Jiahui
Zhou‡
a,
Tao
Han‡
b,
Shahbaz
Ahmad
acd,
Derek
Quinn
cd,
Thomas S.
Moody
cd,
Qi
Wu
*b and
Meilan
Huang
*a
aSchool of Chemistry and Chemical Engineering, Queen's University, David Keir Building, Stranmillis Road, Belfast BT9 5AG, Northern Ireland, UK. E-mail: m.huang@qub.ac.uk
bDepartment of Chemistry, Zhejiang University, Hangzhou, 310027, P. R. China. E-mail: wuqi1000@163.com
cAlmac Sciences, Department of Biocatalysis and Isotope Chemistry, Almac House, 20 Seagoe Industrial Estate, Craigavon BT63 5QD, Northern Ireland, UK
dArran Chemical Company Limited, Unit 1 Monksland Industrial Estate, Athlone, Co. Roscommon, Ireland
First published on 6th November 2023
Alcohol dehydrogenases (ADH) are a family of enzymes that catalyse the interconversion between ketones/aldehydes and alcohols in the presence of NADPH cofactor. It is challenging to desymmetrise the substituted cyclopentane-1,3-dione by engineering an ADH, while the reaction mechanism of the metal independent ADH remains elusive. Here we measured the conversion of a model substrate 2-benzyl-2-methylcyclopentane-1,3-dione by LbADH and found it predominately gave the (2R,3R) product. Binding mode analysis of the substrate in LbADH from molecular dynamics simulations disclosed the origin of the enantioselectivity of the enzyme; the opening and closing of the loop 191–205 above the substrate are responsible for shaping the binding pocket to orientate the substrate, so as to give different stereoisomer products. Using QM/MM calculations, we elucidated the reaction mechanism of LbADH. Furthermore, we demonstrated the reaction profile corresponding to the production of different stereoisomers, which is in accordance with our experimental observations. This research here will shed a light on the rational engineering of ADH to achieve stereodivergent stereoisomer products.
There are two types of ADHs, including Zinc containing and non-metal ADHs. The mechanism of zinc-ADH catalysed reduction mechanism was reported previously,6,7 however, to the best of our knowledge, the mechanism of the reduction of carbonyl compounds with metal-independent ADHs remains elusive. In this research, we expressed an ADH enzyme from Lactobacillus brevis (LbADH) and measured its conversion of the substrate. 2-Benzyl-2-methylcyclopentane-1,3-dione was chosen as a model substrate because the reduction would give four different enantiodivergent products and hence enable the desymmetrisation of the substrate. Our kinetic experiments showed that the (2R,3R) and (2S,3S) products were observed with a ratio of 9 to 1, while the other two chiral products (2S,3R) or (2R,3S) were not observed.
MD simulations disclosed that the loop 191–205 above the substrate could exhibit either open or closed conformation, which is responsible for the production of different stereoisomer products. Further hybrid quantum mechanics/molecular mechanics (QM/MM) calculations disclosed the mechanism of the reduction catalysed by LbADH. We elucidated that the reaction starts with the rate-limiting step, i.e., the concerted hydride transfer from NADPH to the substrate carbonyl carbon and proton transfer from Tyr155 to the substrate carbonyl oxygen. The reaction then proceeds with a concerted proton relay; deprotonated Tyr155 phenolate abstracts a proton from NADP+ ribose, which in turn abstracts a proton from the nearby catalytic acid Lys159. The second step is barrierless for the major product (2R,3R), but it needs to overcome a notable barrier for the minor product (2S,3S) and an even higher barrier for the inaccessible product (2S,3R) or (2R,3S), implicating the origin of the enantioselectivity of the enzyme. Additionally, we found the reaction coordinate in the reactant (i.e., the distance between the NADPH hydride to the carbonyl carbon) is correlated to the barrier for the rate-limiting step.
The enzyme-substrate complexes were solvated in a pre-equilibrated cuboid box of TIP3P17 water molecules, and any protein atom was at least 10 Å to from the edge of the cuboid box. The system was then neutralised by adding 8 Na+ counterions using the tLEaP program implemented in AMBER 20. A harmonic restraint force constant of 100 kcal−1 Å−2 was applied to the solute molecules and ions to minimise the solvent molecules, followed by 1000-step steepest descend and 1000-step conjugate gradient unrestrained minimisation. A cut-off of 10 Å was applied for non-bonded Lennard-Jones potential and electrostatic interactions. Hydrogen bonds were constrained using the SHAKE algorithm during all MD simulations. Under constant volume and periodic boundary conditions, a progressive heating was performed from 0 K to 300 K in 100 ps (5000 steps with a step size of 0.02 ps), followed by 1 ns equilibration using NPT ensemble at 300 K. A harmonic restraint of 5 kcal mol−1 was applied to the solute at both equilibration stages. After the equilibration stages, a 200 ns production MD simulation was run using an NPT ensemble at 300 K and 1 bar. Three replicas of MD simulations were run for each system.
The substrate, truncated NADPH, Tyr155, Lys159, and Ser142 are included in the QM region. During the geometry optimization, residues within 6 Å of the QM region were allowed to move freely, while the remaining residues were kept frozen (Fig. S1, ESI†). The hydrogen link atoms were used to saturate the dangling bond at the QM/MM boundary. The QM/MM calculations were performed using the electronic embedding model within the Gaussian13 ONIOM. Geometries of all intermediates and transition states were optimized at the B3LYP/6-31G(d) level. Frequency calculations within the harmonic approximation were used to verify the nature of all intermediates and transition states. Free energy and enthalpic corrections were carried out by computing harmonic frequencies analytically at 298.15 K. For each step on the reaction profile, thermochemical correction terms δEG were carried out as a difference of the reaction energy (ΔEB3LYP/6-31G(d)) and the corresponding free energy (ΔGB3LYP/6-31G(d)):
δEG = ΔGB3LYP/6-31G(d) − ΔEB3LYP/6-31G(d) |
Potential energy surface scans were conducted to locate the transition states at the same level, i.e., B3LYP-D3(BJ)/6-31G(d). Transition states were confirmed through visual inspection of the imaginary frequency modes as well as intrinsic reaction coordinate (IRC) calculations.
The corrected free energies (ΔG) were calculated as follows:
ΔG = ΔESP + ΔEG |
The single point energy of the optimized intermediates and transition states were also compared using the large basis set, i.e., 6-311+G(2d,2p) and def-2TZVPP.
The stereoisomeric distribution of the products for WT-LbADH was measured as following. Sodium phosphate buffer (1 ml, 100 mM, pH 7.0) containing 10% DMSO (v/v), 10 mM substrate, 30 mM glucose, 0.5 mM NADP+, 0.5–2 mg of LbADH and 4 U GDH were shaken at 30 °C for overnight. Then the reaction was stopped by the addition of an equal amount of ethyl acetate. The organic phase was separated and the solvent was removed; the resulting sample was analysed by chiral HPLC to determine the enantiomeric excess value of the alcohol products.
The reduction of the substrate yielded 90% and 10% stereoisomeric (2R,3R) and (2S,3S) products, respectively, whereas no (2S,3R) nor (2R,3S) product was obtained (Fig. 1 and Scheme 1).
In the proximity of catalytic sites of all the structures, Ser142 stabilizes the substrate carbonyl group by a hydrogen bond (Fig. 2 and Fig. S3, ESI†); Tyr155 is positioned toward the substrate carbonyl to serve as an H-bond donor and also forms H-bond with the NADPH ribose hydroxyl group, which is, in turn, H-bonded to Lys159.
The binding of the substrate in the enzyme is mediated by Glu144, which results in different substrate orientations. In the substrate pose that leads to the main (2R,3R) product, Glu144 forms a hydrogen bond with Tyr189 (Fig. 2A and Fig. S3, ESI†), which is stabilized by the methyl group of Met205. The benzene ring of the substrate is nested between the surrounding Ala93 and Leu194, forming hydrophobic interactions with them. Thr192 forms H-bond with the amide NH2 of the NADPH nicotinamide. The loop 191–205 adjacent to Tyr189 displays a closed conformation, with Asp196 interacting with Lys191 by ionic interactions, which in turn tethered by Glu202 (Fig. 2).
In the substrate pose that leads to the minor (2S,3S) product (Fig. 2B), the sidechain of Glu144 is turned away from the catalytic pocket. The loss of the H-bond between Glu144 and Tyr189 caused the loop K191-M205 to display a significant conformational change to adopt an open conformation. The opening of the loop makes Leu194 become far away from the substrate, such that the substrate was able to move freely in the binding pocket and eventually stabilized by the favourable hydrophobic interactions with the surrounding Leu152 and Ala193 with its benzene ring. The opening of the loop also caused the loss of the H-bond between Thr192 and NADPH. Meanwhile, Asp196 on the loop becomes exposed and its interaction with Lys191 is lost, leaving the latter interacting only with Glu202. A water flux moved in the catalytic site with the opening of the loop (Fig. 2B).
Thus we demonstrated that the loop191–205 region is largely stabilized in the substrate pose leading to the dominant (2R,3R) product (Fig. S2B, ESI†), compared to other substrate poses that lead to minor product (2S,3S) or do not give the corresponding chiral product such as (2S,3R).
LbADH is homologous to RasADH3 and there is 5 AA difference between the two homologous enzymes. Previously, Chen et al. studied RasADH and suggested that Ser137, Tyr150, and Lys154 (corresponding to Ser142, Tyr155, Lys159 in LbADH) may participate in the catalytic process, along with the nicotinamide ring of NADPH.3 However, the exact catalytic process of ADH is not known.
From our MD simulations, Tyr155 and Lys159 are located in the proximity of the substrate and NADPH; therefore, these two residues were included in the QM region, along with the substrate, the NADPH cofactor. Tyr155 and Lys159 were truncated to keep the sidechain and NADPH was truncated to keep the nicotinamide riboside part in the QM/MM calculations.
Potential energy surface scans were conducted for the enzyme-substrate complex structures that would lead to different stereoisomers, i.e. the most dominant (2R,3R) (90% distribution among all products), the less dominant (2S,3S) (10% distribution), and also a chiral product (2S,3R) that was not observed from the kinetic experiments.
Benchmarking calculations were conducted for the DFT methods, namely B3LYP (6-31G*) and M06-2X (6-31G*) (Fig. S4, ESI†). For the substrate pose that lead to the dominant (2R,3R) stereoisomer, after the carbonyl group of the 1,3-dione substrate is reduced and Tyr155 is deprotonated, the deprotonated Tyr155 in I1 abstracts a proton from the nicotinamide riboside hydroxyl, which spontaneously abstract a proton from Lys159 (barrierless). Interestingly, for the substrate pose that leads to the minor product (2S,3S) or the unachievable product (2S,3R), the deprotonated Tyr155 in I1 needs to overcome a notable barrier to be reprotonated (Fig. S4b and S5, ESI†).
The dispersion in the DFT method may affect the energy barrier.21–23 For example, QM/MM calculations for cytochrome P450 catalysed reactions showed that dispersion correction may reduce the barrier of hydrogen abstraction significantly by around 5 kcal mol−1.21 To examine if dispersion is necessary to consider the polarisation effect for ADH, we conducted bench marking calculations with dispersion effect using B3YLP-D3(BJ) and M06-2X-D3 with the 6/31G* basis set.
Since M06-2X implicitly includes dispersion energy, as expected, including Grimme's correction did not cause significant change, whereas notable differences were observed with the B3LYP functional (Fig. 3 and Fig. S4, ESI†), highlighting the importance of including dispersion for B3LYP.
Energy barrier values calculated by M06-2X-D3 and ωB97X-D are similar, whereas lower barrier values were obtained by using B3LYP-D3(BJ) (Fig. 3). It is worth noting that for (2R,3R), the TS2 couldn’t be located using B3LYP or M06-2X/M06-2X-D3 (Fig. S4a, ESI† and Fig. 3a), thus combining B3LYP with dispersion is recommended for geometry optimizations. The two transition states TS1 and TS2 were subsequently verified by frequency calculations and IRC analysis (Fig. S6, ESI†). Further, energy corrections were conducted with two larger basis sets for the optimized geometries, namely, Def-2TZVPP and 6-311+G(2d,2p), and consistent energies were obtained (Fig. 4). For the following discussion, the results obtained by B3LYP-D3(BJ)/6-31G* with ZPE correction at the same level and single point energy corrections at Def-2TZVPP basis set (i.e. B3LYP-D3(BJ)/Def-2TZVPP//B3LYP-D3(BJ)/6-31G*) were employed.
The kinetic study of LbADH showed a KM value of 3.18 (mM) and a kcat value of 0.0114 s−1, which corresponds to a reaction barrier of 20.10 kcal mol−1 through the calculation of the Eyring equation. Our calculations show that the rate determining step of the ADH-catalysed reduction is the concerted hydride transfer from 4CH2 of NADPH nicotinamide to the substrate and proton transfer from Tyr155 to the substrate carbonyl. The formation of the major product (2R,3R) proceeds through a transition state TS1 with a barrier of 13.5 kcal mol−1 (Fig. 4). Then the reaction proceeds with a barrierless and concerted proton relay composed of the proton abstraction from NADPH ribose to the phenolate Tyr155 and proton transfer from Lys159 to the sugar. The bond-breaking distance between the attacking hydrogen and C4 increases to 1.36 Å in the transition state TS1, whereas the bond-formation distance between the C4H and the substrate carbon decreases to 1.44 Å in TS1 (Fig. 5).
The formation of the minor product (2S,3S) also goes through two transition states with similar barriers of 8.7 kcal mol−1 and 8.8 kcal mol−1, respectively (Fig. 4). The difference in the product distribution may be attributed to the different reaction pathways; the formation of the (2S,3S) requires surpassing a second barrier involving the reprotonation of Tyr155 facilitated by NADP+ and Lys159; hence the process is less energetically favourable compared to the formation of the major (2R,3R) product. We also ran a potential energy surface scan for the substrate pose that would lead to the (2S,3R) product, which was not obtained from the kinetic experiment. Our calculations show the production of the (2S,3R) product needs to go through a much high energy barrier of 24.2 kcal mol−1 for the rate limiting step and a second barrier of 2.6 kcal mol−1 (Fig. S7, ESI†), indicating the formation of such product is energetically prohibited and thus not liable, which is in agreement with our experimental observations.
Interestingly, during the formation of the main product (2R,3R), we observed that I1 transformed into a slightly lower-energy intermediate I1′. Notably, the catalytic acid Lys159 in I1′ became closer to the nicotinamide riboside hydroxyl of NADP+ than I1 (1.53Å vs. 2.12 Å), and also the bond angle is more favourable for proton transfer (165° vs. 110°), making the deprotonation from Lys159 to the NADPH ribose and the reprotonation of the Tyr155 easier to occur (Fig. 5a). The bond angle is 144° (the distance is 1.83Å) in the minor product (2S,3S) (Fig. 5b) and 126° (the distance is 2.43 Å) in the inaccessible product (2S,3R) (Fig. S8, ESI†). This indicates that the bond distance and bond angle involved in the H-bond between Lys159 and NADPH ribose is positively related to the reaction barrier of the second reaction step the proton reply.
![]() | ||
Fig. 6 Energy barrier for the rate-limiting step is linear correlated with the reaction coordinate (the distance between the C4 methylene hydrogen of NADPH and the substrate carbonyl carbon atom). |
We studied the binding of a model substrate 2-benzyl-2-methylcyclopentane-1,3-dione in an ADH enzyme represented by LbADH by molecular docking and MD simulations and demonstrated the orientation of the substrate is regulated by the opening/closing of the loop 191–205, which would lead to the different stereoisomer products. The calculations are in agreement with our kinetic studies.
We further elucidated the reaction mechanism of LbADH using by QM/MM calculations. In the first reaction step, the hydrogen of NADPH C4 methylene attacks the carbonyl carbon of the substrate, which is accompanied by a concerted proton transfer from the hydroxyl group of Tyr155. The second reaction step is also a concerted process, where the proton of the ribose hydroxyl of NADP+ is relayed between Lys159 and the phenolate Tyr155, giving rise to neutral Tyr155 and neutral Lys159. A correlation was observed between the reaction coordinate and the reaction barrier of the rate-determining step, indicating the reaction coordinate can be used as an important but easily achievable feature for large-scan screening of the enantioselective ADH variants in enzyme engineering. The understanding of preferred substrate poses in the ADH enzyme leading to different stereoisomers would provide structural basis for further rationally engineering of the enzyme to acquire enantiodivergent products.
Footnotes |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d3cp04019d |
‡ Equal contributions. |
This journal is © the Owner Societies 2023 |