | 3.3 (0.1) | 1.1 | 110 (10) | ND | 3.0 (0.2) × 104 | ND | ND |
Mutagenesis experiments with proteinogenic amino acids are limited to crude replacements of Met. We therefore created two 17X-PTDH mutants with global replacement of methionine by the non-proteinogenic amino acids norleucine (Nle) and selenomethionine (SeMet). These analogues are closer in size to Met than any of the 19 other proteinogenic amino acids and have a methylene group or selenium atom in place of the sulfur atom, providing a clearer indication of electronic/electrostatic (rather than structural) effects. The protein synthesis machinery in E. coli tolerates methionine analogues,26,27 allowing for substitution of Met53 in PTDH by Nle or SeMet (see Methods and Fig. S1†). The kcat of Nle–PTDH is reduced by 25-fold, comparable to the Met53Ala mutant, with essentially identical KM values for both substrates compared to the parent 17X-PTDH (Table 1). This finding indicates that loss of an interaction with Met53 is responsible for the reduction in activity, rather than a change in the structure of the active site. The recombinant 17X-PTDH contains eight Met residues, including one in the hexa-histidine tag. Of these, only Met53 is near the active site. We therefore attribute the reduction in kcat in Nle–PTDH to replacement of Met53 with Nle. This conclusion is supported by the similarity of the kinetic parameters determined for Nle–PTDH–Met53Ala to those observed for the Met53Ala mutant (Table 1). Given the similar KM values of Nle–PTDH and 17X-PTDH, we cannot completely rule out that the observed activity with Nle–PTDH is the result of contamination with 17X-PTDH. In the absence of a good active site titration agent, it is difficult to experimentally address this possibility. However, if contamination with 17X-PTDH were responsible for the observed activity, the Nle substitution would have to be even more deleterious for activity than the observed 25-fold reduction in kcat. In contrast to Nle substitution, SeMet-substituted 17X-PTDH exhibited a kcat very similar to 17X-PTDH. As discussed below, the stabilizing contributions of selenium and sulfur in the transition state are also calculated to be similar. Altogether, these findings point to a catalytic contribution of the sulfur of Met53.
For 17X-PTDH, hydride transfer is fully rate limiting,9 as shown by KIE and pre-steady state kinetic measurements (Table 1). The KIEs for the Met53Ala mutant and Nle–PTDH were also obtained to determine if the rate-limiting step of the reaction changed with the substitution of Met53 (Table 1). In each case the KIEs, Dkcat and Dkcat/KM,Phosphite (i.e. the ratios kcat,H/kcat,D and (kcat,H/KM,H/kcat,D/KM,D)), are very similar in magnitude, as would be expected if the chemical step remained fully rate-limiting combined with a small commitment to catalysis.9 The observation that hydride transfer remains rate-limiting despite a large decrease in kcat has been reported previously for other mutants of PTDH.9 The slightly different magnitudes of the KIEs point to somewhat different transition states, which is not unexpected when mutating a residue that plays a stabilizing role in the transition state (vide infra). Solvent kinetic isotope effects (SIEs) can in some cases provide additional insight into the kinetic mechanism of a reaction that involves a proton transfer step. However, SIEs for PTDH are small (1.5) and complicated by changes in pKa values in H2O and D2O.10 Because SIEs have not been informative in these previous studies on PTDH, SIEs were not determined for the Met53 mutants.
The mutagenesis experiments on Met53 indicate that this residue is important in catalysis, rather than substrate binding. The Nle mutant has reduced activity, similar to the alanine mutant, providing good evidence that the sulfur atom in Met is important for catalysis. We cannot rule out that the different conformational preferences of the side chain of Nle compared to Met28 also play a role, e.g. affecting the ability of the Nle to take on the catalytically important conformation. Attempts to globally replace Met residues with their oxygen analogues were unsuccessful, presumably because of the known toxicity of methoxinine (O-methyl homoserine).29 On the other hand, substitution of Met53 with SeMet did not change kcat: this is consistent with a specific role of the chalcogen atoms in catalysis, and suggests that the catalytic effect is electrostatic/electronic.
Structures from the reaction simulations were analysed to identify important interactions in the enzyme. The interactions between the QM and MM regions can be split into electrostatic and van der Waals components. Here, only the electrostatic energies are discussed in detail (the van der Waals energies are small and do not change significantly during the course of the reaction; the QM/MM van der Waals interaction with Met53 remains relatively constant along the path at ∼−3 kcal mol−1). Average electrostatic and van der Waals interaction energies along the reaction coordinate are given in Table S3.† In agreement with the hydrogen bonding patterns indicated in Fig. 2b, the largest contributions to the electrostatic stabilization energy were provided by Arg237, Arg301, Glu266, Lys76 and Gly77. In the simulations, the Met53 sidechain is observed to change position between Int and TS2, moving from a position close to the phosphite to one near the His292 ring (Fig. 3a). This stabilizes the positively charged histidine, as shown by a change in the interaction energy with Met53. The latter is on average slightly destabilizing in the first half of the reaction but becomes significantly stabilizing at TS2, providing an average of −7.9 kcal mol−1 stabilization (relative to R), the largest contribution of any neutral residue.
Further decomposition of the interaction between Met53 and the QM region in TS2 showed that this stabilizing interaction is mostly due to interaction between the QM region and the sidechain of Met53, rather than its backbone (see Table S3 in ESI†). A C–H⋯O hydrogen bond is present between the methyl group of Met53 and the amide oxygen of the NAD+ co-factor (Fig. 3a). This is a weak interaction, which does not make a significant contribution to stabilization; this interaction is lost in optimization of complexes in the gas phase. Rotation of the methionine methyl group also means that this C–H⋯O interaction is not present in all structures in the enzyme. See the ESI† for further discussion of the interaction between Met53 and NAD+. These interactions were not present in structures of R or Int.
Mulliken population analysis along the pathway shows the changes in charge distribution during the reaction (see Table S4 in ESI†). As expected, the atoms of phosphite and the nucleophilic water molecule undergo significant changes in charge at all stages of the reaction. The charge redistribution is less dramatic for the atoms of the NAD+ cofactor and catalytic base His292. The most significant changes in the partial atomic charges of His292 occur in the first step as it becomes protonated [partial atomic charge Nε: −0.275e in R, −0.238e in TS1 and −0.135e in Int]. There is some redistribution of charge around the histidine ring in the second half of the reaction, but the changes are small e.g. from −0.135e in Int to −0.126e in TS2 (Nε). The presence of Met53 also very weakly polarizes the partial atomic charges of the QM atoms (see Table S4 in ESI†).
The interaction between the Met53 sulfur atom and the positively charged histidine can be described as an n → π* interaction32 because it involves, at least in part, the delocalization of the lone pairs of electrons (n) on the sulfur atom into the antibonding orbitals of the histidine (π*). Clearly, the interaction here is largely electrostatic because it can be modelled by QM/MM methods; its primarily electrostatic nature is also demonstrated by NBO analysis (see ESI†).
Met53 stabilizes the transition state for the hydride transfer step (TS2), but not via hydrogen bonding. The present simulations indicate that this interaction specifically stabilizes TS2 only. Met53 is actually slightly destabilizing in R, TS1 and Int and provides an average of only −2.0 kcal mol−1 electrostatic stabilization to P as His292 and Met53 move further apart. Experimental data show that the step involving the hydride transfer is rate limiting, and hence stabilization of TS2 by Met53 is in accord with the observed reduction in kcat when this stabilization is removed by mutagenesis. In the modelled mechanism His292 becomes positively charged earlier in the mechanism (TS1), thus the interaction with Met53 might be expected to stabilize TS1 and Int also. However, our present simulations indicate that the required conformational change does not occur until TS2.
We cannot rule out the possibility of a concerted mechanism for the reaction in PTDH, but if that were the case, Met53 could also provide similar stabilization to the transition state, which geometrically would resemble the intermediate (Int) modelled here. In a dissociative process, however, protonation of His292 would occur after the rate-limiting hydride transfer event, and hence Met53 would not be expected to stabilize the transition state that governs kcat. As such, the findings here provide some indirect evidence for an associate mechanism and thus against a dissociative mechanism in PTDH if His292 is the catalytic base. The catalytic interaction between Met53 and His292 we have identified in PTDH also provides some support for the proposal that His292 is the catalytic base. According to the pH rate profile, His292 would be protonated throughout the reaction if a different residue acted as the catalytic base. Met53 might then be expected to form this type of interaction with His292 throughout and thus would not be catalytic. However, it is possible that the experimentally observed catalytic effect could be due to a change in position of Met53 during the reaction, i.e. moving to stabilize His292 during the reaction even if His292 does not function as the base. Thus, while the results provide some indirect support for His292 functioning as the base in the reaction, other potential mechanisms were modelled for comparison. Preliminary simulations of dissociative mechanisms, or with other residues as the base, led to unstable results, possibly due to limitations of the AM1 method for treating phosphorus.30 Met53 was found to give little or no stabilization in these simulations. The results presented here with the reaction modelled as an associative process, with His292 as the base, are most consistent with the experimental results. Results at this level of QM/MM theory should not be considered definitive, and therefore it will be useful to investigate the reaction mechanism further in future.
The calculated interaction energies for these complexes at the MP2 (ref. 33)/aug-cc-pVTZ and SCS-MP2 (ref. 36)/aug-cc-VTZ levels of theory, including counterpoise correction for basis set superposition error, are given in Table 2. The strongest interaction was found for the hydrogen-bonded complex between 4-methylimidazolium and methylthioethane: −16.2 kcal mol−1 (MP2) and −14.9 kcal mol−1 (SCS-MP2). The face-on complex of the same fragments had a somewhat smaller, but still significant interaction energy: −10.4 and −9.0 kcal mol−1 at the MP2 and SCS-MP2 levels, respectively. These numbers are in quite good agreement with the average QM/MMele interaction energy of −5.0 kcal mol−1 with the sidechain of Met53 at TS2 in the QM/MM simulations, although indicating that this interaction energy is underestimated in the QM/MM simulations. It is encouraging that, despite the well-known limitations of the AM1 method, QM/MM methods at this level were able to identify this face-on interaction in good agreement with ab initio methods, in terms of both geometry and energy. It is important to point out, though, that these calculations show that the interaction with Met53 is likely to be underestimated at the AM1/CHARMM27 level of theory.
Ab initio interaction energies for methylselenoethane complexes differed by less than 1 kcal mol−1 from the values obtained for the methylthioethane complexes [−9.6 kcal mol−1 face-on complex; −15.5 kcal mol−1 hydrogen-bonded complex at the MP2/aug-cc-pVTZ level of theory]; this finding is consistent with our experimental finding that the SeMet-mutant and 17X-PTDH have similar activity. The interaction energy for the neutral model complex (Fig. 3e) was considerably smaller (−5.8 and −4.4 kcal mol−1, MP2 and SCS-MP2). The interaction energy of a neutral complex of 4-methylimidazole and methylthioethane, constrained to the geometry of the face-on complex, was even smaller (−0.57 kcal mol−1). The significantly smaller interaction energy for the neutral complex with the same geometry is consistent with this type of interaction only being observed between doubly protonated (charged) histidine and methionine. The hydrogen-bonded complex between positively charged His292 and Met53 (Fig. 3c), although stronger than the face-on interaction in the ab initio calculations, was not observed in the enzyme due to the constraints imposed by the other active site residues. The interaction energies calculated here are larger than the 2–3 kcal mol−1 contribution of Met53 to TS stabilization indicated by the experimental data, but are of the correct order of magnitude. It is important to note that the interaction energies calculated here are not directly comparable to the experimental ΔΔG‡ values for the mutations; e.g. Met53 forms other interactions in the active site, which are lost during the reaction.
For MM and QM/MM simulations, a ‘reaction region', not subject to any positional restraints, was defined as a 21 Å sphere centred on O3 of NAD+. The region between 21 and 25 Å was defined as the ‘buffer region’: protein heavy atoms in this region were restrained using force constants scaled to increase with distance from the centre of the system.50,51 Atoms further than 25 Å from the centre of the system were held fixed. A stochastic boundary approach was used,30 and water was restrained to remain within the simulation system using a spherical deformable potential of radius 25 Å.52 The model system was heated to a temperature of 300 K and equilibrated for 100 ps before 1 ns of production MD was performed. Langevin dynamics were applied for atoms in the buffer region (the list of atoms in the buffer was updated every 50 steps), using friction coefficients of 250 ps−1 for protein heavy atoms and 60 ps−1 for water oxygen atoms,53 and a 1 fs time step. Non-bonded interactions, calculated using a 13 Å cut-off, were updated every 25 steps.
The QM region was defined as phosphite, a crystallographic water molecule (Wat61) positioned for reaction, the nicotinamide ring of NAD+ and 4-methylimidazole from His292: a total of 37 QM atoms, including 2 hydrogen (HQ) link atoms54 to satisfy the QM/MM boundary (with charges of MM host groups set to zero). The overall charge of the QM system was zero, from the −1e of monoprotonated phosphite and +1e of the NAD+ cofactor.
Two 1 ns QM/MM MD simulations were performed using two different starting geometries from the MM MD simulations. Umbrella sampling MD simulations were then used to model the reaction, using protocols similar to those applied successfully previously to other enzymes.50,51 Reaction coordinates for an associative mechanism were defined as Zstep1 = d(OH2–H2T) − d(H2T–NE2) − d(OH2–P)/Å for the nucleophilic attack on phosphite and Zstep2 = d(P–H1) − d(H1–C4N)/Å for the hydride transfer step (see Fig. 1). At each point along the relevant reaction coordinate, 5 ps of equilibration was carried out, followed by 50 ps of production QM/MM MD. Step 1 was sampled between reaction coordinate values of Zstep1 = −5.4 to 0.1 Å in 0.1 Å intervals, using a force constant of 200 kcal mol−1 Å−2. Each subsequent simulation was started from the 5 ps point of the preceding simulation. The value of Zstep1 was recorded at each step of the 50 ps dynamics simulation for statistical analysis. Step 2 was modelled in a similar way, in 0.1 Å intervals between Zstep2 = −3.8 to 2.2 Å. The reaction coordinate statistics for each step were combined using the weighted histogram analysis method (WHAM)55 to give the AM1/CHARMM27 free energy profile for the reaction.
Footnote |
† Electronic supplementary information (ESI) available: Additional materials and methods, QM/MM model preparation, analysis methods for QM/MM structures, discussion of the Met–NAD+ interaction, NBO calculations, PDB and CSD searches, Tables S1–S6 and Fig. S1–S5. See DOI: 10.1039/c3sc53009d |
This journal is © The Royal Society of Chemistry 2014 |