Rui
Lai
*a and
Hui
Li
*bcd
aCollege of Chemistry, Jilin University, Changchun, 130021, China. E-mail: lairui@dicp.ac.cn
bDepartment of Chemistry, University of Nebraska-Lincoln, Lincoln, NE 68588, USA. E-mail: hli4@unl.edu
cNebraska Center for Materials and Nanoscience, University of Nebraska-Lincoln, Lincoln, NE 68588, USA
dCenter for Integrated Biomolecular Communication, University of Nebraska-Lincoln, Lincoln, NE 68588, USA
First published on 10th March 2025
Understanding the catalytic mechanism of histone deacetylases can greatly benefit the development of targeted therapies that are safe and effective. Combined quantum mechanical and molecular mechanical (QM/MM) Møller–Plesset second-order perturbation theory (MP2) geometry optimizations are performed to investigate the catalytic mechanism of the deacetylation reaction of a tetrapeptide catalyzed by human Histone Deacetylase 8. A three-step catalytic mechanism is identified: the first step is the formation of a negatively charged tetrahedral intermediate via nucleophilic addition of the activated water to the amide C atom and a proton transfer from the water to His143; the second step is the formation of a neutral tetrahedral intermediate with an elongated amide C–N bond via a proton transfer from His143 to the amide N atom. The third step is the complete cleavage of the amide C–N bond, accompanied by a proton transfer from the newly formed carboxylic group of the neutral tetrahedral intermediate to His142. These three steps have similar computed energy barriers, with the second step having the highest calculated activation free energy of 19.6 kcal mol−1. When there is no potassium ion at site 1, the calculated activation free energy is 17.7 kcal mol−1. Both values are in good agreement with an experimental value of 17.5 kcal mol−1. Their difference implies that there would be a 25-fold increase in the enzyme's activity, in line with experiments. The solvent hydrogen–deuterium kinetic isotope effect was computed to be ∼3.8 for the second step in both cases. It is also found that the energy barriers are significantly and systematically higher on the QM/MM B3LYP and QM/MM B3LYP-D3 potential energy surfaces. In particular, the QM/MM B3LYP and B3LYP-D3 methods fail to predict the neutral tetrahedral intermediate and a meaningful transition state for the third step, leading to a two-step mechanism. With a sufficiently large basis set such as aug-cc-pVDZ, QM/MM M05-2X, M06-2X, M06, and MN15 methods can give results much closer to the QM/MM MP2 method. However, when a smaller basis set such as 6-31G* is used, these methods can lead to errors as large as 10 kcal mol−1 on the reaction pathway. These results highlight the importance of using accurate QM methods in the computational study of enzyme catalysis.
Eukaryotic HDACs are grouped into four classes. Class I, II, and IV HDACs are zinc-dependent enzymes containing a zinc catalytic binding domain with amino acid residues in similar conformations.15,16 Due to their biomedical relevance, class I HDACs have been widely studied. Many crystal structures have been reported for class I HDACs, especially Human histone deacetylase 8 (HDAC8).17 Buried at the bottom of a narrow hydrophobic tract (Fig. 1), the active site in the HDAC8 enzyme has a zinc ion coordinating with two aspartates, a histidine, and a catalytic water molecule in the first coordination sphere. The acetamide carbonyl group of the acetylated lysine in the substrate coordinates with the Zn ions and interacts with the hydroxyl group of a tyrosine residue (Tyr306), which can rotate and undergo a conformational transition to accommodate substrate binding. It is worth noting that there are two K+ binding sites (site 1 and site 2) in HDAC8, as shown in Fig. 1. Site 1 is about 7 Å away from the catalytic zinc ion, while site 2 is near the periphery of the protein. The functional roles of the K+ ions at these two sites in HDAC8 catalysis have been probed experimentally. Vannini et al. found that K+ ions can increase the thermal stability of HDAC8 but did not differentiate the contributions from site 1 and site 2.18 Fierke and co-workers found that the enzyme activity is sensitive to changes in the K+ ion concentration in solution.19
Two possible mechanisms for HDACs have been proposed in the literature (Fig. 2A and B). Based on X-ray crystallographic analysis, Finnin et al. proposed the first mechanism (mechanism 1 in Fig. 2A).20 In this two-step mechanism, the active site involves a singly protonated His142 and a doubly protonated His143, and the His142–Asp176 dyad acts as a general base to accept a proton from the active site water molecule. The water oxygen then attacks the carbonyl carbon of the acetyl-lysine substrate followed by the His143–Asp183 dyad protonating the amine N atom. Zhang and co-workers proposed another two-step mechanism (mechanism 2 in Fig. 2B) based on combined quantum mechanical and molecular mechanical (QM/MM) geometry optimization and molecular dynamics (MD) simulations.21,22 They found that both His142 and His143 are singly protonated on the δ nitrogen position and the His143–Asp183 dyad of human HDAC8 acts as a general base for accepting a water proton in the initial nucleophilic attack to form the negative tetrahedral intermediate. Chen et al. revisited the two two-step mechanisms with a QM/MM geometry optimization method and found it difficult to energetically distinguish which histidine residue, His143 or His142, should be protonated in the initial step, raising controversy over the catalytic mechanism of HDAC8.23 Interestingly, Chen et al. suggested that removing the K+ ion from site 1 increases the reactivity of HDAC8,23 while Wu et al. found a decrease.21,22 Using a QM/MM geometry optimization method, Gleeson et al.24 examined several possible pathways and found that the most favorable catalytic pathway is mechanism 2, where the His143 is protonated in the initial step with a comparable but lower energy barrier than that of His142.
In general, the discrepancies in the catalytic mechanism obtained with different QM/MM methods may result from the use of different starting structures, QM methods, QM region sizes, and the treatment of QM and MM coupling. The choice of QM methods is often the most prominent factor that causes the differences in QM/MM computed energies. It is well known that popular density functional theory (DFT) methods, which have been used in the above QM/MM studies, tend to underestimate weak interactions in chemical systems.25 For example, the B3LYP26,27 method tends to underestimate the intermolecular dispersion interaction, especially for systems involving ligand coordination of transition metals.28 By directly introducing electron correlation into Hartree–Fock energy, Møller–Plesset second-order perturbation theory (MP2)29 methods are considered more accurate in describing intermolecular interactions than DFT methods for many closed-shell chemical systems.30 A multi-configuration self-consistent-field (MCSCF)31 method is another type of post-Hartree–Fock method that is the best choice for describing static electronic correlation but is insufficient for dynamic electronic correlation without going to a very large active space or using a subsequent perturbation treatment. More accurate post-Hartree–Fock methods such as coupled cluster methods32 are prohibitively expensive for the geometry optimization and transition state search of enzyme active sites consisting of ∼100 atoms. The MP2 method is probably the only practical post-Hartree–Fock method that can be routinely used for QM/MM geometry optimization and transition state search of enzymatic systems.
In this work, we performed high-level QM/MM MP2 calculations to explore the HDAC8 catalytic reaction pathway and used the coupled-cluster singles, doubles, and non-iterative triples [i.e., CCSD(T)]33–35 method on small model systems extracted from the QM/MM MP2 optimized active site structures on the pathway to assess and confirm the accuracy of the MP2 method. A diacetylated tetrapeptide derived from a p53 tumor suppressor protein was used as the substrate. A crystal structure of an HDAC8 mutant containing this tetrapeptide is available (details in the next section). This substrate has been extensively used in experimental studies because it retains sufficient binding and catalytic relevance for histone deacetylases.36–38 In particular, it preserves the key interactions at the active site, thus is a good model substrate for the study of the hydrolytic mechanism. Computational mechanistic studies in the literature also used the same tetrapeptide substrate.21–24 Using a common substrate ensures a direct comparison between our results and the experimental and theoretical results in the literature. The hydrolytic mechanism obtained for this substrate can also be compared to those found for other substrates.
For HDAC8 with a K+ ion at site 1, we found that an additional intermediate is involved in the catalytic pathway, leading to a three-step catalytic mechanism. When no K+ ion is at site 1, it reduces to a two-step mechanism. Due to variations in the settings of QM/MM calculations, it is usually difficult to directly compare different QM/MM results in the literature. Our QuanPol39 program is a seamless and full-spectrum QM/MM program that can work with different QM methods and MM methods within the same framework, providing a great opportunity for comparing the accuracies of different QM methods. We performed QM/MM B3LYP26,27 (with and without an empirical dispersion correction) calculations and found significant structural and energetic differences from the QM/MM MP2 results. The QM/MM MP2 method appears to offer more accurate descriptions of the structures and energetics, largely due to a better description of the long-range dispersion interaction among the groups at the active site. We also compared the QM/MM results of the M05-2X,40 M06,41 M06-2X,41 and MN1542 density functional methods to QM/MM MP2 results and found that these functionals work reasonably well when a sufficiently large basis set is used.
The X-ray crystal structure of a Tyr306Phe mutated HDAC8-substrate complex (2V5W.PDB, resolution 2.0 Å)36 was obtained from the Protein Data Bank (PDB44). The file was modified by deleting all chains and the corresponding water molecules except for chains A and I.36 Chain A is a monomer of the asymmetric dimer crystal structure containing two structurally essential K+ ions and the active site coordinating Zn2+; Chain I is a tetrapeptide substrate (a p53 tumor suppressor protein derived diacetylated peptide): (acetyl)-L, Arg-L, His-L, Lys (e-acetyl)-L, Lys(e-acetyl). The Phe306 was changed back to Tyr306 as in the wild type.45 The Chimera program46 was used to add hydrogen atoms to the protein portion (chain A) and the water molecules, with all histidines in the active site (His142, His143, and His180) singly protonated on the δ nitrogen. The QuanPol program was used to assign the AMBER47–49 ff12SB50 force field to chain A. The QuanPol three-point flexible water model QP301 was used for the water molecules.
Using the QuanPol option LOUT = 1, the tetrapeptide substrate was assigned a universal force field for all 120 atoms with a zero net charge. Then the five H atoms of the Arg residue were manually changed from +0.30e to +0.50e so the total charge of the tetrapeptide was +1e. This universal force field has Lennard-Jones (LJ) potential parameters identical to the AMBER force field used here. The atomic charges and covalent parameters are similar to the MMFF9451 force field. Using a slightly different force field for the substrate would not cause significant differences in the results as the substrate has much fewer atoms than the enzyme. The differences in the force fields have essentially no direct (but some indirect) impacts on the subsequent QM/MM calculations, in which most of the substrate atoms are in the QM region. For an arbitrary organic molecule such as the diacetylated tetrapeptide substrate studied here, there is no standard AMBER force field, so an ad hoc force field or a universal force field like MMFF94 needs to be generated and mixed with the standard AMBER protein force field.
The QuanPol option ICOMBIN = 1 was used to combine the enzyme (chain A) and substrate (chain I) force field files. After combination, the total number of atoms was 6857, with a net charge of −2e. The enzyme–substrate system was then solvated in an 84 Å × 72 Å × 83 Å rectangular periodic boundary condition (PBC) box and filled with 25 Na+ atoms, 23 Cl− ions, and 13673 water molecules. The QuanPol three-point flexible water model QP301 was used for all water molecules. The whole system had a total of 47
924 atoms and a zero net charge.
The system was equilibrated for 10 ns using MD simulations. To preserve the experimentally measured X-ray structure of the zinc center, seven atoms were fixed in their original Cartesian coordinates in the crystal structure: the Zn2+ ion, the O of the catalytic H2O, the HZ of Tyr306, the N-term O and C atoms of the tetrapeptide substrate, and the two K2+ ions. Temporarily fixing these atoms in a force field calculation was necessary because the zinc ion did not have covalent force field parameters. Our experiences show that merely using charge–charge and LJ potentials to describe the Zn–ligand coordination in force field MD simulation and geometry optimization can lead to unrealistic active site structures (especially the bond angles and dihedral angles) that are far away from the experimental observation or QM/MM calculation. In earlier studies of Zn enzymes, we also found that for non-reactive Zn–ligand coordinate bonds, QM/MM optimized structures are often very close to experimentally measured crystal structures for the corresponding states.52,53 Therefore, when a covalent force field for the Zn active site is not available, the default practice would be fixing several active site atoms in their original crystal coordinates to prevent unpredictable disruptions of the active site. These atoms were not fixed in the later QM/MM geometry optimization. A shifting function (QuanPol keyword ISHIFT = 4) was used to adjust the long-range charge–charge interactions up to 12 Å. A switching function (QuanPol keyword ISWITCH = 1) was used to adjust the long-range Lennard-Jones (LJ) interactions from 10 to 12 Å. The Berendsen thermostat and barostat were used with a temperature scaling time constant of 200 fs and a pressure scaling time constant of 200 fs, and with a bath temperature T of 298.15 K and bath pressure P of 1.00 bar. The MD simulations were run using the Beeman integration algorithm.54 At the end of the 1 ns MD simulation, the volume of the PBC box was 81.91 Å × 70.21 Å × 80.94 Å. At the end of the 10 ns MD simulation, the volume of the PBC box was 81.95 Å × 70.24 Å × 80.98 Å. A superposition of the geometries from the 1 ns and 10 ns MD simulation shows that the two geometries are similar (Fig. S3, ESI†). The last geometries from the 1 ns and 10 ns MD simulation were used for subsequent QM/MM B3LYP,26,27 B3LYP with Grimme's empirical dispersion corrections (denoted B3LYP-D355) and MP229 geometry optimizations.
The QM/MM geometry optimizations for the 1 ns snapshot were performed for 1844 atoms within 16 Å of the oxygen atom of the catalytic water molecule in the presence of the rest of the 46080 atoms. The same 1844 atoms were optimized in all QM/MM calculations for the 1 ns snapshot. For the 10 ns snapshot, 1876 atoms within 16 Å of the oxygen atom of the catalytic water molecule were optimized, and the same 1876 atoms were optimized in all QM/MM calculations for the 10 ns snapshot. A comparison of the QM/MM structures and energies obtained with 1 ns and 10 ns snapshots (Table S6, ESI†) suggests that the thermal fluctuations in the protein matrix have an insignificant effect on the reaction energies and do not alter the catalytic mechanism. The details can be found in the ESI.† Discussions in the rest of the paper will focus on the results obtained with a 1 ns snapshot unless specifically mentioned.
In the QM/MM calculations, 105 atoms were in the QM region (Fig. 1): the Zn2+ ion, the catalytic H2O, His142, His143, His180, Asp176, Asp178, Asp183, Asp267, and the acetyl-lysine side chains of the tetrapeptide substrate. The net charge of the QM region was −2e. The QM/MM boundary was fragmented at the Cα–Cβ bonds of each respective side chain and the default capping H atom method implemented in QuanPol was used to treat the covalent bonds between the QM and MM regions.39 The 6-31G* basis set56 was used for 82 QM atoms. The aug-cc-pVDZ basis set57,58 was used for the 23 most relevant atoms, which were the Zn ion, the three atoms of the H2O molecule, four carboxylate O atoms of Asp178 and Asp267, the two carboxylate O atoms of Asp176 and Asp183, the four N and H atoms of His142 and His143, one N atom of His180, and the four N, H, C, and O atoms of the tetrapeptide substrate. In all the QM/MM calculations, the QM–MM LJ interaction was used (the QM atoms used their original MM LJ parameters). The force field charges of the MM atoms interact directly with the electrons and nuclei of the QM atoms and affect the Hartree–Fock wavefunction and subsequent MP2 energy and gradient.59–61 A shifting function (QuanPol keyword ISHIFT = 4) was used for MM charge–charge interaction up to 12 Å. A switching function (QuanPol keyword ISWITCH = 1) was used for MM LJ interaction from 10 to 12 Å. A QM–MM switching function (QuanPol keyword ISWITCH = 1) was used for QM–LJ, QM–charge interactions from 22 to 32 Å. The parallel MP2 program implemented by Ishimura, Pulay, and Nagase et al.62,63 was used together with the QuanPol routines that add MM interactions to the MP2 method.60
All QM/MM geometry optimizations and transition state searches were performed with the Hessian guided optimization method we implemented in the QuanPol program. First, all 1844 (for 1 ns snapshot) or 1876 (10 ns snapshot) atoms were optimized using a steepest descent scheme at the MM level. The Hessian for these atoms was then computed at the MM level and used to guide subsequent QM/MM geometry optimization. The QM/MM geometry optimization was started at the Hartree–Fock level (with the 6-31G* basis set and Grimme's dispersion correction) for the ES state, using the MM Hessian as the guide. The Hessian was systematically updated and improved as the QM/MM geometry optimization proceeded to finish. Based on the optimized ES structure and the improved Hessian, other states were constructed (by repositioning a few atoms) and optimized. All transition states were optimized with a saddle point search method by preparing a trial structure and supplying a Hessian matrix containing a leading negative mode. We have implemented a convenient and low-cost method in the QuanPol program to compute and update part of an existing and all-positive-mode Hessian to produce a new Hessian that (if lucky) contains one negative mode for a trial TS structure. The QM/MM optimized structures and the improved Hessian matrices obtained at the Hartree–Fock level of theory can be used as starting points for subsequent B3LYP, B3LYP-D3, and MP2 optimization. All transition states found at the B3LYP-D3 level of theory were confirmed with the partial Hessian vibration analysis method implemented in QuanPol. The results are shown in Table S1 (ESI†). The relative electronic energies of each species are in Table 1. The corresponding QM/MM MP2 and spin component scaled MP2 (SCS-MP264) single point energies (using four different basis sets) were calculated for all QM/MM MP2 optimized structures. The relative single-point energies are given in Table S2 (ESI†). The optimized coordinates of the enzyme and substrate at various reaction stages are available as PDB files in the ESI.†
ES | TS1a | EI1a | TS1 | EI1 | TS2 | EI2 | TS3 | EI3 | EP | |
---|---|---|---|---|---|---|---|---|---|---|
With a K+ ion at site 1 | ||||||||||
B3LYP/MM | 0.0 | 18.3 | 17.1 | 24.1 | 21.3 | 30.6 | 17.2 | 18.9 | 10.8 | 1.5 |
B3LYP-D3/MM | 0.0 | 16.9 | 15.5 | 22.9 | 19.0 | 25.8 | 16.5 | 19.4 | 13.5 | 6.1 |
MP2/MM | 0.0 | 13.7 | 10.5 | 17.3 | 13.7 | 19.7 | 10.5 | 17.1 | 13.4 | 9.6 |
No K+ ion at site 1 | ||||||||||
B3LYP-D3/MM | 0.0 | 9.5 | 5.5 | 19.9 | 16.5 | 24.4 | 15.1 | 16.7 | 4.9 | −1.7 |
MP2/MM | 0.0 | 5.4 | −1.8 | 12.6 | 10.4 | 17.4 | 8.9 | 10.5 | 2.6 | 0.0 |
The first step is the nucleophilic addition of the Zn-bound water molecule to the acetamide carbonyl carbon atom (Csub). To activate the water molecule, either His142 or His143 (both with only δ-N protonated) could accept a proton from the water molecule. These two possibilities are examined with the MP2/MM geometry optimization method. Similar to the results of Chen et al. and Gleeson et al.,23,24 we found that a water proton (H2wat) can pass a transition state (TS1, Fig. 4B) and bind to His143 (EI1 in Fig. 4C). The electronic energy for the transition state TS1 is 17.3 kcal mol−1. In the ES state (Fig. 4A), the distance between the Zn ion and the O atom of the carbonyl group of the substrate (see Zn–Osub in Table S4, ESI†) from the MP2/MM geometry optimization is 2.16 Å. In TS1, this distance is shortened to 1.93 Å. The proton transfer from the water to His143 is accompanied by the nucleophilic attack of the partially formed hydroxide to the carbonyl carbon atom (Csub), leading to a negatively charged tetrahedral intermediate (EI1 in Fig. 4C). The relative electronic energy for EI1 is 13.7 kcal mol−1. The Owat–Csub distance is shortened from 2.55 Å in ES (Fig. 4A) to 1.50 Å in TS1 (Fig. 4B). The interaction between the Zn ion and Owat is significantly weakened in EI1, as shown by the elongation of the distance between the two atoms (from 2.09 Å in ES to 2.51 Å in TS1). The distance between Owat and the C atom of the carbonyl group of the substrate (Owat–Csub in Table S4, ESI†) is 1.48 Å in the negative tetrahedral intermediate EI1, longer than the standard carboxylic C–O distance of 1.34 Å.
We also found that a water proton (H1wat) can transfer to His142. The electronic energy for the transition state (TS1a, Fig. S7B, ESI†) is 13.7 kcal mol−1 (relative to ES). While the proton transfers from the water to His142, the partially formed hydroxide approaches the carbonyl carbon atom (Csub), leading to a negatively charged tetrahedral intermediate (EI1a in Fig. 5A and Fig. S7C, ESI†). The relative electronic energy for EI1a is 10.5 kcal mol−1. However, because the protonation of His143 (i.e., EI1) is the prerequisite for the subsequent protonation of the amide N atom, EI1a must change into EI1. Chen et al.23 found a transition state (denoted as 2TS3a′ in their paper) between EI1a and EI1. Here, our transition state search on the MP2/MM potential energy surface (PES) shows that the transition state between EI1a and EI1 is simply TS1 (Fig. 4B), the same transition state bridging ES and EI1. Therefore, on the MP2/MM PES, there are two pathways from ES to EI1: one is ES – TS1a – EI1a – TS1 – EI1; the other is ES – TS1 – EI1.
The second step is the protonation of the amide N atom by His143 and the formation of a neutral tetrahedral intermediate (EI2 in Fig. 4E). The distance between Owat and the C atom of the carbonyl group of the substrate (Owat–Csub) is shortened from 1.47 Å in EI1 (Fig. 4C) to 1.45 Å in TS2 (Fig. 4D) and finally to 1.41 Å in EI2 (Fig. 4E). The protonation of Nsub leads to a weakened amide C–N bond (from 1.46 Å in EI1 to 1.49 Å in TS2, and finally to 1.57 Å in EI2). The electronic energy of TS2 is 19.7 kcal mol−1 higher than ES. The electronic energy of the neutral tetrahedral intermediate EI2 is 10.5 kcal mol−1, 3.2 kcal mol−1 lower than the negative tetrahedral intermediate EI1 (13.7 kcal mol−1).
The third step is the complete cleavage of the amide C–N bond in the tetrapeptide substrate, accompanied by a proton transfer from the newly formed carboxylic group to His142. The proton has already transferred to His142 (H1wat–NH142 1.10 Å) when the C–N distance is at the transition state (TS3, Fig. 4F) length of 1.99 Å. The electronic energy barrier from EI2 to the transition state TS3 (Fig. 4F) is 6.6 kcal mol−1, and the electronic energy of TS3 is 17.1 kcal mol−1 higher than ES. The amide C–N distance is elongated from 1.57 Å in EI2 to 1.99 Å in TS3, and finally to 2.57 Å in the intermediate EI3 (Fig. 5C). The electronic energy of EI3 is 13.4 kcal mol−1 higher than ES. However, as an immediate local minimum just passing TS3, EI3 is not very stable because Tyr306 forms a probably weak hydrogen bond with the newly formed carboxylic group. Passing a low and flat electronic energy barrier (0.26 kcal mol−1), Tyr306 can rotate and form a (presumably stronger) hydrogen bond to the newly formed amine group (–CNH2) of the substrate, leading to a more stable structure EP (Fig. 3 and 5E). EP has a relative electronic energy of 9.6 kcal mol−1 (relative to ES), 3.8 kcal mol−1 lower than EI3 (Table 1 and Fig. 3). EP is likely the final enzyme–product complex minimum state. Because the transition state between EI3 and EP has a very low and flat electronic energy barrier, it may not be a meaningful free energy barrier when zero-point energy and thermal free energy are considered. Similarly low and flat electronic energy barriers (0.01 and 0.14 kcal mol−1) were also observed on the QM/MM B3LYP and B3LYP-D3 potential energy surfaces. Therefore, in the rest of the paper, this transition state is not discussed.
The first step in the three-step mechanism identified in this work is identical to the first step in mechanism 2 (Fig. 2B) proposed in the literature for Zn(II)-HDAC8 with two K+ ions: a negatively charged tetrahedral intermediate is formed after the initial hydroxide attack. In the two-step mechanism (mechanism 2 in Fig. 2B), the negative tetrahedral intermediate directly decomposes into the final product (via a concerted protonation of amide N by His143, C–N cleavage, and protonation of His142). In the three-step mechanism (Fig. 2C), the negative tetrahedral intermediate evolves into a neutral tetrahedral intermediate (via protonation of amide N by His143), then decomposes into the final product (via C–N cleavage and protonation of His142).
To verify the accuracy of the MP2 and SCS-MP2 methods, we performed coupled-cluster singles, doubles with non-iterative triples [i.e., CCSD(T)]33–35 calculations for a set of 34-atom active site models (Fig. S6, ESI†) extracted from the QM/MM MP2 optimized structures. These models include the core reactive groups: the Zn ion, the water, the substrate acetamide group, and the imidazole groups of His142 and His143. The SPK-ADZP65 basis set was used as it was similar to but slightly smaller than the aug-cc-pVDZ basis set and led to a significant reduction of computer resource requirements, as well as good convergence. The relative CCSD(T) corrections across the models are small, 2.1 kcal mol−1 for regular MP2 and 1.3 kcal mol−1 for SCS-MP2, at maximum (Table S7, ESI†). These results suggest that for the reaction pathway studied here, the MP2 and SCS-MP2 methods have an accuracy of about 1–2 kcal mol−1.
According to our MP2/MM calculations (as shown in Fig. 3), the relative electronic energies of TS1, TS2, and TS3 are 17.3, 19.7, and 17.1 kcal mol−1, respectively. Though these values are comparable, the second step (TS2) has the highest electronic energy. Zero-point energy (ZPE) and free energy corrections are estimated with partial Hessian harmonic frequencies. Due to the high computational cost of the MP2 method, the partial Hessian harmonic frequency calculations are performed with the B3LYP-D3/MM method. Twenty-six key reactive and zinc-binding atoms are included in the partial Hessian calculations: the Zn ion, NHis142, HHis142, NHis143, HHis143, OAsp176, OAsp178, OAsp183, NHis180, OAsp267, O, and H atoms of the catalytic water, as well as the reactive N, H, O, and C atoms in the substrate. The ZPE and thermal free energy corrections are −1.7, −0.1, and −0.5 kcal mol−1, respectively, for TS1, TS2, and TS3 (relative to ES). After correction, the activation free energies for TS1, TS2, and TS3 are 15.6, 19.6, and 16.6 kcal mol−1, respectively. TS2 still has the highest value. Therefore, our best estimation of the free energy barrier is 19.6 kcal mol−1 for 298 K, which is in good agreement with the experimental value of 17.5 kcal mol−1 for Zn(II)-HDAC8 in 137 mM NaCl and 2.7 mM KCl. If the estimated experimental value 18.1 kcal mol−1 for Zn(II)-HDAC8 (with two K+ ions) is used for comparison, the agreement is also good.
Consider the following general enzyme-catalyzed reaction and the standard (physical chemistry standard state, i.e., activity = 1.0 when solute concentration is near 1.0 M) Gibbs free energy changes on the pathway:
![]() | (1) |
ΔGo1 + ΔGo2 + ΔGo3 < 0 (can be slightly > 0) | (2) |
E + S to ES is typically exergonic at the standard state, but endergonic is not a problem:
ΔGo1 < 0 (can be >0) | (3) |
Here we use EIlowest to denote the state with the lowest standard free energy (global minimum) on the reaction pathway. It can be either ES or EP, or any other state before E + P, including E + S. EIlowest to E + P is typically exergonic at standard state, but slightly endergonic is also common, because P is always at concentrations orders lower than the standard state concentration 1 M. In case the EIlowest is ES, we should have:
ΔGo2 + ΔGo3 < 0 (can be slightly > 0) | (4) |
ΔGo3 <0 (can be slightly > 0) | (5) |
For the Zn(II)-HDAC8 (with two K+ ions) catalyzed deacetylation reaction studied here, the standard binding free energy change ΔGo1 for the R-H-K(Ac)-K(Ac) tetrapeptide and enzyme has been experimentally measured to be −4.0 kcal mol−1 (KM = 1.1 × 10−3 M) at 298.15 K.66 It was also experimentally found that the ΔGo1 is insensitive to site-1 K+ depletion and mutations of some active site residues.38 Literature results on the standard free energy change or equilibrium constant for the hydrolysis of the tetrapeptide substrate are not available. It has been experimentally measured that the standard free energy of hydrolysis of acetamide (CH3CONH2) to acetate and ammonium is −2.1 kcal mol−1,68 and is −2.4 kcal mol−1 for N,N-dimethyl-acetamide.69 The hydrolysis of the acetamide group in the R-H-K(Ac)-K(Ac) tetrapeptide substrate studied here should be close to these two values because all acetamide hydrolyses have similar free energy changes. For convenience of discussion, here we use an estimated value of −2.2 kcal mol−1. The following experimental results for HDAC8 catalysis can be established:
ΔGo1 + ΔGo2 + ΔGo3 = −2.2 kcal mol−1 | (6) |
ΔGo1 = −4.0 kcal mol−1 | (7) |
ΔGo2 + ΔGo3 = +1.8 kcal mol−1 | (8) |
The experimental relationship in eqn (8) can be used to examine ΔGo2 values from theoretical calculations. In this work, ΔGo2 is computed to be +8.8 kcal mol−1 (electronic energy 9.6 kcal mol−1 with −0.8 kcal mol−1 ZPE and thermal free energy correction, Table 2 and Table S1, ESI†) for HDAC8 with two K+ ions, so ΔGo3 must be around −7.0 kcal mol−1. Since here ES is computed to be the global minimum, the slightly positive ΔGo2 + ΔGo3 is the overall free energy of dissociation and satisfies eqn (4). This ΔGo2 + ΔGo3 value corresponds to a thermodynamic dissociation equilibrium constant of 0.05 (unit-less, using activity and physical chemistry standard states) for ES = E + P1 + P2 (note the products are acetate and ammonium, denoted as P1 and P2). This equilibrium constant allows for a highly favorable dissociation of ES into E + P at typical biological and experimental concentrations such that [E]total is 1.2–50 μM and [S] and [P] are 5–2000 μM.38 Therefore, the computed ΔGo2 value of +8.8 kcal mol−1 in this work (the direct consequence of this result is that ES is the global minimum) is reasonable and not contradictory to experimental facts. In contrast, in an earlier work by Corminboeuf et al.,21 ΔGo2 was computed to be around −8.9 kcal mol−1 for HDAC8 with two K+ ions, so the EP state was the global minimum (Table 2). Based on the experimental relationship in eqn (8), ΔGo3 must be around +10.7 kcal mol−1, which is too large to allow for a favorable dissociation of EP into E + P. Similarly, in an earlier work by Chen et al.,23 ΔGo2 was computed to be around −4.4 kcal mol−1 for HDAC8 with one K+ ion (Table 2), also leading to a large ΔGo3 and unfavorable dissociation of EP into E + P. To be discussed later, these two problematic ΔGo2 values were computed with inaccurate QM methods and/or a small basis set.
QM method | ES | TS1a | EI1a | TS1 | EI1 | TS2 | EI2 | TS3 | EI3 | EP |
---|---|---|---|---|---|---|---|---|---|---|
With a K+ ion at site 1 | ||||||||||
MP2/BS1 | 0.0 | 13.7 | 10.5 | 17.3 | 13.7 | 19.7 | 10.5 | 17.1 | 13.4 | 9.6 |
SCS-MP2/BS1//MP2/BS1 | 0.0 | 15.5 | 12.1 | 20.9 | 15.6 | 22.1 | 12.2 | 18.1 | 12.5 | 7.6 |
B3LYP/BS1 | 0.0 | 18.3 | 17.1 | 24.1 | 21.3 | 30.6 | 17.2 | 18.9 | 10.8 | 1.5 |
B3LYP-D3/BS1 | 0.0 | 16.9 | 15.5 | 22.9 | 19.0 | 25.8 | 16.5 | 19.4 | 13.5 | 6.1 |
M05-2X/BS1//MP2/BS1 | 0.0 | 15.9 | 13.4 | 20.6 | 16.9 | 22.4 | 13.4 | 19.0 | 15.0 | 10.3 |
M06/BS1//MP2/BS1 | 0.0 | 16.9 | 13.6 | 20.2 | 16.8 | 22.5 | 13.5 | 20.0 | 15.3 | 11.5 |
M06-2X/BS1//MP2/BS1 | 0.0 | 17.8 | 15.4 | 23.0 | 18.3 | 25.5 | 15.9 | 20.1 | 13.6 | 10.2 |
MN15/BS1//MP2/BS1 | 0.0 | 15.0 | 11.4 | 20.1 | 14.7 | 21.5 | 13.5 | 19.0 | 13.6 | 10.5 |
B3LYP/6-31G*//MP2/BS1 | 0.0 | 15.4 | 12.9 | 22.2 | 16.7 | 23.3 | 13.0 | 13.2 | 6.4 | −1.2 |
M05-2X/6-31G*//MP2/BS1 | 0.0 | 13.7 | 9.0 | 16.9 | 12.1 | 18.0 | 8.6 | 14.1 | 10.5 | 7.1 |
M06/6-31G*//MP2/BS1 | 0.0 | 15.6 | 9.4 | 17.0 | 12.2 | 18.3 | 8.9 | 14.8 | 10.7 | 8.3 |
M06-2X/6-31G*//MP2/BS1 | 0.0 | 15.6 | 10.8 | 19.2 | 13.6 | 21.4 | 11.0 | 15.0 | 8.4 | 6.4 |
MN15/6-31G*//MP2/BS1 | 0.0 | 12.3 | 5.8 | 15.3 | 8.7 | 16.2 | 7.7 | 12.8 | 7.0 | 4.9 |
Corminboeuf et al.21 | 0.0 | — | — | 15.0 | 10.6 | 13.7 | — | — | −8.9 | — |
Wu et al.22 | 0.0 | — | — | 18.3 | ∼13.0 | ∼15.3 | — | — | −7.1 | — |
Chen et al.23 | 0.0 | 14.0 | 9.8 | 15.2 | 10.2 | 9.6 | 4.4 | 10.8 | 8.5 | — |
Gleeson et al.24 | 0.0 | 17.9 | 10.8 | — | — | 12.1 | — | — | −2.2 | — |
No K+ ion at site 1 | ||||||||||
MP2/BS1 | 0.0 | 5.4 | −1.8 | 12.6 | 10.4 | 17.4 | 8.9 | 10.5 | 2.6 | 0.0 |
SCS-MP2/BS1//MP2/BS1 | 0.0 | 6.5 | −0.7 | 16.3 | 12.4 | 19.9 | 10.6 | 13.0 | 1.4 | −2.3 |
B3LYP/BS1 | 0.0 | 9.5 | 5.5 | 19.9 | 16.5 | 24.4 | 15.0 | 16.7 | 4.9 | −1.7 |
M05-2X/BS1//MP2/BS1 | 0.0 | 8.5 | 1.3 | 16.4 | 13.7 | 20.2 | 11.7 | 13.6 | 4.8 | 1.3 |
M06/BS1//MP2/BS1 | 0.0 | 8.5 | 0.6 | 16.4 | 13.4 | 20.3 | 11.7 | 13.9 | 4.3 | 2.1 |
M06-2X/BS1//MP2/BS1 | 0.0 | 7.8 | 1.4 | 19.9 | 14.5 | 22.1 | 12.9 | 15.7 | 1.1 | −0.3 |
MN15/BS1//MP2/BS1 | 0.0 | 6.5 | −1.2 | 15.9 | 11.4 | 18.8 | 11.3 | 13.7 | 2.4 | 0.8 |
M05-2X/6-31G*//MP2/BS1 | 0.0 | 4.6 | −5.8 | 12.4 | 8.8 | 15.6 | 6.8 | 9.2 | −2.2 | −4.1 |
M06/6-31G*//MP2/BS1 | 0.0 | 4.7 | −6.1 | 12.4 | 8.6 | 15.7 | 6.9 | 9.5 | −2.6 | −3.0 |
M06-2X/6-31G*//MP2/BS1 | 0.0 | 4.2 | −5.5 | 15.7 | 9.6 | 17.8 | 7.9 | 11.3 | −6.4 | −6.1 |
MN15/6-31G*//MP2/BS1 | 0.0 | 2.5 | −9.2 | 10.7 | 5.2 | 13.5 | 5.4 | 8.7 | −6.3 | −6.6 |
Wu et al.22 | 0.0 | — | — | 22.6 | — | — | — | — | — | — |
Chen et al.23 | 0.0 | 8.5 | −0.6 | 11.1 | 7.4 | 8.9 | 2.9 | 6.2 | −4.4 | — |
The results are shown in Table 1 and Fig. 3. In general, the reaction pathways with and without a K+ ion at site 1 are very similar and can be characterized by the same set of intermediates and transition states (with the same names).
The first step is the nucleophilic addition of the Zn-bound water molecule to the amide carbonyl carbon atom (Csub). The ES (Fig. S8A, ESI†) can pass TS1 (Fig. S8D, ESI†) and evolve into EI1 (Fig. S8E, ESI†) by transferring a proton from water to His143. It can also pass TS1a (Fig. S8B, ESI†) and evolve into EI1a (Fig. 5B and Fig. S8C, ESI†) by transferring a proton from water to His142. The EI1a state has an electronic energy −1.8 kcal mol−1 lower than ES. When zero-point energy and thermal free energy (+1.7 kcal mol−1, Table S1, ESI†) are included, the free energy of EI1a is −0.1 kcal mol−1 lower than ES. SCS-MP2 calculation with a larger basis set (Table S2, ESI†) suggests that the free energy of EI1a should be a couple of tenths kcal mol−1 higher than ES. In EI1a the proton on δ-NH142 is already transferred to Asp176, suggesting that the removal of the K+ ion from site 1 leads to an increment in the electronegativity of the oxygen atom of Asp176. Similar to the case with a K+ at site 1, our transition state search on the MP2/MM PES shows that the transition state between EI1a and EI1 is again TS1 (Fig. S8D, ESI†), the same transition state bridging ES and EI1.
The second step is the protonation of the amide N atom by His143 and the formation of a neutral tetrahedral intermediate (EI2 in Fig. S8G, ESI†). The electronic energy of TS2 (Fig. S8F, ESI†) is 17.4 kcal mol−1 (relative to ES). The electronic energy of the neutral tetrahedral intermediate EI2 is 8.9 kcal mol−1, 1.5 kcal mol−1 lower than the negative tetrahedral intermediate EI1 (10.4 kcal mol−1).
The third step is the complete cleavage of the amide C–N bond in the tetrapeptide substrate. Different from the one with K+ ions at site 1, this step involves a concerted transition state (TS3 in Fig. S8H, ESI†), where C–N bond breaking is associated with a proton transfer from the newly formed carboxylic group to His142. In TS3, the proton is between N of His142 and water (H1wat–NH142 1.25 Å and H1wat–Owat 1.26 Å). Passing TS3, as H1wat is completely transferred to His142, the proton on δ-NH142 would transfer to Asp176, finally leading to the EI3 state (Fig. 5D and Fig. S8I, ESI†). Therefore, the EI3 state for a K+-free site 1 is different from that for a K+-bound site 1 due to an increased proton binding ability of Asp176 upon K+ removal (Fig. 5C and D). The electronic energy barrier from EI2 to the transition state TS3 (Fig. S8H, ESI†) is 1.7 kcal mol−1. The ZPE and thermal free energy correction is −2.9 kcal mol−1 from EI2 to TS3. After the correction, the free energy barrier from EI2 to TS3 is −1.3 kcal mol−1 (not really a barrier). Therefore, after K+ depletion from site 1, it is likely a two-step mechanism. The electronic energy of EI3 (Fig. S8I, ESI†) is 2.6 kcal mol−1 higher than ES. Passing a low and flat (0.47 kcal mol−1) electronic energy barrier, Tyr306 can rotate and form a hydrogen bond with the newly formed amine group (–CNH2) of the substrate, leading to a more stable structure EP (Fig. 3 and 5F). Because the transition state between EI3 and EP is low and flat, it may not be a meaningful free energy barrier. In the rest of the paper, this transition state is not discussed. EP has a relative electronic energy of 0.0 kcal mol−1 (relative to ES, Table 1 and Fig. 3). In both EI3 and EP (Fig. 5D and F) the proton on δ-NH142 is already transferred to Asp176 due to an increased proton binding ability of Asp176 upon K+ removal. As discussed earlier, when a K+ ion is at site 1, the relative electronic energy for EP is 9.6 kcal mol−1. Clearly, a K+-free site 1 significantly favors the EP state, due to a proton transfer from His142 to Asp176. However, the stability of the EP state is irrelevant to whether the catalytic reaction can proceed or not. HDAC8 can catalyze the reaction in both cases: with and without K+ at site 1. The only difference is the catalytic efficiency, as discussed below.
The relative electronic energies of TS1a, EI1a, TS1, and TS2 are 5.4, −1.8, 12.6, and 17.4 kcal mol−1, respectively. TS2 remains to be the rate-determining transition state. Since EI1a (–1.8 kcal mol−1 in relative electronic energy, −0.1 kcal mol−1 in relative free energy) is the lowest energy point in the reaction pathway (Fig. 3), the activation free energy should be derived using EI1a and TS2. The ZPE and thermal free energy correction for EI1a to TS2 is −1.5 kcal mol−1 (Table S1, ESI†). Therefore, our best estimation of the free energy barrier is 17.7 kcal mol−1 for 298 K. The free energy barrier (17.7 kcal mol−1) for K+-free site 1 is 1.9 kcal mol−1 lower than that for K+-bound site 1, implicating a 25-fold difference in the catalytic activity of Zn(II)-HDAC8. For comparison, experimental kinetic studies found that the Co(II)-HDAC8 exhibited an 11-fold difference in activity between K+-free and K+-bound site 1.19 The MP2/MM result obtained here is in good quantitative agreement with the site-1 K+ depletion experimental values. The computed free energy barrier of 17.7 kcal mol−1 for Zn(II)-HDAC8 with only a site-2 K+ is in good agreement with the experimental value of 17.5 kcal mol−1 for Zn(II)-HDAC8 in 137 mM NaCl and 2.7 mM KCl. If the estimated experimental value of 16.7 kcal mol−1 for Zn(II)-HDAC8 with only a site-2 K+ is used for comparison, the agreement is also good.
In summary, the K+ depletion from site 1 significantly increases the proton binding ability of Asp176, which then preferentially stabilizes TS1a, EI1a, and TS1 in the first step, and TS3, EI3, and EP in the third step, by providing stronger hydrogen bonding through His142 or even accepting the proton from His142. In contrast, the stabilities of EI1, TS2, and EI2 are much less affected by this K+ depletion because the reactive proton is on the His143 side in these states.
Using the QM/MM B3LYP-D3 optimized structures (which are similar to those from QM/MM MP2 optimization), the sKIEs were computed to be 8.9 and 4.0, respectively, for TS1 and TS2 (relative to EI1a). Because the computed free energy (with H atoms, not D) of EI1a is only lower than the ES state by 0.1 kcal mol−1, when computational errors are considered, ES may be lower than EI1a. Using ES as the reference, the computed sKIEs are 8.5 and 3.8, respectively, for TS1 and TS2 (relative to ES). These two sets of sKIEs (relative to EI1a and ES, respectively) are close to each other. Here, the chance of TS1 being the rate-determining step is very low because it is 7 kcal mol−1 lower than TS2 in computed free energy. Therefore, the experimentally measured sKIE should be mainly due to TS2 when site 1 has no K+. As described earlier, when there is a K+ ion at site 1, the TS2 isotope effect was computed to be 3.8. Therefore, the computed sKIEs are insensitive to K+ occupation at site 1, implying that the apparent sKIE would remain at roughly a constant (computed to be ∼3.8 here) when the K+ concentration varies (the ratio of mono-K+-HDAC8 to bi-K+-HDAC8 changes). This prediction can be verified by experiments. If the experimentally measured sKIE remains constant at ∼3.8 (or another value as experimentally measured, such as 6.2 or 5.3), it may suggest that the TS2 is the rate-determining transition state in both cases (site 1 has K+ or not). If the sKIE increases at higher K+ ion concentrations, it may suggest that the TS1 is rate-determining when both site 1 and site 2 have K+ ions (in this case, the TS1 value is 5.5). If the sKIE decreases at higher K+ ion concentrations, it may suggest that the TS3 is rate-determining when both site 1 and site 2 have K+ ions (in this case, the TS3 value is 1.2).
The Zn–Owat and Zn–Osub coordination bond distances optimized with B3LYP/MM are generally longer than those optimized with B3LYP-D3/MM and MP2/MM (Fig. 6). For example, the Zn–Owat distances in EI1 (Fig. 6A and Table S5, ESI†) optimized with B3LYP/MM (2.51 Å) are 0.08 Å and 0.14 Å longer than that with B3LYP-D3/MM (2.43 Å) and MP2/MM (2.37 Å), respectively. In the ES, the Zn–Osub distances (Fig. S1, ESI†) optimized with B3LYP/MM (2.40 Å) and B3LYP-D3/MM (2.17 Å) are longer than 2.16 Å optimized with the MP2/MM method. For comparison, the Zn–Osub distance in the crystal structure (2V5W.PDB) is 2.02 Å. In the crystal structure, Tyr306 is replaced by Phe306, so the hydrogen bond between the Osub atom and Tyr306 does not exist. This would leave a slightly stronger interaction and a shorter distance between the Osub atom and the Zn ion. Indeed, the MP2/MM optimized Zn–Osub distance (2.16 Å, with Tyr306) is slightly longer than the 2.02 Å (with Phe306) in the crystal structure. The B3LYP/MM optimized Zn–Owat and Zn–Osub distances in other reaction species (i.e. TS1a, EI1a, TS1, EI1, TS2, EI2, TS3, EI3, and EP) are slightly longer than those optimized with B3LYP-D3/MM and MP2/MM methods. The B3LYP method tends to underestimate the interactions between the Zn ion and its ligands and predict longer coordinate distances.
The relative electronic energies obtained with the B3LYP/MM method are 24.1, 30.6, and 18.9 kcal mol−1, respectively, for TS1, TS2, and TS3 (Fig. 7, all compared to ES). TS2 is the highest, thus rate-determining. However, it is ∼13 kcal mol−1 larger than the experimental value of ∼17.5 kcal mol−1. Inclusion of Grimme's empirical dispersion correction to B3LYP (i.e., B3LYP-D3) leads to optimized structures and energies closer to the MP2 results: the B3LYP-D3/MM optimized relative electronic energies are 22.9 kcal mol−1 and 25.8 kcal mol−1 for TS1 and TS2, respectively (Fig. 7). Here, the empirical dispersion correction leads to more than 4 kcal mol−1 energy improvement for the rate-determining step, which is significant. Our previous work on cytochrome P450cam obtained similar corrections.70 Even with the empirical dispersion correction, the B3LYP-D3/MM relative energies of TS1 and TS2 are still 5.6 and 5.9 kcal mol−1 higher than the MP2/MM values 17.3 and 19.7 kcal mol−1.
For TS3, the B3LYP/MM and B3LYP-D3/MM energies, 18.9 and 19.4 kcal mol−1, respectively, are comparable to the MP2/MM value of 17.1 kcal mol−1. However, the B3LYP/MM and B3LYP-D3/MM computed electronic energy barriers from EI2 to TS3 are only 1.7 and 2.9 kcal mol−1, respectively. For comparison, the MP2/MM computed barrier from EI2 to TS3 is 6.6 kcal mol−1 (SCS-MP2/MM results are similar, see Table S2, ESI†). We note that, in MP2/MM optimized TS3, the proton has already transferred to the N atom of His142, while in B3LYP/MM and B3LYP-D3/MM optimized TS3 structures, the proton is still on the oxygen atom of water. When ZPE and free energy correction (which is −1.7 kcal mol−1 ongoing from EI2 to TS3 as computed with the B3LYP-D3/MM partial Hessian method) is applied, the free energy barrier becomes 0.0 and +1.3 kcal mol−1, respectively, on the B3LYP/MM and B3LYP-D3/MM PESs. Therefore, due to the very low electronic energy barriers, TS3 is not a meaningful transition state on the B3LYP/MM and B3LYP-D3 PESs. Even when only electronic energy is considered, there is a good chance that the TS3 would not appear in a QM/MM B3LYP (or B3LYP-D3) calculation with settings different from those used in this work, especially when different starting structures and basis sets are used. For example, when the small 6-31G* basis set is used, the QM/MM B3LYP relative single-point energies of EI2 and TS3 (based on QM/MM MP2 optimized structures) are 13.0 and 13.2 kcal mol−1, respectively, virtually no difference (Table 2). It is not surprising that the two-step mechanism (mechanism 2 in Fig. 2B) was obtained with B3LYP/MM.21,22 Other QM/MM DFT methods may also predict the same two-step mechanism.24 Using a QM/MM DFT method, Chen et al. obtained a N-protonated neutral tetrahedral intermediate similar to the neutral tetrahedral intermediate EI2 in this work, but the proton transfer from His143 to the amide N atom is barrier-less, and the subsequent C–N bond cleavage requires a very low activation energy, thus not a meaningful transition state.23 Their results resemble the B3LYP/MM and B3LYP-D3/MM results in this work.
Some key interatomic distances involved in the reaction are shown in Fig. S2 (ESI†). Most of the distances optimized with B3LYP/MM are longer than those optimized with B3LYP-D3/MM and MP2/MM. For example, the B3LYP/MM optimized Csub–Owat distance is 2.73 Å in ES and 1.70 Å in TS1. The B3LYP-D3/MM optimized Csub–Owat distances are shorter, 2.67 Å in ES and 1.60 Å in TS1. The MP2/MM optimized Csub–Owat distances are the shortest, 2.55 Å in ES and 1.50 Å in TS1. The B3LYP/MM optimized H1wat–NH142 and H2wat–NH143 distances (Fig. S2C and S2D, ESI†) are longer than B3LYP-D3/MM and MP2/MM values on the reaction pathway. These longer distances are a general reflection of the fact that the B3LYP functional lacks long-range dispersion. Empirical dispersion correction (B3LYP-D3) can only partially remedy this problem.
Interestingly, the relative energies obtained with MP2/MM single-point energy calculation based on B3LYP/MM and B3LYP-D3/MM optimized structures resemble the MP2/MM optimized relative energies (Table S4, ESI†). Using MP2/MM optimized values (with mixed aug-cc-pVDZ and 6-31G* basis set) as references, the largest unsigned differences in relative energies are 2.4 kcal mol−1 and 1.0 kcal mol−1, respectively, for B3LYP/MM and B3LYP-D3/MM optimized structures, with average unsigned differences being 0.8 kcal mol−1 and 0.5 kcal mol−1. Therefore, if the PESs and the optimized structures are similar, MP2 single point energy calculations can be used to obtain a quick estimation of the effects of many-body dispersion energy on the catalytic pathway. However, if some minimum or maximum points are missing, this strategy cannot be used.
For the catalytic pathway without K+ at site 1, we examined the differences in the geometries and energies optimized with the B3LYP-D3/MM and MP2/MM methods. The results are shown in Table 1 and Table S5 (ESI†). Again, the relative energies for TS1 and TS2 (19.9 kcal mol−1 and 24.4 kcal mol−1) obtained from B3LYP-D3/MM geometry optimizations are significantly higher than those from MP2/MM (12.6 and 17.4 kcal mol−1). For geometries, most of the distances optimized with B3LYP-D3/MM are slightly longer than those optimized with MP2/MM (Table S5, ESI†).
We also compared the structures and energies obtained with the MP2/MM, B3LYP/MM, and B3LYP-D3/MM methods. It is found that the B3LYP/MM method predicts too high transition state energies. Including Grimme's empirical dispersion correction to the B3LYP can lower the reaction barriers by ∼4 kcal mol−1, highlighting the importance of active site intermolecular dispersion interactions in enzyme catalysis. Even with the empirical dispersion correction (i.e., B3LYP-D3/MM), the computed reaction energy barriers for the first and second steps are still ∼5 kcal mol−1 higher than those obtained with the MP2/MM method. Although both B3LYP/MM and B3LYP-D3/MM predict the same third transition state (TS3) as MP2/MM does, the B3LYP/MM and B3LYP-D3/MM predicted electronic energy barriers from EI2 to TS3 are too low (1.7 kcal mol−1 and 2.9 kcal mol−1, respectively). With ZPE and thermal free energy corrections, the activation free energies are even lower or negative on the B3LYP/MM and B3LYP-D3/MM potential energy surfaces, reducing the three-step mechanism to a two-step mechanism. In contrast, MP2/MM predicts a significant electronic energy barrier of 6.6 kcal mol−1 from EI2 to TS3, solidifying a three-step mechanism for systems with a K+ ion at site 1. When a sufficiently large basis set such as aug-cc-pVDZ is used, the M05-2X, M06, M06-2X, and MN15 functionals can generate results closer to MP2. However, when smaller basis sets such as 6-31G* are used, the errors can be as large as 10 kcal mol−1 in the computed relative energies, causing incorrect perspectives of the HDAC8 catalytic mechanism. Caution must be used when choosing the QM method and basis set in QM/MM study of enzyme catalysis.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5cp00002e |
This journal is © the Owner Societies 2025 |