P.
Paiva
ab,
L. M. C.
Teixeira
ab,
R.
Wei
c,
W.
Liu
d,
G.
Weber
e,
J. P.
Morth
bf,
P.
Westh
bf,
A. R.
Petersen
bg,
M. B.
Johansen
bg,
A.
Sommerfeldt
bg,
A.
Sandahl
bg,
D. E.
Otzen
bh,
P. A.
Fernandes
ab and
M. J.
Ramos
*ab
aLAQV@REQUIMTE, Departamento de Química e Bioquímica, Faculdade de Ciências, Universidade do Porto, Rua do Campo Alegre s/n, 4169-007, Porto, Portugal. E-mail: mjramos@fc.up.pt
bEnZync Center for Enzymatic Deconstruction of Thermoset Plastics,
cJunior Research Group Plastic Biodegradation, Department of Biotechnology & Enzyme Catalysis, Institute of Biochemistry, University of Greifswald, Felix-Hausdorff-Str. 8, 17489, Greifswald, Germany
dTianjin Institute of Industrial Biotechnology, Chinese Academy of Sciences, 32 West Seventh Avenue, Tianjin Airport Economic Area, Tianjin 300308, China
eMacromolecular Crystallography, Helmholtz-Zentrum Berlin, Alber-Einstein-Straße 15, 12489 Berlin, Germany
fDepartment of Biotechnology and Biomedicine, Technical University of Denmark, Søltofts Plads, DK-2800, Kongens Lyngby, Denmark
gTeknologisk Institut, Kongsvang Alle 29, DK-8000, Aarhus, Denmark
hInterdisciplinary Nanoscience Center (iNANO). Aarhus University, Gustav Wieds Vej 14, DK-8000 Aarhus, Denmark
First published on 18th December 2024
The recently discovered metagenomic urethanases UMG-SP1, UMG-SP2, and UMG-SP3 have emerged as promising tools to establish a bio-based recycling approach for polyurethane (PU) waste. These enzymes are capable of hydrolyzing urethane bonds in low molecular weight dicarbamates as well as in thermoplastic PU and the amide bond in polyamide employing a Ser-Sercis-Lys triad for catalysis, similar to members of the amidase signature protein superfamily. Understanding the catalytic mechanism of these urethanases is crucial for enhancing their enzymatic activity and improving PU bio-recycling processes. In this study, we employed hybrid quantum mechanics/molecular mechanics methods to delve into the catalytic machinery of the UMG-SP2 urethanase in breaking down a model PU substrate. Our results indicate that the reaction proceeds in two stages: STAGE 1 – acylation, in which the enzyme becomes covalently bound to the PU substrate, releasing an alcohol-leaving group; STAGE 2 – deacylation, in which a catalytic water hydrolyzes the enzyme:ligand covalent adduct, releasing the product in the form of a highly unstable carbamic acid, expected to rapidly decompose into an amine and carbon dioxide. We found that STAGE 1 comprises the rate-limiting step of the overall reaction, consisting of the cleavage of the substrate's urethane bond by its ester moiety and the release of the alcohol-leaving group (overall Gibbs activation energy of 20.8 kcal mol−1). Lastly, we identified point mutations that are expected to enhance the enzyme's turnover for the hydrolysis of urethane bonds by stabilizing the macrodipole of the rate-limiting transition state. These findings expand our current knowledge of urethanases and homolog enzymes from the amidase signature superfamily, paving the way for future research on improving the enzymatic depolymerization of PU plastic materials.
Polyurethanes (PU) are among the most widely used types of recalcitrant plastics, due to their broad scope of properties. PU accounts for 5.3% of the 400 million tons of plastic produced globally each year, placing it sixth in terms of synthetic polymer production.3 The carbamate bond, commonly referred to as the urethane bond in polyurethane chemistry, R1OCOCNR2R3, is formed by combining an isocyanate (typically methylene diphenyl diisocyanate or toluene diisocyanate) with a polyol (e.g., polyethers and polyesters). Depending on the formulation, polyurethane materials can be either thermoplastic or thermoset polymers, allowing for a wide range of applications, including adhesives, coatings, foams, elastomers, and sealants.7
In 2018, hard and flexible foams constituted 68% of the PU market share, indicating that most of the applications used thermoset PU.8 Unlike thermoplastics, thermoset PUs have highly cross-linked structures and are thus insoluble in both water and organic solvents. Thermosets cannot be repeatedly melted and reshaped upon heating, but only thermally decomposed through processes like pyrolysis or chemical depolymerization at very high temperatures.7,9 Consequently, recycling thermoset PUs is limited to grinding, adhesive bonding, or chemical methods such as glycolysis. The first two are secondary recycling processes, which do not depolymerize the waste but simply repurpose the recycled polymer for less demanding applications. At the same time, the third option is a tertiary recycling method that molecularly disassembles the waste polymers and transform them to produce other chemicals.10 The current options for thermoset PU recycling are far from satisfactory, as the mechanically-recycled polymers can only be used for alternative applications (e.g., as fillers) with significantly reduced market value, or the chemical depolymerization process consumes high amounts of energy (ca. 817 kg CO2-eq. per t PU waste).11,12 Therefore, developing a more efficient and environmentally friendly recycling method is also economically beneficial.
Enzymatic depolymerization is one of the most promising strategies for addressing the end-of-life of recalcitrant hydrolyzable plastics.6,13,14 Unlike chemical recycling, the enzymatic process does not require harsh conditions (e.g., high temperatures and/or of toxic compounds). Still, it can suffer from low catalytic efficiency provided by the native biocatalysts. Since Müller et al. reported the first PET hydrolase (PETase) in 2005,15 an enzyme capable of depolymerizing polyethylene terephthalate (PET) into its monomers including terephthalate and ethylene glycol, there has been a rise in interest in discovery and design more powerful PETases.16–19 Building on this, the company Carbios engineered an enzyme that hydrolyzes around 90% of pretreated post-consumer PET waste within 10 hours, demonstrating the potential of enzymatic depolymerization for industrial recycling applications.20,21
The commercial success of enzymatic PET recycling has sparked renewed interest in biocatalytic recycling of other mass-produced plastics, with PU emerging as a logical next target due to its hydrolyzable backbone linked by carbamate bonds. The history of searching for PU-degrading microorganisms and enzymes is not necessarily shorter than that for polyesters; nonetheless, the results have been less promising, as most reported enzymes are polyester hydrolases that are exclusively active on polyester-based PU.7,22,23 In 2023, Branson et al.24 discovered three urethanase enzymes (UMG-SP1, UMG-SP2, and UMG-SP3), and demonstrated their effectiveness in PU recycling via a two-step chemoenzymatic process, achieving complete conversion of post-consumer soft foam waste composed of toluene diisocyanate (TDI)-based polyether-PU into the respective polyols and aromatic diamines. These enzymes were identified through a metagenomic screening from an isolated soil sample exposed to PU-related chemicals for a long period. Remarkably, UMG-SP2 hydrolyzed more than 90% of the low molecular weight dicarbamate TDI-diethylene glycol within 24 h.
The crystal structure of UMG-SP1, which shares 52.4% sequence identity with UMG-SP2, was recently solved along with a series of characterizations and engineering, demonstrating its depolymerization activity on pretreated polyamide and thermoplastic PU.25 Simultaneously, we elucidated ligand-free and ligand-bound crystal structures of UMG-SP2 and validated its depolymerization ability on PU polymer in a separate study.26 Consequently, our research established a robust foundation for enhancing UMG-SP2's catalytic efficiency to fulfill the demands of its applications in industrial PU recycling. To achieve this goal, understanding the structure-to-activity relationship will allow us to establish the catalytic mechanism of UMG-SP2 and subsequently identify “prejudicial residues” that destabilize the rate-limiting transition state (TS) in relation to the reactant state. Such undesirable residues increase the reaction activation energy and consequently decrease the rate, making them the most promising targets for mutations aimed at enhancing UMG-SP2's catalytic efficiency rationally.
UMG-SP2 belongs to the amidase signature superfamily, which has a highly conserved active site. Most members of this superfamily share the Sernuc-Sercis-Lys catalytic triad, including UMG-SP2 (Ser190nuc-Ser166cis-Lys91).27–29 In this type of catalytic triad, the mechanism typically starts with a proton transfer from Sercis to Lys, enabling the remaining Sernuc residue to perform a nucleophilic attack on the substrate's amide bond. For this to happen, the catalytic Lys must adopt a neutral state, as only in this state can it accept a proton donated by the Sercis residue. Later in the reaction, Lys returns the proton to Sercis, which in turn transfers its proton to the substrate's leaving group, culminating in the formation of the acyl-enzyme state and the release of an amine. To regenerate the enzyme, a water molecule enters the active site and conducts a nucleophilic attack on the substrate's carbonyl carbon that is bound to the Sernuc. Consequently, the Sernuc residue becomes deacylated, releasing the product as a carboxylic acid, and the enzyme regenerates for a new catalytic cycle.30 Even though the typical mechanism for Sernuc-Sercis-Lys has been studied before, the atomistic and energetic details vary throughout the family members. Additionally, to the best of our knowledge, the hydrolysis of the urethane bond catalyzed by this type of enzyme has not been elucidated. Thus, it is essential to study the catalytic mechanism of urethanases in detail, instead of relying on general studies for the amidase family. Therefore, we set out to establish the mechanism of the urethane bond hydrolysis catalyzed by UMG-SP2 in order to propose mutations to enhance its efficiency. For this purpose, we used the symmetric di-urethane ethylene 4,4′-methylenedianiline (DUE-MDA) as a substrate (Scheme 1).
![]() | ||
Scheme 1 The chemical structure of the substrate used in this work to mimic a PU segment, di-urethane ethylene 4,4′-methylenedianiline (DUE-MDA). |
This compound was chosen because it is a dicarbamate with two of the most often employed structural compositions (MDA as the isocyanate and polyether-based polyol) in industrial PU monomer formulations. We employed an adiabatic quantum mechanics/molecular mechanics methodology (QM/MM). Our calculations provided a detailed atomistic and energetic description of the mechanism and identified potential productive mutations to improve UMG-SP2's catalytic efficiency for urethane bond hydrolysis.
We estimated the protonation state of all residues, at pH 8.0, with the empirical pKa predictor PROPKA 3.5.0 (Table S1†).31 PROPKA predicted that the side chain of the catalytic triad Lys91 should exist mostly in its neutral form (–NH2). Visual inspection revealed that Lys91, buried in the active site cavity and therefore not solvated by water, was not close to any negatively charged residue (i.e., Asp or Glu) that could stabilize a hypothetical positive state of this residue, thus explaining why PROPKA estimated a pKa value of 6.09. In its neutral form, the side chain of Lys91 is able to donate two hydrogen bonds to two nearby serine residues (Ser167 and Ser185, located at ≈3.0–3.1 Å) and simultaneously accept a hydrogen bond from the catalytic Ser166cis (2.35 Å). We further analyzed visually the active site residues and other titratable residues in their surroundings to thoroughly verify the selected protonation states. Among the nine existing histidine residues, six were Nε-protonated (His159, His196, His215, His271, His376, and His439), and three were Nδ-protonated (His108, His111, and His159).
We employed the GROMACS software (version 2021.5)41,42 to minimize the energy of the assembled system via a three-step protocol that used the steepest descent algorithm.43 During the first minimization step, all water molecules were minimized. In the second step, the backbone atoms of UMG-SP2 were restrained while the remaining atoms of the system were minimized; in the final step, the whole system was minimized. The final minimized UMG-SP2 structure exhibited an RMSD of 0.38 Å (all enzyme non-hydrogen atoms were considered in the calculation) compared to the original crystallographic structure.
We then performed a classical molecular dynamics (MD) simulation to assess the overall stability of the UMG-SP2:DUE-MDA complex and to gather a structure that should correspond to a catalytically competent conformation. Throughout the entire MD protocol, all bonds involving hydrogen atoms were maintained fixed using the LINCS algorithm,44 which permitted the use of an integration time of 2 fs. The non-bonded interactions were explicitly calculated if under the cutoff of 10 Å, beyond which the Particle-Mesh Ewald scheme45 was applied to treat non-bonded Coulomb interactions. The system was initially heated to 29 °C for 100 ps at constant volume conditions (NVT ensemble), which was accomplished by using the V-rescale thermostat46 and by randomly generating initial velocities according to a Maxwell distribution. During this phase, all solute atoms were kept restrained. This heating phase was followed by a 2 ns-long NPT phase, in which the density of the solvent was relaxed at 1 bar and 29 °C, using the V-rescale thermostat and the Berendsen barostat.47 Again, all solute atoms were kept fixed while the solvent was allowed to equilibrate.
Subsequently, we performed a 100 ns-long NPT phase to equilibrate the overall structure of the UMG-SP2 enzyme at 29 °C and 1 bar, while preserving the geometry of the active site residues (i.e., Lys91, Ser166cis, Ile187, Gly188, and Ser190nuc) and the carbonyl group of the ligand's urethane bond with positional restraints. We followed this initial run with a second 100 ns-long NPT equilibration, which was conducted with fewer positional restraints (those affecting Ile187 and Gly188 were released). Finally, we carried out a 100 ns-long NPT production run without any restraints, controlling the temperature and pressure with the V-rescale thermostat and the Parrinello–Rahman barostat.48 During this stage, we saved the configurations of the system every 200 ps, and followed five distances throughout the production phase (Fig. S1†). We used them as selection criteria to choose a catalytically competent UMG-SP2:DUE-MDA conformation: Hγ(Ser190nuc)–Oγ(Ser166cis), Hγ(Ser166cis)–Nζ(Lys91), Oγ(Ser190nuc)–Ccarbonyl(DUE-MDA), Ocarbonyl(DUE-MDA)–NHbackbone(Ile187), and Ocarbonyl(DUE-MDA)–NHbackbone(Gly188). The first three distances are related to the putative activation/deprotonation of the catalytic Ser190nuc and the nucleophilic attack it should perform on the carbonyl carbon of the DUE-MDA's target urethane bond, respectively. The last two distances, calculated between two backbone amides and the ligand's carbonyl group, can be related to a successful accommodation of the target urethane bond in the cavity that hypothetically can act as an oxyanion hole throughout the enzymatic reaction. Configurations that exhibited interatomic distances inferior to 3.5 Å for Oγ(Ser190nuc)–Ccarbonyl(DUE-MDA), and under 2.5 Å for the remaining four metrics were considered as being catalytically competent and potential candidates for the final UMG-SP2:DUE-MDA model. Approximately 32% of all configurations fulfilled such criteria. These configurations were ranked in ascending order of the sum of the five distances, and the top-ranked configuration (i.e., the one with the smallest sum) was selected for QM/MM calculations, excluding those obtained during the first 20 ns of the production phase to ensure the use of a properly equilibrated structure.
We continued our study using the ONIOM subtractive scheme49 with electrostatic embedding in all QM/MM calculations, as implemented in the Gaussian 09 software. The valences of the bonds that crossed the boundary between the QM and MM layers were saturated with hydrogen atoms using the link-atom approach.50,51 All of the solvent molecules located in the MM region were frozen using the molUP plugin for the VMD software.52,53 The QM/MM model was initially optimized and later submitted to linear transit scans along putative reaction coordinates to investigate the steps underlying the catalytic mechanism. The linear transit scans' maxima were considered guesses for subsequent full transition state (TS) optimizations. Then, resorting to Intrinsic Reaction Coordinate (IRC) calculations, we obtained the minima associated with the optimized TSs, and their structure was subsequently optimized. We verified the nature of all stationary states as either TS (single imaginary frequency) or minima (absence of imaginary frequencies) with vibrational frequency calculations. All geometry optimizations, IRC, and vibrational frequency calculations were carried out using the B3LYP density functional54,55 and the 6-31G(d) basis set. Single-point QM/MM energy calculations were conducted at the B3LYP/6-311+G(2d,2p)-D3(BJ):ff14SB level of theory in the fully optimized stationary states. Grimme's D3 dispersion with Becke–Johnson damping56 was included in all calculations, as implemented in the Gaussian 09 software. The zero-point energy, as well as the thermal and entropic contributions to the Gibbs energy (calculated with the particle in a box/rigid rotor/harmonic oscillator formalism) were subsequently added to the final electronic energies of each stationary state to yield the corresponding Gibbs energy. Only vibrational temperatures larger than 120 K (≈100 cm−1) were considered for the calculation of entropic and enthalpic corrections, as validated elsewhere.57
To determine whether the calculated Gibbs free energy profile is independent of the selected density functional, we conducted single-point calculations on the isolated QM layer of each stationary point using ORCA 4.2.1 software.58 We employed B3LYP/6-311+G(2d,2p)-D3(BJ) and three additional theoretical methods: PWPB95, DSD-PBEB95, and SCS-MP2, all with the def2-TZVPP basis set. The energy difference between the three theoretical methods and B3LYP was then added as a correction to the QM/MM free energies previously calculated with Gaussian 09, yielding the final Gibbs free energies, which are presented and discussed in Table S2 and Fig. S4.†
We studied the catalytic mechanism of the deacylation reaction (STAGE 2) using a similar procedure. We built the reactant state of the deacylation stage from the product of the acylation reaction after removing the leaving group, i.e., the triethylene glycol monomethyl ether. The QM/MM system used to study the deacylation reaction comprised 9905 atoms, among which 117 formed the QM layer, and the remaining 9788 atoms were included in the MM region.
Finally, we obtained the reactant structure to start exploring the reactional state, from and after optimizing the assembled structure.
Further analysis revealed that the RMSD of the DUE-MDA ligand exhibited larger fluctuations compared to that of the enzyme. This disparity reflects the intrinsic flexibility of DUE-MDA, whose polyether tails possess high conformational freedom due to multiple rotatable bonds. Despite these fluctuations, the target urethane bond region of the substrate remained firmly lodged in the active site, with minimal deviation from the catalytically relevant positioning. Meanwhile, the enzyme displayed only minor loop movements, which did not alter the active site architecture or the binding of the substrate.
A UMG-SP2:DUE-MDA conformation was selected from the MD simulation according to a set of distance-based criteria (methods section). The structural alignment of the chosen conformation and the original X-ray structure of UMG-SP226 (Fig. S3†) revealed that the two structures are very similar, both in their overall fold (backbone RMSD of 1.25 Å) and in the organization of the active site region (all-atom RMSD of 0.60 Å for the Lys91, Ser166cis, Ser167, Ser185, Ile187, Gly188, and Ser190nuc set of residues). Together, these findings highlight the quality and stability of the assembled UMG-SP2:DUE-MDA model.
As UMG-SP2 belongs to the amidase signature superfamily, we would expect that its catalytic reaction involved a nucleophilic attack mediated by the conserved Ser190nuc-Ser166cis-Lys91 triad. Our calculations revealed that before the attack on the target urethane bond of DUE-MDA takes place, the nucleophilic Ser190nuc side chain hydroxyl must become activated, i.e. ionized. We observed that this occurs through a concerted and asynchronous reaction, in which Lys91 deprotonates Ser166cis, thus activating it so that Ser166cis in turn can abstract a proton from the nucleophilic Ser190nuc. In the TS associated with this step (TS1, imaginary frequency of 909.1i cm−1), the Hγ(Ser166cis) was nearly equidistant to Oγ(Ser166cis) and Nζ(Lys91) (1.23 Å vs. 1.28 Å), whereas Hγ(Ser190nuc) remained closer to its original position (1.04 Å to the Ser190nuc's Oγ) – Fig. 2B. Throughout the concerted proton transfer, the Oγ(Ser190nuc) builds up electron density and increases its negative character, as shown by the variation in its atomic charge when moving from the reactant (−0.22 a.u.) to the first intermediate state, INT1 (−0.33 a.u.). In the latter, the Bürgi–Dunitz angle, measured from Ser190nuc's side chain to the target urethane bond, adopted a value of 108°. Together, these findings reveal that the proton-transfer events rendered Ser190nuc fully competent to conduct the attack on the substrate's carbonyl carbon. The calculations have shown that INT1 (Fig. 2C) corresponds to a stationary state in the potential energy surface, but not to a minimum in the Gibbs free energy profile (activation Gibbs free energy of 0.3 kcal mol−1 and reaction Gibbs free energy of 1.5 kcal mol−1), meaning that it is not a stable intermediate of the reaction cycle. This suggests that the activation of Ser190nuc should be concerted with the nucleophilic attack, although occurring in an early phase of the mechanistic step. Nevertheless, as the energy difference is very small, the enzyme should be able to easily switch between the two states.
Once activated, Ser190nuc carried out the attack on the carbonyl carbon of the DUE-MDA's urethane bond (Fig. 3). This reaction exhibited an activation Gibbs free energy (ΔG‡) of 17.4 kcal mol−1, and it was endergonic in 16.0 kcal mol−1. The TS of this catalytic step was characterized by an imaginary frequency of 151.1i cm−1 (TS2 – Fig. 3B) that was mostly associated with the stretching of the forming Oγ(Ser190nuc)–Ccarbonyl(DUE-MDA) bond.
As Oγ(Ser190nuc) approached the substrate's carbonyl carbon (1.69 Å in TS2), the negative character of the atoms that comprise the urethane bond increased, specifically that of Ocarbonyl atom (DUE-MDA), which changed from −0.28 a.u. (reactant) to −0.32 a.u. (TS2). The negative charge that built up on the carbonyl group was stabilized by the oxyanion hole backbone amide groups of Ile187 and Gly188 (ca. 1.76 Å in TS2) and, to a lower extent, by the backbone amide group of the Ser190nuc nucleophile (2.30 Å in TS2). The calculations revealed that the product of this step corresponded to a tetrahedral intermediate (INT2-TI – Fig. 3C), in which the attacking Ser190nuc was covalently bound to the DUE-MDA substrate (1.52 Å vs. 2.72 Å in the reactant state). In this stationary state, the substrate's carbonyl C–O bond exhibited a bond length (1.30 Å) that fell between those typical of single and double bonds, evidencing the increased negative character of this group. Consequently, it was in INT2-TI where the interactions with the oxyanion hole amide groups were stronger.
Throughout the nucleophilic attack, the ion–dipole interaction between Ser190nuc and Ser166cis became progressively weaker until it was lost in INT2-TI (Fig. 3C). This occurred concomitantly to the establishment of a new hydrogen bond between Ser166cis and the Oester of the urethane bond under attack (1.74 Å in INT2-TI). As a result, Ser166cis moved much closer to the urethane bond's Oester than to its Namide (1.74 Å vs. 3.10 Å), thus adopting a position that should favor the protonation of Oester(DUE-MDA) rather than the Namide(DUE-MDA) when the urethane bond breaking occurred. Such a role in facilitating the leaving group's protonation by the bridging Sercis of Sernuc-Sercis-Lys triads was also reported for other enzymes of the amidase signature superfamily, such as the fatty acid amide hydrolase.59 The novelty here lies in Ser166cis that appears to play an important role in selecting which moiety should be first released as the leaving group. Specifically, our calculations showed that Ser166cis cleaves the ester moiety rather than the amide side of the urethane bond, as we will discuss subsequently.
The next catalytic step involved the tetrahedral intermediate breakdown and the consequent cleavage of the urethane bond, concerted with a proton transfer from Ser166cis to the Oester(DUE-MDA), yielding an alcohol-leaving group (Fig. 4). This step was characterized by a ΔG‡ of 4.8 kcal mol−1 and was exergonic in −3.2 kcal mol−1.
As the reaction proceeded from INT2-TI to TS3 (Fig. 4B), the Oester(DUE-MDA)–Ccarbonyl(DUE-MDA) bond elongated from 1.48 Å to 2.09 Å and the substrate's carbonyl group became more positively charged. As a result, this group re-acquired the double bond character, its interactions with the oxyanion hole became weaker, and the Oγ(Ser190nuc)–Ccarbonyl(DUE-MDA) bond shortened (1.52 Å in the INT2-TI vs. 1.38 Å in TS3). On the other hand, Oester(DUE-MDA) built up electronic density (charge changed from −0.15 a.u. in INT2-TI to −0.25 a.u. in TS3) and became more prone to receive a proton from Oγ(Ser166cis); in the TS3, the proton was virtually equidistant to both atoms (Fig. 4B). The TS of this step was characterized by an imaginary frequency at 849.6i cm−1 that was mostly dominated by the stretching of the atoms involved in the proton transfer. This indicates that, although concerted, the collapse of the tetrahedral intermediate (i.e., urethane bond cleavage) preceded the proton transfer reaction, in an asynchronous event. When going from TS3 to the INT3 state (Fig. 4C), the Oester(DUE-MDA)–Ccarbonyl(DUE-MDA) bond became completely cleaved (2.79 Å), the proton was successfully transferred from Ser166cis to the (now) alcohol-leaving group, and the Oγ(Ser190nuc) became fully attached to the substrate's carbonyl carbon (1.33 Å). Moreover, the negative character of Oγ(Ser166cis) increased substantially (−0.21 a.u. to −0.30 a.u.), which allowed it to become closer to the proton it previously transferred to Lys91 in the first catalytic step (1.40 Å in INT3).
The final step of the enzyme acylation stage consisted of a return of the proton from Lys91 to Ser166cis, restoring their initial protonation state (Fig. 5). Although we managed to characterize a transition state for this reaction (TS4, 841.7i cm−1 – Fig. 5B), our QM/MM calculations showed that the activation barrier required to go from INT3 to the final acyl-enzyme state (AE – Fig. 5C) vanished upon the introduction of the thermal and entropic contributions to the Gibbs energy. This means that this catalytic step is barrierless and that once INT3 was formed, the proton transfer from Lys91 to Ser166cis occurred spontaneously, yielding the AE with a reaction Gibbs free energy of −1.9 kcal mol−1. In the TS4 structure, the proton being shuttled to Ser166cis was nearly halfway from the donor Lys91 residue (closer to Ser166cis by 0.01 Å), and at the AE state the proton transfer was completed and both residues became neutral. The structural arrangement of these two states was virtually identical (all-atom RMSD of the QM layer atoms of 0.07 Å), which explains their resemblance in energy terms (Gibbs free energy difference of just 0.1 kcal mol−1).
After the acylation reaction, the resulting alcohol-leaving group, triethylene glycol monomethyl, was ready to abandon the active site, making it accessible to the solvent molecules required for the hydrolysis of the AE complex.
In the reactant state (AE* – Fig. 6A), the catalytic water rested at 3.32 Å of the substrate's carbonyl carbon, and its position was mainly dictated by the hydrogen bonds it established with Ser166cis (1.85 Å) and with Ser185 (2.15 Å). Similarly to the reactant state of the acylation stage (R – Fig. 2A), Lys91 acted as a hydrogen bond donor to Ser167 (2.01 Å) and Ser185 (2.15 Å), while simultaneously being an acceptor of a strong hydrogen bond from Ser166cis (1.54 Å). The substrate's carbonyl group remained lodged in the oxyanion hole cavity (at 1.79 and 2.20 Å of the amide groups) and interacted with the backbone amide of Ser190nuc (2.04 Å).
The first step of the deacylation stage consisted of a concerted reaction that combined three elementary steps: a proton transfer from Ser166cis to Lys91, a second proton transfer from Watcat to Ser166cis, and a nucleophilic attack conducted by the Watcat on the carbonyl carbon of the acylated Ser190nuc. This step was characterized by a ΔG‡ of 17.8 kcal mol−1 and was exergonic in −6.0 kcal mol−1. Despite the concerted nature of these catalytic events, our calculations showed that they occurred asynchronously, as discussed below. The transition state of this step (TS5 – Fig. 6B) was characterized by an imaginary frequency at 403.3i cm−1, which was largely dominated by the stretching of the O(Watcat)–Ccarbonyl(DUE-MDA) and H(Watcat)–Oγ(Ser166cis) bonds. In this state, the proton transfer from Ser166cis to Lys91 was completed (1.09 Å vs. 1.54 Å in the AE* state), the Watcat shared one of its hydrogen atoms with Ser166cis (1.32 Å vs. 1.85 Å in the AE* state), and the nucleophilic O(Watcat) rested at 1.89 Å from the target carbonyl carbon atom. This indicated that, within the same catalytic step, the activation of Ser166cis by Lys91 preceded the remaining two events. We believe that this should be required for Ser166cis to build up more electronic density on its side chain oxygen (charge changed from −0.24 a.u. to −0.28 a.u. from AE* to TS5) and to become more prone to activate the nucleophilic Watcat.
As the reaction proceeded from TS5 to INT5 (Fig. 6C), the hydroxyl group originating from Watcat became bound to the substrate's carbonyl carbon (1.33 Å in the INT5 structure), and the bond between Oγ(Ser190nuc) and the substrate was cleaved (2.55 Å in INT5 vs. 1.40 Å in TS5). In the acylation stage, we characterized a tetrahedral intermediate that originated from Ser190nuc's nucleophilic attack on the carbonyl group of the substrate's urethane bond (INT2-TI – Fig. 3C). However, in the current stage, we did not observe the formation of a stable tetrahedral intermediate resulting from the nucleophilic attack conducted by Watcat. Indeed, the IRC calculation performed to obtain INT5 from TS5 showed that the system passed through a tetrahedral geometry that immediately decayed to the INT5 state, and that, during this process, Ser190nuc became deacylated and the product was released in the form of a carbamic acid (Fig. 6C). Furthermore, it also revealed that the negative character of the Oγ(Ser190nuc) atom increased as Ser190nuc became deacylated (charge varied from −0.13 a.u. in the TS5 to −0.32 a.u. in the INT5), which induced the establishment of a new ion–dipole interaction with Ser166cis (1.53 Å), specifically with the hydrogen atom that Ser166cis previously received from Watcat.
The second step of the deacylation stage concluded the overall catalytic mechanism of UMG-SP2. It corresponded to the regeneration of the enzyme's initial state and proceeded via two concerted and asynchronous proton transfers between the residues of the Ser190nuc-Ser166cis-Lys91 triad (Fig. 7): Ser166cis shuttled a proton to Ser190nuc and accepted a proton from Lys91 (the one transferred in the opposite direction during the previous catalytic step). A closer look at the TS of this reaction (TS6 – Fig. 7B) revealed that when the proton was nearly halfway from Lys91 and Ser166cis (1.30 Å to Lys91 and 1.21 Å to Ser166cis), the bond between Oγ(Ser190nuc) and the proton transferred from Ser166cis was already established (1.05 Å), corroborating the asynchronous nature of these events. TS6 exhibited a single imaginary frequency at 822.0i cm−1, largely dominated by the stretching of the atoms involved in the proton transfer between Lys91 and Ser166cis.
The calculations revealed that this reaction led to a state (P – Fig. 7C) that corresponded to a minimum in the potential energy surface, but not to a stable intermediate in the thermal Gibbs free energy profile (ΔG‡ = 0.5 kcal mol−1 and ΔGreaction = 1.6 kcal mol−1). This phenomenon was also observed in the first catalytic step of the entire mechanism, where we hypothesized that the enzyme should be able to easily switch between the charged (Ser190nuc−-Ser166cis-Lys91+, observed in INT1) and neutral (Ser190nuc-Ser166cis-Lys91, observed in R) states. We believe that the same applies in this step and that both states are interchangeable.
The catalytic reaction occurred in two well-defined stages: in STAGE 1, the enzyme becomes covalently bound to the PU substrate, and an alcohol-leaving group is released; in STAGE 2, the enzyme:ligand covalent adduct is hydrolyzed by a catalytic water molecule and the product is released in the form of a carbamic acid (Scheme 2).
According to Fig. 8, the complete catalytic reaction involves a Gibbs activation free energy of 20.8 kcal mol−1 for the highest barrier identified. While this represents the primary barrier characterized in our study, additional phenomena such as product dissociation and solvent diffusion, which occur under real-world conditions, may also contribute to the overall reaction energetics. Even though the turnover rate for this enzyme has not been reported, we find reassurance in the fact that the theoretically calculated energy barrier laid within the range of values (∼13–23 kcal mol−1) corresponding to the experimental catalytic rates of most known hydrolases.60 The rate-limiting transition state of the overall reaction (TS3) corresponded to the concerted reaction that combined the tetrahedral intermediate breakdown and the cleavage of the substrate's urethane bond by its ester moiety with the subsequent formation of the alcohol-leaving group.
Our calculations revealed that a carbamic acid product resulted from the urethane bond cleavage of DUE-MDA by UMG-SP2. Carbamic acids are known to be unstable at room temperature and to quickly decompose to form an amine and carbon dioxide (CO2).61,62 Therefore, we believe the product we characterized should quickly eliminate CO2 and give rise to a monosubstituted MDA-urethane compound, specifically mono-urethane ethylene 4,4′-methylenedianiline (MUE-MDA). A second catalytic cycle of UMG-SP2 would complete the cleavage of the remaining urethane bond of MUE-MDA, releasing the alcohol-leaving group, triethylene glycol monomethyl, and the carbamic acid form of 4,4′-methylenedianiline (MDA). Again, the latter should quickly decompose into CO2 and MDA (the amine), which, interestingly, was detected as the end product of UMG-SP1's activity on a synthetic MDA diisocyanate-based thermoplastic polyester-PU.25
By definition, residues that destabilize the rate-limiting TS (TS3) in relation to the reactant state increase the energy barrier, while those that stabilize it will decrease the barrier. Based on the nature of TS3 and the reactant state, we predict that positive residues closer to the Oester(DUE-MDA) than to the Oγ(Ser190nuc) will stabilize the TS3 in relation to the reactant state. Conversely, positive residues nearer the Oγ(Ser190nuc) will do the opposite. For negative residues, those closer to the Oγ(Ser190nuc) are expected to stabilize TS3 in relation to the reactant state, while those nearer to the Oester(DUE-MDA) will do the opposite. The residues that destabilize TS3 are the most promising targets for mutation.
Fig. 9 shows the impact of each deleted MM residue on the energy barrier. Additionally, the energy contribution of each residue is plotted as a function of the difference between the distance to the positive side of the TS3 macrodipole (Oγ(Ser190nuc)) and the distance to the negative side (Oester(DUE-MDA)). For this measurement, we assigned a reference carbon atom for each type of residue (Table S4†).
We observed that most residues with a significant energy contribution to the barrier were charged. Additionally, the majority of those charged residues were closer to the Oester(DUE-MDA) than to Oγ(Ser190nuc). Furthermore, most negative residues were located closer to the negative side of the TS3 macrodipole. Consistent with our prediction, residues Asp95 and Asp226 destabilized TS3, leading to an increase of the energy barrier. Since these residues were closer to the negative side of the TS3 macrodipole, we suggest mutating them to polar residues (e.g., Asn) for a more conservative strategy, or to positive residues (Lys or Arg) for a structurally riskier approach. Conversely, Asp142 was closer to Oγ(Ser190nuc) than to Oester(DUE-MDA), placing it closer to the positive side of the TS3 macrodipole, thereby lowering the energy barrier.
We also predicted that positive residues closer to the Oester(DUE-MDA) than to Oγ(Ser190nuc) stabilize the TS3, contributing to a lower energy barrier. In line with our rationale, Lys324 and Arg325, located near the negative side of the macrodipole, stabilized TS3. On the other hand, we predicted that positive residues closer to the Oγ(Ser190nuc) than to Oester(DUE-MDA) destabilize TS3. Accordingly, Arg192 and Lys204, being near the positive side of the TS3 macrodipole, destabilized TS3. Notably, Arg192 significantly increased the energy barrier, making it the most promising mutational target. Arg192 is surrounded by neutral residues, with a lack of strong electrostatic interactions in its vicinity. The absence of stabilizing interactions for the positive charge resulted in the destabilization of its surroundings. This charged residue is close to the backbone of Ser190nuc, Ile187, and Gly188, which are key residues for the reaction, meaning their stability is essential. The presence of a nearby unstable positive charge in the vicinity of these residues led to their destabilization, consequently destabilizing TS3. Hence, Arg192 should be mutated to a polar or hydrophobic residue (e.g., Gln, His, or Met) to stabilize the neutral residue network. Concerning Lys204, as it is closer to the positive side of the TS3 macrodipole, we propose mutating it to a neutral residue (e.g., Gln or Met) for a more conservative strategy or to a negative residue (e.g., Asp or Glu) for a more aggressive strategy.
Based on our method, we believe that mutating the four aforementioned targets should stabilize the TS3 macrodipole. This stabilizing effect will lower the energy barrier, leading to an increase in the kcat term, achieving our goal of improving the UMG-SP2 hydrolysis of PU substrates.
Our calculations demonstrated that UMG-SP2 cleaves urethane bonds in two mechanistic stages, acylation and deacylation, and that the first comprises the rate-limiting step with an overall ΔG‡ of 20.8 kcal mol−1. The catalytic cycle culminates with the release of the final product in the form of a carbamic acid (Fig. 8A). Notably, the enzyme does not fully degrade the target urethane bond, but rather cleaves its ester moiety (esterase-like activity), leading to a highly unstable product. The latter should, in an enzyme-independent manner, rapidly decompose to form CO2 and an amine, causing the complete degradation of the urethane bond.
The QM/MM calculations shed light on the specific role of the active site residues during the urethane bond hydrolysis. They corroborated that Ser190nuc is the nucleophile, whose reactivity is controlled by the remaining two residues of the catalytic triad (Lys91 and Ser166cis). Lys91 establishes persistent hydrogen bond interactions with two nearby residues (Ser167 and Ser185) and, by deprotonation/protonation events, respectively activates/deactivates the bridging Ser166cis. The latter acts as a catalytic base, activating the nucleophilic Ser190nuc, and is also responsible for the protonation of the substrate's leaving group. In addition, Ser166cis plays a major role in selecting which moiety (ester vs. amine) should be released as the leaving group. The structural rearrangement of Ser166cis's side chain during the nucleophilic attack by Ser190nuc facilitates the proton transfer to Oester(DUE-MDA) and the subsequent urethane bond cleavage by its ester side. Throughout these events, the enzyme's oxyanion hole accommodates the urethane bond's carbonyl and stabilizes the tetrahedral reaction intermediates.
Finally, we have identified four mutational targets predicted to stabilize the TS3 macrodipole, which are expected to decrease the energy barrier and enhance the catalytic efficiency of UMG-SP2 for urethane bond cleavage.
Overall, the findings reported herein offer valuable insight into the catalytic mechanism underlying the hydrolysis of PU substrates by UMG-SP2. We hypothesize that, to some extent, many of the reported phenomena may be common to the other two metagenomic urethanases (UMG-SP1 and UMG-SP3) and to other enzymes of the amidase signature superfamily, although further studies are needed to corroborate this assumption. We hope this work encourages future research on the enzymatic depolymerization of PU that aims to address the environmental issues arising from the widespread use of plastics.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4sc06688j |
This journal is © The Royal Society of Chemistry 2025 |