A.
Tomić
a,
B.
Kovačević
*b and
S.
Tomić
*a
aDepartment of Organic Chemistry and Biochemistry, Rudjer Boskovic Institute, Bijenicka cesta 54, 10000, Zagreb, Croatia. E-mail: Antonija.Tomic@irb.hr; Sanja.Tomic@irb.hr
bDepartment of Physical Chemistry, Rudjer Boskovic Institute, Bijenicka cesta 54, 10000, Zagreb, Croatia. E-mail: Borislav.Kovacevic@irb.hr
First published on 2nd September 2016
Human dipeptidyl-peptidase III (h.DPP III) is a zinc-exopeptidase that hydrolyses dipeptides from the N-terminus of its substrates. Its mechanism of action was assumed to be similar to that of thermolysin, but was never thoroughly investigated. This study presents the first insight into the reaction mechanism of h.DPP III, determined on the model and real (hydrated enzyme with Leu-enkephalin bound in the active site) systems. The Glu451-assisted water addition on amide carbon atoms and nitrogen inversion (i.e. change of pyramidalization on the leaving nitrogen) are shown to be the rate-determining steps with the activation energies in a good agreement with the experimental results for the Leu-enkephalin hydrolysis. The energy barrier for nucleophilic attack is about 28 kJ mol−1, while barriers for the N-inversion differ as a consequence of the number of hydrogen bonds that have to be changed, which is smaller in the model active site than in the solvated enzyme. Although precisely defined geometry of the enzyme binding site puts an additional restraint on the hydrogen bonding interactions, at the same time it stimulates the forward reaction towards the final hydrolytic product. Namely, different from the model, the N-inversion is in a concerted fashion followed by favourable hydrogen bonding with Glu451 that immediately “locks” the system into the configuration where reversion to the enzyme–substrate complex is hardly achievable. Therefore we propose that the functional significance of DPP III is dual: to lower the energy barrier of the peptide hydrolysis and to suppress the reverse reaction.
In mammalian tissues, DPP III is broadly distributed and thought to contribute to the final steps of normal intracellular protein catabolism. There are strong indications of its role in the endogenous pain-modulation system,3,4 as well as in the endogenous defense against oxidative stress.5 Pathophysiological roles of DPP III are indicated in cataractogenesis,6 malignant growth7 and influenza virus infection.8 The most recent studies showed that DPP III is a member of a six-gene signature that accurately predicts human breast cancer patient survival.9 Hast et al. revealed that DPP III contributes to the activation of transcription factor Nrf2 by binding to its negative regulator, KEAP1, a member of the Keap1 functional E3 ubiquitin ligase complex.10 The Nrf2-Keap1 signalling pathway is a major regulator of the cytoprotective response to oxidative and electrophilic stress. All these findings characterize DPP III as a valuable drug target.
There are several crystal structures of h.DPP III available. The structure of the ligand-free enzyme (pdb codes: 3FVY and 3EGY) was determined in the so-called open and closed form, while that of the liganded enzyme only in a more compact, closed, form (pdb codes: 3T6B, 3T6J, 5E33, 5E3A, 5E2Q, 5EHH and 5E3C). The structure of the open ligand-free enzyme has a wide cleft (∼40 Å wide and ∼40 Å high) between the two domains (so-called upper and lower domain) which can be easily accessed by peptide substrates. The increased protein globularity, which is a result of protein closure, has also been observed during molecular dynamic (MD) simulations of the ligand-free h.DPP III.11 During the simulations the enzyme transformed from an extended (open) form to the compact (closed) form while the conformation of each protein domain remained conserved.
In the closed (chemically active) enzyme structure the catalytic zinc ion, bound to the upper domain at the “roof” of the cleft, interacts with the carbonyl oxygen from the scissile amide bond of the substrate molecule bound to the lower protein domain.12–14 The substrate is bound in the form of a β-strand to the five-stranded β-core from a lower protein domain (of an enzyme) in an antiparallel fashion.
Although the DPP III structure represents a novel fold, the mode of zinc binding closely resembles that observed in other structurally unrelated metallopeptidases, such as neprilysin and thermolysin. These two enzymes also coordinate with the metal ion through two histidines and one glutamate, and superposition of the zinc-binding residues reveals very similar coordination geometry.15 Neprilysin and thermolysin have the structurally equivalent residues to evolutionarily 100% conserved His568 and Glu451 from the h.DPP III. It was shown that in the thermolysin catalyzed reaction the glutamic acid residue acts as a general base during the deprotonation of the water molecule that attacks the scissile amide bond, while the role of histidine is to stabilize the tetrahedral intermediate.16,17 The kinetic data on h.DPP III showed that the His568 to Asn mutation significantly decreases enzyme activity, for example the kcat value for hydrolysis of synthetic substrate Arg–Arg–2-naphthylamide (RR–2-NA) is three orders of magnitude lower in the His568Asn mutant than in the wild type protein, however the Km values are similar.18 In addition, the Tyr318 to Phe substitution (Tyr318 is also 100% conserved residue) resulted in a significant reduction of the enzymatic activity, the kcat for RR–2-NA hydrolysis decreased two orders of magnitude while the ligand binding affinity was not affected.19 These results indicate that these highly conserved residues could be involved in transition state stabilization during the catalytic action of h.DPP III.
We approached the study of human DPP III hydrolysis progressively. First we performed a quantum mechanical (QM) study on a small model system shown in Fig. 1 that was followed by a quantum mechanics/molecular mechanics (QM/MM) study of the reaction in the solvated enzyme – Leu-enkephalin (Tyr–Gly–Gly–Phe–Leu) complex. Using DFT calculations we elucidated the role of the individual enzyme residues in the catalytic reaction and determined the protonation state of the 100% conserved His568 residue.
QM calculations were carried out for a minimal model of the active site consisting of the zinc ion ligated by one water molecule, two 4-methylimidazoles, and one propanoate group, replacing the enzyme residues His450, His455 and Glu508, respectively. The conserved residues Tyr318 and Glu451 were modelled by 1-methly-4-hydroxyphenyl and butyrate groups, while His568 by 4-methylimidazol or 4-methylimidazolium cations, in the case of the neutral (histidine with protons on the Nε atom) or protonated histidine residue, respectively (Fig. 1). The substrate was modelled by CH3CO–Gly–NHCH3, i.e. glycine was capped with acetyl (Ace) at the N-terminal and methylamide (Mam) at the C-terminal (Ace–Gly–Mam). The initial position of the water molecule coordinating the zinc ion was chosen based on our previous QM/MM calculations on the zinc ion coordination.12 Since the previous theoretical studies of the peptide hydrolysis by thermolysin had revealed the Glu143-assisted water addition as a preferable mechanism, the water molecule attached to Zn2+ was hydrogen bonded to Glu451 in our model system as well. This is in line with our QM/MM calculations on the Zn2+ ion coordination12 which clearly indicated that one water molecule is always hydrogen bonded to Glu451 in the closed structure of h.DPP III. Another water molecule found close to the metal ion12 coincides with the substrate oxygen location and was deleted.
In the crystal structures of DPP III–peptide complexes, the amide group of the second amino acid from the substrate N-terminus makes a hydrogen bond with Tyr318, while its hydroxyl group is hydrogen bonded to His568. The particular hydrogen-bond pattern was preserved in our model structure (see Fig. S1, ESI†).
The B97D20 density functional method as implemented in GAUSSIAN 0921 program package was used for all the present optimization calculations. Stationary points were determined using the 6-31G(d) basis set for the H, N, C and O atom and the LANL2DZ-ECP basis set for the Zn2+ ion.22 During the optimization of the stationary points (ES, INT and P) the terminal methyl group carbon atoms of the model enzyme residues (Y318, H450, E451, H455, E508 and H568) were fixed. However, to search for the minima and transition states, the unconstrained relaxed potential energy surface scan was used. To guess the transition state structure the Ow–Cs and Cs–Ns distances were scanned since these bond lengths most closely describe the imaginary frequencies of the first and second transition states proposed in the study of thermolysin hydrolysis.17 No constraints were applied on the substrate during the optimization. The transition state structures were obtained either using the Synchronous Transit-Guided Quasi-Newton (STQN) method or the transition state optimization method.23–25
Each transition state was confirmed by the existence of one negative eigenvalue (one imaginary frequency), with the vibrational vector leading to the desired neighbouring minima. In addition, transition state geometries were connected to the ground state geometries using intrinsic reaction coordinate (IRC) calculations that allow one to follow the path downhill from the transition state in both forward and reverse directions.26
Calculations were performed in the presence of the low dielectric media (dielectric constant 4) utilizing the polarizable continuum model (CPCM).27,28
QM/MM geometry optimization of the enzyme binding site was performed using the 2-layer ONIOM (Our own N-layered Integrated molecular Orbital and molecular Mechanics) methodology implemented in program GAUSSIAN 09:
EONIOM = EMM(S) + EQM(SM) − EMM(SM) | (1) |
For the purpose of 2-layer ONIOM structure optimization the system was divided into two layers which were handled at different levels of theory. At the places where borders “cut” covalent bonds (Cα–Cβ), the link atoms were placed. The QM layer, consisting of Y318, H450, E451, H455, E508 and H568 side chains, Leu-enkephalin, the zinc ion and a water molecule, was treated by the B97D DFT method and two different basis sets: 6-31G(d) for the H, N, C and O atoms, and LANL2DZ-ECP for the Zn2+ atom. The MM part (the rest of the protein) was treated using the AMBER force field (parm96).31 To take into account solvation, water molecules from the enzyme first and second solvatation sphere (additional 3501 water molecules closest to the protein molecule) were considered as a part of the MM layer as well. The initial cycle of geometry optimizations was performed using mechanical embedding (ME), and the final geometry optimization was performed using electronic embedding (EE).30 The net charge of the whole system was either −24 e or −23 e, while the QM region had a zero or plus one charge depending on whether the H568 is neutral or protonated, respectively. During QM/MM geometry optimization and potential energy surface scan, protein residues and water molecules found within 8 Å from the substrate molecule were allowed to move. Minima and saddle points were designated according to the vibrational analysis. Transition states were confirmed by the existence of one negative eigenvalue (one imaginary frequency), with the vibrational vector leading to the desired neighbouring minima. In addition, each transition state was connected to the neighbouring ground states by IRC calculations.
Details of the amide bond hydrolysis mechanism determined using the model system with protonated His568 are summarized in Fig. 2, with the energy profile for individual reaction steps illustrated in Fig. 3a and Table 2. The three-dimensional structures of stationary points are shown in Fig. S2 (ESI†). The selected bond lengths in the optimized structures are given in Table 1.
ES | TS1 | INT1 | TS2 | INT2 | TS3 | INT3 | TS4 | P | P′ | |
---|---|---|---|---|---|---|---|---|---|---|
a Distances between the zinc ion and either nitrogen (Nδ) or oxygen (carboxyl) atoms from the histidine or glutamate amino acid residue, respectively. b Distance between the Tyr318 hydroxyl oxygen and amide hydrogen of the second amino acid residue from the substrate (sub) N terminus. c C and C′ are carbon atoms adjacent to Ns and Cs, respectively, while Hs is a hydrogen bonded to Ns, see Fig. S3 (ESI). | ||||||||||
d(Os–Zn)/Å | 2.42 | 2.09 | 2.03 | 2.04 | 2.02 | 2.05 | 2.06 | 2.07 | 2.12 | 4.17 |
d(Ow–Zn)/Å | 2.06 | 2.18 | 2.36 | 2.26 | 2.40 | 2.26 | 2.25 | 2.26 | 2.36 | 2.01 |
d(Cs–Ns)/Å | 1.34 | 1.39 | 1.43 | 1.40 | 1.45 | 1.46 | 1.50 | 1.55 | 2.76 | 4.15 |
d(Ow–Cs)/Å | 3.23 | 1.79 | 1.55 | 1.60 | 1.49 | 1.48 | 1.44 | 1.42 | 1.27 | 1.28 |
d(Os–Cs)/Å | 1.26 | 1.32 | 1.36 | 1.36 | 1.38 | 1.38 | 1.37 | 1.35 | 1.30 | 1.27 |
d(Ow–Hw2)/Å | 0.98 | 0.99 | 1.00 | 1.00 | 1.00 | 0.99 | 1.00 | 1.02 | 1.67 | 3.66 |
d(Ow–Hw1)/Å | 1.09 | 1.58 | 1.67 | 1.65 | 1.71 | 2.11 | 2.60 | 2.49 | 2.69 | 5.05 |
d(Hw2–Oe2)/Å | 2.82 | 2.83 | 2.84 | 2.76 | 2.74 | 2.50 | 1.71 | 1.60 | 1.01 | 1.06 |
d(Hw1–Oe1)/Å | 1.41 | 1.03 | 1.01 | 1.02 | 1.01 | 0.99 | 1.06 | 1.24 | 2.22 | 2.82 |
d(Ns–Hw1)/Å | 5.07 | 2.95 | 2.96 | 2.72 | 2.76 | 2.19 | 1.61 | 1.30 | 1.02 | 1.02 |
d(Ns–Hw2)/Å | 4.66 | 3.19 | 3.10 | 3.07 | 3.01 | 2.92 | 2.59 | 2.60 | 3.59 | 1.59 |
d(H450–Zn)a/Å | 2.21 | 2.17 | 2.15 | 2.15 | 2.14 | 2.14 | 2.16 | 2.15 | 2.11 | 2.13 |
d(H455–Zn)a/Å | 2.15 | 2.11 | 2.11 | 2.11 | 2.11 | 2.11 | 2.12 | 2.12 | 2.11 | 2.11 |
d(E508–Zn)a/Å | 2.11 | 2.09 | 2.08 | 2.08 | 2.07 | 2.06 | 2.05 | 2.05 | 2.04 | 2.05 |
d(H568[Hε]–Os)/Å | 1.68 | 1.53 | 1.45 | 1.46 | 1.42 | 1.43 | 1.49 | 1.53 | 1.66 | 1.60 |
d(H568[Nε–Hε])/Å | 1.05 | 1.08 | 1.10 | 1.10 | 1.12 | 1.11 | 1.09 | 1.08 | 1.05 | 1.08 |
d(Y318–sup)b/Å | 1.82 | 1.84 | 1.83 | 1.84 | 1.84 | 1.87 | 1.95 | 1.95 | 1.94 | 1.96 |
ω 1(C–Ns–Cs–Hs)c/° | 175.6 | −133.5 | −125.6 | 163.9 | 122.1 | 124.9 | 122.7 | 121.5 | ||
ω 2(C′–Cs–Ns–Os)c/° | −178.6 | 137.6 | 126.5 | 131.1 | 126.3 | 127.1 | 125.9 | 125.0 |
The zinc-coordinated water is polarized by a negatively charged Glu451 (strong Oe1–Hw1–Ow hydrogen bond), which is a strong base, and by the zinc cation (2.06 Å long Zn–Ow coordination bond). Consequently, the Hw1 proton is partially shared between Glu451 and water as indicated by Oe1–Hw1 and Ow–Hw1 distances of 1.41 Å and 1.09 Å, respectively. Apparently the pKa value of the metal bound water molecule is lower than that of the one in solution. It should be noticed that the water Hw2 proton makes an additional hydrogen bond with the substrate terminal acetyl group. Besides by its coordination to the metal ion, the substrate is further stabilized by hydrogen bonds with evolutionally conserved His568 and Tyr318. Tyr318 is additionally hydrogen bonded to Glu508 and this hydrogen bond is preserved throughout the reaction. A similar substrate-stabilization pattern is observed in the active site of thermolysin.16
Once the tetrahedral intermediate is formed (INT1), it can either lose the –NH2CH3 group (giving the hydrolytic product) or the –OH group (reverting to the starting compound). Deslongchamps32,33 has proposed that one of the factors affecting this choice is the conformation of the intermediate; more specifically, the position of the lone pairs. In this view, a leaving group –NH2CH3 or –OH can depart only if each of the remaining heteroatoms bound to the carbon (Cs) atom has an orbital antiperiplanar to the breaking bond, either Cs–Ns or Cs–Ow. This factor was called the stereoelectronic control or effect. From the principle of microscopic reversibility it follows that the lone-pair orbitals on Os and Ns atoms upon OH− nucleophilic attack must be oriented antiperiplanar to the new Cs–Ow bond, i.e. initial pyramidalization on the Ns atom (INT1 structure) with the lone pair pointing away from the Glu451 is influenced by the direction of OH− attack.
The N-inversion from the (R) to (S) isomer through the near planar transition state (ω1 = 163.9° in TS2 structure) occurs in the second reaction step with an activation barrier of 11.6 kJ mol−1. The improper torsion angle ω1, defining the degree of nitrogen pyramidalization, changes by more than 110° as the system evolves from the INT1 to INT2 structure. It should be noticed that the structures INT1 and INT2 differ only in the lone pair position at Ns, namely, the N-inversion (the change in the Ns configuration from (R) to (S)) leads to the new tetrahedral intermediate in which the positions of the N–H bond and the lone pair orbital are interchanged. According to the energy profile, the (R) isomer (INT1) is about 8 kJ mol−1 more stable than the (S) isomer (INT2). Apparently the nitrogen in the INT2 structure can easily surmount the low energy barrier (3.7 kJ mol−1) and interconvert back to the (R) isomer (INT2 → INT1 transition). Instead, the formation of the hydrogen bond between the Ns lone pair and the Glu451 Ow–Hw1 group “locks” Ns in (S) configuration (INT2 → INT3 transition). The energy barrier for this transition is 5.9 kJ mol−1. Once the Ns–Glu451 hydrogen bond is formed, the reverse reaction is hindered. Namely, the energy barrier of 13.5 kJ mol−1 is much higher for the reverse (INT3 → INT2 transition) than for the forward (INT3 → P transition) reaction which is barrierless (Fig. 3a). Furthermore, the lone pair orbitals at Ns and Os are not antiperiplanar to the Cs–Ow bond, and therefore, according to the stereoelectronic effect, this reaction is not feasible. In summary, for the reaction determined using the model system, TS3 is the point from which the reaction can proceed only in the forward direction, i.e. from which the reverse reaction becomes unfavourable. The first three reaction steps can be considered as rate-determining steps with a total activation energy of 45.5 kJ mol−1 (obtained as a sum of 28 kJ mol−1 + 11.6 kJ mol−1 + 5.9 kJ mol−1). During the third reaction step (INT2 → INT3 transition) the hydroxyl group (Ow–Hw2) bound to the substrate reoriented in a way to form a hydrogen bond with the Glu451 (Ow–Hw2–Oe2 hydrogen bond in the INT3 structure). Only at this point, after establishing strong Ow–Hw2–Oe2 and Ns–Hw1–Oe1 hydrogen bonds, does the (S) isomer in the INT3 structure become equally stable as the (R) isomer in the INT1 structure.
In INT1-3 structures, the amide bond has a single bond character as indicated by the Cs–Ns distance (1.43, 1.45 and 1.50 Å). Oxyanion intermediates (INT1-3) and transition states (TS2-3) are stabilized by the strong hydrogen bonding to His568 and Glu451 (Ow–Hw1–Oe1 hydrogen bond in INT1 and INT2 structures, and Ns–Hw1–Oe1 hydrogen bond in the INT3 structure), as well as by bidentate coordination to the zinc cation. Strong hydrogen bonding between the substrate and His568 is indicated by short Hε–Os distances (1.45, 1.42 and 1.49 Å in INT1, INT2 and INT3 structures, respectively), and consequently slightly extended His568 Nε to Hε atom distances (1.10, 1.12 and 1.09 Å in INT1, INT2 and INT3 structures, respectively).
Therefore, the fourth reaction step, during which the system relaxes into the final product (P), consists of: transfer of the proton Hw1 from Glu451 to leaving nitrogen (Ns), followed by cleavage of the amide bond and regeneration of the protonated state of the carboxyl of Glu451 by the second Hw2 proton from the Ow atom. One can notice that upon making a zero-point correction to electronic energy, the energy of TS4 structures becomes lower than the energy of the INT3 structure suggesting that the system more or less continuously, without energy barrier, relaxes from the INT3 structure into the product P.
The final optimized product complex P has a bidentate bond between the zinc and the acetate part of the product of the hydrolysis (although the Ow–Zn coordination bond of 2.36 Å is slightly long). In addition, the acetate part of the product is stabilized by the strong hydrogen bond with His568. The methylamine part of the product is only weakly bound to the model of the enzyme active site via a single hydrogen bond to the Glu451 Oe1 oxygen (with the Oe1–Hw1 distance equal to 2.22 Å) and weak van der Waals interactions with His450 and His568.
As suggested for thermolysin17 the second proton (Hw2) transfer by Glu451 towards the Ns nitrogen could probably be considered as a next step of the product conversion, although this step is not crucial for peptide bond cleavage. We were not able to locate protonated amine as a final product. Instead, the product P′, where the Hw1 proton is placed on Glu451, and where Zn2+ is tetrahedrally coordinated (with one coordination to acetate part of the product), was found to be the most stable product structure.
The mechanism for the peptide bond hydrolysis obtained for the hydrated enzyme is summarized in Fig. 5, with the energy profile for individual reaction steps given in Fig. 3b and Table 2. Three-dimensional structures of the energy optimized stationary points are shown in Fig. S4 (ESI†), and the selected bond lengths in these structures are given in Table 3.
ΔE (model)/kJ mol−1 | ΔE (real)/kJ mol−1 | ||||
---|---|---|---|---|---|
A | B | A | B | ||
ES | 0.00 | 0.00 | ES | 0.00 | 0.00 |
TS1 | 26.38 | 27.99 | TS1 | 24.16 | 27.94 |
INT1 | 23.19 | 26.68 | INT1 | 23.11 | 30.25 |
TS2 | 39.66 | 38.30 | TS2 | 61.72 | 59.48 |
INT2 | 31.24 | 34.62 | INT2 | 44.65 | 44.22 |
TS3 | 38.74 | 40.52 | TS3 | 45.44 | 38.39 |
INT3 | 24.32 | 26.96 | INT3 | 37.82 | 35.44 |
TS4 | 28.64 | 22.97 | TS4 | 38.35 | 30.80 |
P | 3.13 | 3.32 | P | −12.61 | 3.71 |
P′ | −38.51 | −38.94 |
ES | TS1 | INT1 | TS2 | INT2 | TS3 | INT3 | TS4 | Pd | P′d | |
---|---|---|---|---|---|---|---|---|---|---|
a Distances between the zinc ion and either nitrogen (Nδ) or oxygen (carboxyl) atoms from the histidine or glutamate amino acid residue, respectively. b Distance between the Tyr318 hydroxyl oxygen and amide hydrogen of the second amino acid residue from the substrate (sub) N terminus. c C and C′ are carbon atoms adjacent to Ns and Cs, respectively, while Hs is a hydrogen bonded to Ns, see Fig. S3 (ESI). d Corresponds to the enzyme-product complex. | ||||||||||
d(Os–Zn)/Å | 2.89 | 2.17 | 2.09 | 2.09 | 2.10 | 2.11 | 2.15 | 2.13 | 2.85 | |
d(Ow–Zn)/Å | 2.02 | 2.15 | 2.26 | 2.17 | 2.16 | 2.16 | 2.14 | 2.15 | 2.03 | |
d(Cs–Ns)/Å | 1.36 | 1.42 | 1.44 | 1.45 | 1.53 | 1.57 | 1.69 | 1.74 | 3.18 | |
d(Ow–Cs)/Å | 2.56 | 1.74 | 1.58 | 1.52 | 1.46 | 1.45 | 1.41 | 1.39 | 1.31 | |
d(Os–Cs)/Å | 1.24 | 1.31 | 1.33 | 1.34 | 1.34 | 1.33 | 1.32 | 1.31 | 1.24 | |
d(Ow–Hw2)/Å | 0.99 | 0.99 | 0.99 | 0.98 | 1.00 | 1.01 | 1.07 | 1.14 | 1.69 | |
d(Ow–Hw1)/Å | 1.07 | 1.56 | 1.61 | 2.10 | 2.53 | 2.48 | 2.41 | 2.42 | 3.58 | |
d(Hw2–Oe2)/Å | 1.88 | 2.58 | 2.63 | 2.45 | 1.71 | 1.64 | 1.45 | 1.31 | 1.01 | |
d(Hw1–Oe1)/Å | 1.45 | 1.03 | 1.02 | 1.00 | 1.08 | 1.18 | 1.57 | 1.64 | 3.56 | |
d(Ns–Hw1)/Å | 2.78 | 2.83 | 2.80 | 1.95 | 1.53 | 1.34 | 1.08 | 1.06 | 1.03 | |
d(Ns–Hw2)/Å | 3.87 | 3.21 | 3.12 | 2.92 | 2.54 | 2.55 | 2.64 | 2.72 | 4.46 | |
d(H450–Zn)a/Å | 2.12 | 2.15 | 2.14 | 2.13 | 2.12 | 2.12 | 2.11 | 2.12 | 2.10 | |
d(H455–Zn)a/Å | 2.12 | 2.15 | 2.15 | 2.14 | 2.14 | 2.14 | 2.13 | 2.14 | 2.14 | |
d(E508–Zn)a/Å | 2.04 | 2.08 | 2.08 | 2.07 | 2.07 | 2.07 | 2.07 | 2.08 | 2.04 | |
d(H568[Hε]–Os)/Å | 1.78 | 1.60 | 1.56 | 1.50 | 1.55 | 1.57 | 1.61 | 1.61 | 1.81 | |
d(H568[Nε–Hε])/Å | 1.04 | 1.07 | 1.08 | 1.10 | 1.09 | 1.08 | 1.07 | 1.07 | 1.04 | |
d(Y318–sub)b/Å | 2.30 | 2.29 | 2.30 | 2.27 | 2.15 | 2.15 | 2.20 | 2.21 | 2.47 | |
ω 1(C–Ns–Cs–Hs)c/° | −151.5 | −132.0 | −128.3 | 158.1 | 130.1 | 127.9 | 128.0 | 127.9 | ||
ω 2(C′–Cs–Ns–Os)c/° | 173.7 | 138.6 | 131.2 | 132.0 | 129.8 | 128.9 | 126.4 | 125.0 |
Antiparallel binding of the β-stranded substrate to the protein β-core influences mild initial pyramidalization on the Ns atom (ω1 = −151.5°) hydrogen bonded to Gly389.
Nucleophilic attack of OH− on Cs is accompanied by pyramidalization on the carbon (Cs), see Table 2, resulting not only in the formation of the tetrahedral intermediate INT1, but also in the further pyramidalization of Ns. Namely, during the system evolution from ES to INT1, ω2 and ω1 dihedral angles reduce more than 40° and 20°, respectively. Due to the nitrogen Ns atom pyramidalization its lone pair electrons orient away from Glu451 in ES, TS1 and INT1 structures. As already emphasized in the discussion about the reaction obtained using the model system, the (R) isomer of Ns is stereoelectronically controlled by the direction of OH− attack. In the enzyme this control is further supported by the favourable hydrogen bonding to the protein, namely the Gly389 carbonyl group.
By adding the zero-point correction to electronic energy, TS1 becomes lower in energy than INT1 (Fig. 3b) suggesting that at this level of theory the TS1 structure cannot be considered as a real transition state.
The energy cost for the system to have a favourable (S) isomer on the leaving nitrogen atom is higher in the enzyme (59.5 kJ mol−1) than in the model system (45.5 kJ mol−1). This is a result of the N-inversion step (as energies for the water addition step are equal) that costs 29.2 kJ mol−1 in the enzyme active site, while 17.5 kJ mol−1 in the model system. A difference of ∼12 kJ mol−1 is mostly an outcome of more hydrogen bonds being broken in the enzyme active site than in the model during transformation of the (R to S) isomer. In both real and model systems, by interconversion to the favourable (S) isomer on the Ns atom, two strong, Ow–Hw2–Oe2 and Ns–Hw1–Oe1, hydrogen bonds are being formed, while the one between the added Ow–Hw2 group and the substrate (carbonyl group from the first amino acid or the terminal acetyl group) disappears. Meanwhile in the enzyme active site transformation from the INT1 to INT2 structure leads to weakening of the hydrogen bond between the leaving amide group and Gly389, as well as the total loss of the hydrogen bond between the lone pair orbital on the leaving Ns atom and the water molecule found in its vicinity (see Fig. 6).
From the energetic point of view we can conclude that once the system passes the TS2 transition state it more or less continuously develops towards the final product P, i.e. once the configuration of the leaving nitrogen is suitable for proton transfer, the system relaxes in the final product without passing any energy barrier. This shows that forward reaction, leading to a final product, is even more preferred in real than in model systems, i.e. that reversible reaction is more difficult to appear in the enzyme active site than in a model system.
Calculations have shown that emerging negative charge on the Os atom in the rate-determine steps is mostly stabilized by coordination to the metal dication and by the strong hydrogen bonding with protonated His568 as indicated by the short His568–Os distances. This is in agreement with the experimental measurement that has shown a 1296-fold decrease in the kcat value after mutation of His568 to Asn. Kinetic data also suggested the importance of Tyr318 in enzymatic hydrolysis, i.e. Tyr318 to Phe mutation resulted in a 100-fold decrease in the kcat value. The presented results clearly indicate stabilization of the transition state by His568, and participation of Tyr318 in positioning of both the substrate and Glu508.
According to the calculations the reaction rate is determined by two consecutive steps, the nucleophilic attack and the nitrogen (Ns) inversion. While the nucleophilic attack (by Glu451-assisted deprotonated water molecule) energy barriers determined on the systems of different complexities are of the same height (28 kJ mol−1), the energy barriers for the N-inversion differ by 12 kJ mol−1. The reason for this difference is the larger number of hydrogen bonds that need to be broken during the N-inversion in the real system than in the model system. In order to properly position toward the Glu451 hydroxyl (Oe1–Hw1) and to enable the proton (Hw1) transfer, the scissile peptide bond nitrogen (Ns) has to interconvert. While unconstrained in the model system, this interconversion is hindered by two hydrogen bonds, Ns–Hs to Gly389 and Ns to H2O, in the enzyme. This is probably the reason why in a real system, the N-inversion step occurs in a concerted fashion followed by the hydrogen bonding to Glu451, and consequently why the interconversion immediately “locks” the system in configuration from which the backward reaction in more difficult. From the energetic point of view, once the system passes the TS2 transition state it almost barrierlessly develops towards the final product P. On the other hand, in the model system, one can observe equilibrium between the (R) and (S) isomers on the Ns atom, i.e. the substrate is more free to explore different conformations. To be more precise, from the INT2 structure, the model system can equally easy revert to ES or go forward to P passing over energy barriers of 3.7 and 5.9 kJ mol−1, respectively. Therefore, a greater possibility of reversible reaction leading to a starting compound is observed in the model system than in the real system.
These findings can also be explained by stereoelectronic theory. In the model system, in the INT2 structure, the NS atom has a favourable configuration for a Hw1 proton transfer (the lone pair on the Ns atom is pointing toward Glu451, although the Ns–Hw1 distance of 2.76 Å is still quite long for proton transfer to occur) but there is no lone pair orbital on the Ow atom antiperiplanar to the C–N bond to allow its cleavage. So, from this point the system can easily transform into an ES complex. Only after the Ow–Hw2–Oe2 hydrogen bond is formed, orientation of the lone pair on the Ow atom becomes antiperiplanar to the C–N bond, and the Ns–Hw1 distance reduces to 1.61 Å. This transformation “locks” the system in a configuration where the reversion to the enzyme–substrate complex is not favourable. Moreover, the lone pair orbitals antiperiplanar to the newly formed Cs–Ow bond (needed for the reversible reaction) do not exist at Ns and Os atoms. Therefore only after the favourable hydrogen bond interactions between the substrate and Glu451 are established, the equilibrium between two tetrahedral intermediates is likely to be largely on the side of the (S) isomer than the (R) isomer in the model system.
On the other hand in the enzyme active site, water addition and reorientation of the lone pair orbitals on Ns (N-inversion) and Ow atoms occur in a concerted fashion. Once the system is transformed into the favourable (S) configuration at the nitrogen, favourable hydrogen bonds with the Glu451 are formed that according to the stereoelectronic principle allow the C–N bond breakage. Thus one can say that the functional significance of the enzyme lies in the protein-controlled suppression of the reverse reaction.
Although most of the reaction steps that define the mechanism of action of thermolysin and neprilysin are similar to the ones determined for human DPP III, the importance of the nitrogen inversion was not reported in their case. However, we must emphasize that the importance of the N-inversion step was already found in catalytic pathways of aspartic peptidases.38
The QM/MM calculations on the enzyme–substrate complex carried out in this work indicate that the substrate carbonyl oxygen atom is not coordinated to the zinc ion in the Michaelis complex. Namely, the Zn2+ ion is tetrahedrally coordinated by His450, His455, Glu508 and water molecules. In the model system, the zinc coordination number is five since the substrate molecule also coordinates with the zinc ion, although rather weakly. Stabilization of the oxy-anion transition state is achieved through the formation of the strong coordination bond between the substrate carbonyl oxygen and the zinc ion as well as by its hydrogen bonding to the protonated His568. This suggests that both the transition state for nucleophilic attack and the oxy-anion intermediate are more tightly bound to the enzyme than the substrate itself.
Kinetic data obtained for the Leu-enkephalin hydrolysis by human DPP III are in a good agreement with our QM/MM calculations. Moreover, those obtained for His568 and Tyr318 DPP III mutants can be well explained by the results presented in this paper as well. A 1296-fold decrease in the kcat value after mutation of His568 to Asn is a consequence of direct stabilization of the transition state by histidine, while a 100-fold decrease in the kcat value after mutation of Tyr318 to Phe shows participation of tyrosine in positioning of both the substrate and Glu508.
Footnote |
† Electronic supplementary information (ESI) available: QM and QM/MM calculations of the systems with neutral His568, and QM/MM calculations of the system obtained directly from MD simulations. See DOI: 10.1039/c6cp04580d |
This journal is © the Owner Societies 2016 |