Next Article in Journal
Rh-Catalyzed Environmentally Benign Selective Hydrogenation of a Broad Variety of Functional Groups Using Al-Water as a Hydrogen Source
Next Article in Special Issue
Discovery and Heterologous Expression of Unspecific Peroxygenases
Previous Article in Journal
Structurally Modified MXenes-Based Catalysts for Application in Hydrogen Evolution Reaction: A Review
Previous Article in Special Issue
Not a Mistake but a Feature: Promiscuous Activity of Enzymes Meeting Mycotoxins
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Combined Computational–Experimental Study on the Substrate Binding and Reaction Mechanism of Salicylic Acid Decarboxylase

1
College of Biotechnology, Tianjin University of Science and Technology, Tianjin 300457, China
2
Haihe Laboratory of Synthetic Biology, Tianjin 300308, China
3
Tianjin Institute of Industrial Biotechnology, Chinese Academy of Sciences, Tianjin 300308, China
4
School of Pharmacy, North Sichuan Medical College, Nanchong 637100, China
5
National Center of Technology Innovation for Synthetic Biology and National Engineering Research Center of Industrial Enzymes, Tianjin 300308, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Catalysts 2022, 12(12), 1577; https://doi.org/10.3390/catal12121577
Submission received: 26 October 2022 / Revised: 17 November 2022 / Accepted: 30 November 2022 / Published: 4 December 2022
(This article belongs to the Special Issue Advances in Biocatalysis and Enzyme Engineering)

Abstract

:
Salicylic acid decarboxylase (SDC) from the amidohydrolase superfamily (AHS) catalyzes the reversible decarboxylation of salicylic acid to form phenol. In this study, the substrate binding mode and reaction mechanism of SDC were investigated using computational and crystallographic methods. Quantum chemical calculations show that the enzyme follows the general mechanism of AHS decarboxylases. Namely, the reaction begins with proton transfer from a metal-coordinated aspartic acid residue (Asp298 in SDC) to the C1 of salicylic acid, which is followed by the C–C bond cleavage, to generate the phenol product and release CO2. Interestingly, the calculations show that SDC is a Mg-dependent enzyme rather than the previously proposed Zn-dependent, and the substrate is shown to be bidentately coordinated to the metal center in the catalysis, which is also different from the previous proposal. These predictions are corroborated by the crystal structure of SDC solved in complex with the substrate analogue 2-nitrophenol. The mechanistic insights into SDC in the present study provide important information for the rational design of the enzyme.

1. Introduction

Salicylic acid decarboxylase (SDC) catalyzes the reversible decarboxylation of salicylic acid (SA) to form phenol (Scheme 1) [1]. The reaction in the carboxylation direction is a biological alternative to the traditional Kolbe–Schmitt reaction [2]. This method of using enzymes provides a strategy to directly fix CO2 under mild conditions and the reaction rarely produces by-products, showing high potential for industrial applications [3,4]. The product at the carboxylation direction, salicylic acid, is an important chemical raw material which is commonly used in the production of aspirin drugs [5] and cosmetics [6].
SDC is a metal-dependent decarboxylase belonging to the amidohydrolase superfamily (AHS), which can catalyze the decarboxylation reaction independent of cofactors and O2. The AHS enzymes share significant structural and mechanistic similarities, especially the unique (β/α)8 TIM-barrel fold structure of the active site and divalent metal ions [7]. Interestingly, in previous reports [8,9,10], the metal ions in the active center of this type of enzyme were often considered as Zn2+. However, in recent studies, it was found that they are Mg2+ or Mn2+ [11,12,13,14,15,16].
Interestingly, SDC is catalytically active toward various aromatic compounds in addition to the natural substrate [17,18], for example, 1-naphthol [19], p-aminosalicylic acid [20] and 3-methylsalicylic acid [21]. This further enhances the potential of SDC for industrial applications. Although SDC has a wide substrate spectrum, the wide-type enzyme displays low activity toward non-natural substrates. Thus, protein engineering has been used to produce mutants with improved catalytic performance [20,21,22,23].
In a recent study, the crystal structure of SDC from Trichosporon moniliiforme in complex with the natural substrate salicylic acid was reported (Figure 1A) [23]. A Zn2+ cation was proposed to be the catalytic metal in the active site and to be coordinated by one water, Glu8, His169, Asp298 and SA (Figure 1B). Moreover, molecular dynamics simulations and mutation experiments showed that the mutant MT3 (Y64T/P191G/F195V/E302D) enhanced catalytic activity by expanding the substrate binding pocket. The kinetic parameters Km and kcat of this enzyme in catalyzing SA decarboxylation were reported as (1.1 ± 0.1) × 10−3 M and 3.6 ± 0.2 s−1, respectively [23].
In the present study, the substrate binding mode and mechanism of the SDC-catalyzed decarboxylation reaction were studied using quantum chemical and experimental approaches. It was shown that SDC adopts a similar substrate binding mode and reaction mechanism to other metal-dependent AHS decarboxylases. Interestingly, our results suggested that the metal of the active center is Mg2+, rather than the previously proposed Zn2+. The detailed information on the substrate binding mode and reaction mechanism obtained in this study could provide useful information to guide the selection of amino acids in the rational design of SDC mutants with improved catalytic efficiency.

2. Results and Discussion

In the previous structural study on SDC [23], the metal ion in the active site was proposed to be Zn2+. However, in our studies on the other decarboxylases from the same family, namely the AHS superfamily, it was found that the metal ion in the active center is usually Mg2+ or Mn2+ [11,12,13,14,15,16]. We thus considered all three types of divalent metals for the entire pathway in the mechanistic study. It turns out that the reaction mechanisms and energy profiles with different metals are very similar; however, the Mg-containing model has the lowest barriers compared to the other two (as discussed below). The enzyme is thus more likely to be Mg-dependent, rather than the previously proposed Zn-dependent [23]. In the following part, we will first discuss the results concerning the Mg-containing system and then the alternatives with Zn2+ and Mn2+ being the assumed metals in the active site.
In the previously reported crystal structure of SDC in complex with the natural substrate (PDB ID: 6JQX), the substrate was proposed to bind to the metal in a monodentate mode [23]. However, it has been established for the other AHS decarboxylases that the substrate is coordinated to the metal in a bidentate mode and the monodentate mode is in fact unproductive [16]. It is thus interesting to address whether SDC also adopts the bidentate mode in the catalysis. To this end, the structures of the enzyme–substrate complexes with the bidentate mode (called Mode-A) and monodentate mode (called Mode-B) were optimized by using the cluster approach and the corresponding energies were calculated (Figure 2).
In the optimized structure of Mode-A (E:S-AMg in Figure 2A), Asp298 was set to be protonated, while the substrate hydroxyl group was set to be deprotonated due to its coordination to the metal ion. Additionally, an oxygen atom of the substrate carboxylate group was also coordinated to the metal ion. The carboxylate group of the substrate forms hydrogen bonds with a water molecule, Arg235 and Asp298, and the deprotonated hydroxyl group forms a hydrogen bond with another water. In the optimized Mode-B (E:S-BMg in Figure 2B), the hydroxyl group of SA is in the neutral form and Asp298 is in the deprotonated state. In addition to the coordination to the metal ion, the carboxylate group of the substrate also forms a hydrogen bond with the water molecule, which is involved in the hydrogen bond network with another water molecule, Pro66 and Tyr301.
The calculated energies showed that Mode-A is 1.9 kcal/mol lower than Mode-B. In other words, Mode-A is the preferred binding mode in the enzyme–substrate complex. Interestingly, the substrate was previously suggested to bind in a monodentate mode [23]. Since the energy difference between the two modes is not that big, we considered both in the following mechanistic study.
In the pathway with Mode-A (Scheme 2A), the first step of the reaction is the proton transfer from Asp298 to the substrate C1 to form the 2,4-dienone intermediate (Int-AMg, Figure 3). The barrier for this step is calculated to be 16.6 kcal/mol, and the energy of the formed intermediate is 9.4 kcal/mol higher than E:S-AMg (Figure 4). At the transition state of this step (TS1-AMg), the C-H bond distance and the O-H bond distance are 1.35 Å and 1.33 Å, respectively. Subsequently, the C-C bond cleavage takes place to generate the phenol and CO2 products via the transition state TS2-AMg. The energy barrier of this step is 15.8 kcal/mol, and the distance of the breaking C–C bond at TS2-AMg is 2.12 Å (Figure 4). The E:P-AMg is in energy 0.8 kcal /mol higher than E:S-AMg.
According to the calculation results, the energy barrier of the overall reaction is 16.6 kcal/mol, and the rate-limiting step is the proton transfer (Figure 4). The calculated value is in excellent agreement with the experimental data, which is 16.8 kcal/mol converted from the kcat value of 3.6 s−1 according to transition state theory [23]. It is interesting to compare the calculate energies of SDC with those of LigW [12], 2,3-DHBD [13], and γ-RSD [15], which also catalyze the decarboxylation of phenolic acids. The barrier of the rate-limiting proton transfer of SDC is found to be similar compared to the corresponding steps of all three other enzymes. For the step of C-C bond cleavage, the energy differences between SDC and the others are larger. However, this does not contribute to the difference in the reaction rate because the proton transfer is the rate-limiting step.
The kinetic isotopic effect (KIE) is helpful in understanding the nature of the rate-limiting step of the catalyzed reaction. We here predict the KIE values for SDC by re-calculating the zero-point energies (ZPEs) of E:S-AMg and the rate-limiting TS1-AMg by replacing the carboxyl carbon of the substrate with C13 and replacing the proton of Asp298 with deuterium, respectively. The calculated KIE values converted from the energy differences in ZPEs is 5.1 for the proton and 1.0 for the carboxyl carbon.
In the pathway with Mode-B (Scheme 2B), the reaction first goes through two steps of proton transfer, namely from the hydroxyl group of the substrate to Asp298 and from Asp298 to the substrate C1, which is followed by the C–C bond cleavage to form the products (Scheme 2B). Interestingly, this mechanism was calculated to be energetically unfavorable with very high barriers. The calculated energy of the transition state for the second proton transfer is 46.5 kcal/mol higher than that of the E:S-AMg, and the mechanism with Mode-B is thus obviously infeasible (see Figure S5 for optimized structures).
Since the metal in the active site was previously assumed to be Zn2+ [23], we calculated the corresponding energy profiles by using the same active site model as that shown in Figure 2, but the Mg2+ was replaced by Zn2+ (see Figures S6–S11 for optimized structures). The calculation results show that in the case of Zn-enzyme the energy of Mode-A is very similar to that of Mode-B (only 0.7 kcal/mol in favor of the former). However, the energy barrier of the pathway with Mode-A is 20.3 kcal/mol, which is 3.7 kcal/mol higher than that of the corresponding pathway of Mg-enzyme (Figure 4). Moreover, similar to that of the Mg-enzyme, the barrier of the proton transfer from Asp298 to the substrate C1 in the pathway with Mode-B is also very high (42.4 kcal/mol relative to the corresponding enzyme–substrate complex). Since Mode-B has been shown with significantly high barriers for both Mg- and Zn-containing models, for the examined scenario with the Mn-containing model we considered only the pathway with Mode-A (Figures S12–S17). It turns out that the calculated barrier of the overall reaction is also higher than the Mg-enzyme (by 3.0 kcal/mol, Figure 4).
According to the cluster calculations discussed above, the SDC enzyme is more likely to be Mg-dependent and the substrate is coordinated to the metal in a bidentate mode. To verify these predictions, we solved the crystal structure of SDC in complex with substrate analogue 2-nitrophenol (PDB ID: 8H41, see Table S1 for the data collection and refinement statistics). Analyses on the active site structure provide strong support to the prediction on the basis of the calculations.
First, the previously proposed Zn2+ is too negative to fit the electron density map of the metal binding site (Figure 5A). The metal in the active site might be an atom with fewer electrons, such as the computationally predicted Mg2+ (Figure 5C). The fitting result with Mg2+ indeed shows a perfect match. By using Mn2+, the electron density is slightly overestimated (Figure 5B). Thus, SDC is a Mg-dependent enzyme. Furthermore, the overall structure of SDC obtained in the present study (Figure 5D) is almost identical to the previously solved structure shown in Figure 1A. However, interestingly, the close-view of the active site clearly showed that 2-nitrophenol is coordinated in a bidentate mode (Figure 5E), consistent with the lowest energy and productive binding mode for the natural substrate on the basis of calculations. Another interesting point here is that the measured angle between the nitro group and the phenyl ring is only ca 5°. This is different from the other AHS decarboxylase LigW, for which the substrate analog in the solved X-ray structure was observed to be significantly distorted [11].
Taken together, the calculations predict, from an energetical point of view, that SDC is a Mg-dependent enzyme and show that the reaction follows the general mechanism of AHS decarboxylases consisting of the first proton transfer from a metal-coordinated aspartic acid and the following C-C bond cleavage, in which the substrate is coordinated to the metal in a bidentate mode. Crystallographic study provides support to the metal identity and binding mode of the substrate.

3. Computational and Experimental Details

3.1. Computational Details

All calculations in this study were performed using the Gaussian 16 program [24], with B3LYP-D3(BJ) hybrid functional [25,26,27,28]. In the geometry optimizations, the 6-31G (d,p) basis set was used for the C, H, O, and N atoms, and the LANL2DZ pseudopotential basis set was used for the divalent metal ions [29]. Frequency calculations were performed at the same theoretical level as geometric optimization to obtain the zero-point energies (ZPEs). To consider the effect of protein surrounding which is not included in the active site model, the single-point energy calculations were performed using the SMD model with a dielectric ε = 4.0 at the same level of theory [30]. To obtain more accurate energies, single-point energy calculations were carried out with a larger basis set, namely 6-311+G (2d,2p) for C, H, O, and N atoms and LANL2DZ for divalent metal ions. ZPE and solvation effects were added to the single point energies from the large basis set calculations. According to previous studies on the decarboxylation reactions, the entropy gain generated from CO2 release was estimated by its translational entropy, which is calculated to be 11.1 kcal/mol at room temperature. This value was added to the energy of the decarboxylation step [12,13,14,15,31,32,33,34].

3.2. Active Site Model

The quantum chemical cluster approach is employed in the present study. This method has been proven to be very powerful in investigating various aspects of enzymatic reactions [35,36,37,38,39]. The active site model used for the calculations was designed based on the crystal structure of SDC from Trichosporon moniliiforme in complex with the substrate (PDB ID: 6JQX) [23]. The model consists of the metal ion along with its ligands (Glu8, His169, Asp298, a water molecule, and salicylic acid) and other residues making up the active sites (Glu9, Ala10, Tyr27, Tyr64, Ser65, Pro66, Pro170, Gly190, Pro191, Phe195, His224, Glu227, Arg235, His238, Trp239, Ser273, Tyr301, and Glu302). Additionally, three other crystallographic water molecules were also included in the model. The truncations in the model were made at the α-carbons of the amino acid and hydrogen atoms were added to saturate the carbon. To maintain the overall structure of the active site, the truncated carbons and a number of hydrogens were kept fixed during the geometry optimization. The model consists of 322 atoms and has a total charge of 0.

3.3. Cloning and Protein Purification

The TmSdc gene (GenBank accession number DM040453) was cloned into the pQE80L vector. The pQE80L-TmSdc plasmid was transformed into an E. coli BL21(DE3) cell which was grown in LB medium at 37 °C to an OD600 of ~0.8 and then induced by 0.5 mM isopropyl β-D-thiogalactopyranoside (IPTG) at 16 °C for 20 h. Cells were harvested by centrifugation at 5000× g for 15 min and then re-suspended in lysis buffer containing 25 mM Tris-HCl, pH 7.5, 150 mM NaCl, and 20 mM imidazole, followed by disruption with a French Press. Cell debris was removed by centrifugation at 17,000× g for 1 h. The supernatant was then applied to a Ni-NTA column with an FPLC system (GE Healthcare). The target proteins eluted at ~100 mM imidazole when using a 20–250 mM imidazole gradient. Each protein was dialyzed against a buffer containing 25 mM Tris-HCl, pH 7.5, and loaded onto a Q Sepharose column. Target proteins were eluted at ~200 mM NaCl when using a 0–500 mM NaCl gradient. The purified proteins were passed through a Superdex 200 column and further concentrated to 8 mg/mL in buffer containing 25 mM Tris-HCl (pH 7.5) and 150 mM NaCl. The purity of each protein (>95%) was checked using SDS-PAGE analysis.

3.4. Crystallization, Data Collection, Structure Determination, and Refinement

The optimized crystallization condition of apo-TmSdc was 24% PEG1000, 0.2 M Tris, pH 7.5. In general, 1 µL protein (8 mg/mL) was mixed with 1 µL of reservoir solution in 24 well Cryschem Plates, and equilibrated against 300 µL of the reservoir. The TmSdc crystals in complex with 2-nitrophenol were obtained by soaking the apo-TmSdc crystals in mother liquor containing 10 mM ligand for 1 day. All crystallization experiments were conducted at 25 °C using the sitting-drop vapor-diffusion method. All of the X-ray diffraction data sets were tested and collected at beamlines BL02U1/BL10U2/BL17B/BL18U1/BL19U1 of the National Facility for Protein Science in Shanghai (NFPS) and Shanghai Synchrotron Radiation Facility (SSRF). The diffraction images were processed using HKL2000 [40]. The structure was solved using the molecular replacement (MR) method with Phaser program [41] from the Phenix [42] suite using the structure of TmSdc (PDB ID: 6JQW) [23] as the search model. The further model building and refinement was carried out using programs phenix.refine [43] and Coot [44]. Prior to structural refinements, 5% randomly selected reflections were set aside for calculating Rfree as a monitor [45]. Data collection and refinement statistics are summarized in Table S1.

4. Conclusions

In the present study, the metal identity, substrate binding mode, and reaction mechanism of salicylic acid decarboxylase (SDC) were investigated using the quantum chemical cluster approach, in combination with crystallographic study. The enzyme is here demonstrated to follow the general mechanism of the amidohydrolase superfamily (AHS). Namely, the reaction starts with the proton transfer from the metal-coordinated aspartic acid (Asp298) to the C1 position of the substrate, which is the rate-limiting step of the entire reaction, and then the C-C bond is broken to form the product. However, very interestingly, and different from the previous proposal, the metal ion in the active site of SDC is found to be Mg2+, and the substrate binds to Mg2+ in a bidentate mode. Namely, both the carboxylate group and the phenolic hydroxyl group of the substrate are coordinated to the metal ion. These calculation results are corroborated by the solved crystal structure of SDC in complex with the substrate analogue 2-nitrophenol. The obtained information on the substrate binding mode and the reaction mechanism would be helpful in guiding the selection of targeted sites for mutation in protein engineering.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal12121577/s1, the data collection and refinement statistics of TmSDC crystal, additional calculation results for the Mg-, Mn- and Zn-systems and the Cartesian coordinates of the optimized structures.

Author Contributions

Conceptualization, X.S., H.S. and W.L.; Formal Analysis, F.C., Y.Z., C.Z., W.W., J.G., Q.L., H.Q., Y.D., W.L., F.L., H.S. and X.S.; Investigation, F.C., Y.Z., J.G. and Q.L.; Writing—Original Draft Preparation, F.C., Y.Z. and W.L.; Writing—Review and Editing, X.S. and H.S.; Supervision, W.L., F.L., H.S. and X.S.; Funding Acquisition, X.S., C.Z. and Y.D. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the National Key R&D Program of China (2021YFA0911500) and the Tianjin Synthetic Biotechnology Innovation Capacity Improvement Project (TSBICIP-CXRC-026). C.Z. thanks the Bureau of Science and Technology Nanchong City (20SXQT0161) for financial support. Y.D. thanks the Natural Science Foundation of Tianjin (19JCZDJC34800) for the financial support.

Data Availability Statement

Not applicable.

Acknowledgments

We thank the staff from the BL02U1/BL10U2/BL17B/BL18U1/BL19U1 beamlines of the National Facility for Protein Science in Shanghai (NFPS) and Shanghai Synchrotron Radiation Facility (SSRF) for assistance during data collection.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kirimura, K.; Gunji, H.; Wakayama, R.; Hattori, T.; Ishii, Y. Enzymatic Kolbe–Schmitt Reaction to Form Salicylic Acid from Phenol: Enzymatic Characterization and Gene Identification of a Novel Enzyme, Trichosporon Moniliiforme Salicylic Acid Decarboxylase. Biochem. Biophys. Res. Commun. 2010, 394, 279–284. [Google Scholar] [CrossRef] [PubMed]
  2. Lindsey, A.S.; Jeskey, H. The Kolbe-Schmitt Reaction. Chem. Rev. 1957, 57, 583–620. [Google Scholar] [CrossRef]
  3. Luo, J.; Larrosa, I. C−H Carboxylation of Aromatic Compounds through CO2 Fixation. ChemSusChem 2017, 10, 3317–3332. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Plasch, K.; Hofer, G.; Keller, W.; Hay, S.; Heyes, D.J.; Dennig, A.; Glueck, S.M.; Faber, K. Pressurized CO2 as a Carboxylating Agent for the Biocatalytic Ortho-Carboxylation of Resorcinol. Green Chem. 2018, 20, 1754–1759. [Google Scholar] [CrossRef] [Green Version]
  5. Boullard, O.; Leblanc, H.; Besson, B. Salicylic Acid. In Ullmann’s Encyclopedia of Industrial Chemistry; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2000; p. a23_477. ISBN 978-3-527-30673-2. [Google Scholar]
  6. Madan, R.K.; Levitt, J. A Review of Toxicity from Topical Salicylic Acid Preparations. J. Am. Acad. Dermatol. 2014, 70, 788–792. [Google Scholar] [CrossRef] [PubMed]
  7. Seibert, C.M.; Raushel, F.M. Structural and Catalytic Diversity within the Amidohydrolase Superfamily. Biochemistry 2005, 44, 6383–6391. [Google Scholar] [CrossRef]
  8. Song, M.; Zhang, X.; Liu, W.; Feng, J.; Cui, Y.; Yao, P.; Wang, M.; Guo, R.-T.; Wu, Q.; Zhu, D. 2,3-Dihydroxybenzoic Acid Decarboxylase from Fusarium Oxysporum: Crystal Structures and Substrate Recognition Mechanism. ChemBioChem 2020, 21, 2950–2956. [Google Scholar] [CrossRef]
  9. Xu, S.; Li, W.; Zhu, J.; Wang, R.; Li, Z.; Xu, G.-L.; Ding, J. Crystal Structures of Isoorotate Decarboxylases Reveal a Novel Catalytic Mechanism of 5-Carboxyl-Uracil Decarboxylation and Shed Light on the Search for DNA Decarboxylase. Cell Res. 2013, 23, 1296–1309. [Google Scholar] [CrossRef] [Green Version]
  10. Goto, M.; Hayashi, H.; Miyahara, I.; Hirotsu, K.; Yoshida, M.; Oikawa, T. Crystal Structures of Nonoxidative Zinc-Dependent 2,6-Dihydroxybenzoate (γ-Resorcylate) Decarboxylase from Rhizobium Sp. Strain MTP-10005. J. Biol. Chem. 2006, 281, 34365–34373. [Google Scholar] [CrossRef] [Green Version]
  11. Vladimirova, A.; Patskovsky, Y.; Fedorov, A.A.; Bonanno, J.B.; Fedorov, E.V.; Toro, R.; Hillerich, B.; Seidel, R.D.; Richards, N.G.J.; Almo, S.C.; et al. Substrate Distortion and the Catalytic Reaction Mechanism of 5-Carboxyvanillate Decarboxylase. J. Am. Chem. Soc. 2016, 138, 826–836. [Google Scholar] [CrossRef]
  12. Sheng, X.; Zhu, W.; Huddleston, J.; Xiang, D.F.; Raushel, F.M.; Richards, N.G.J.; Himo, F. A Combined Experimental-Theoretical Study of the LigW-Catalyzed Decarboxylation of 5-Carboxyvanillate in the Metabolic Pathway for Lignin Degradation. ACS Catal. 2017, 7, 4968–4974. [Google Scholar] [CrossRef]
  13. Hofer, G.; Sheng, X.; Braeuer, S.; Payer, S.E.; Plasch, K.; Goessler, W.; Faber, K.; Keller, W.; Himo, F.; Glueck, S.M. Metal Ion Promiscuity and Structure of 2,3-Dihydroxybenzoic Acid Decarboxylase of Aspergillus Oryzae. ChemBioChem 2021, 22, 652–656. [Google Scholar] [CrossRef] [PubMed]
  14. Sheng, X.; Plasch, K.; Payer, S.E.; Ertl, C.; Hofer, G.; Keller, W.; Braeuer, S.; Goessler, W.; Glueck, S.M.; Himo, F.; et al. Reaction Mechanism and Substrate Specificity of Iso-Orotate Decarboxylase: A Combined Theoretical and Experimental Study. Front. Chem. 2018, 6, 608. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Sheng, X.; Patskovsky, Y.; Vladimirova, A.; Bonanno, J.B.; Almo, S.C.; Himo, F.; Raushel, F.M. Mechanism and Structure of γ-Resorcylate Decarboxylase. Biochemistry 2018, 57, 3167–3175. [Google Scholar] [CrossRef] [PubMed]
  16. Sheng, X.; Himo, F. Mechanisms of Metal-Dependent Non-Redox Decarboxylases from Quantum Chemical Calculations. Comput. Struct. Biotechnol. J. 2021, 19, 3176–3186. [Google Scholar] [CrossRef]
  17. Wuensch, C.; Glueck, S.M.; Gross, J.; Koszelewski, D.; Schober, M.; Faber, K. Regioselective Enzymatic Carboxylation of Phenols and Hydroxystyrene Derivatives. Org. Lett. 2012, 14, 1974–1977. [Google Scholar] [CrossRef]
  18. Payer, S.E.; Faber, K.; Glueck, S.M. Non-Oxidative Enzymatic (De)Carboxylation of (Hetero)Aromatics and Acrylic Acid Derivatives. Adv. Synth. Catal. 2019, 361, 2402–2420. [Google Scholar] [CrossRef] [Green Version]
  19. Ren, P.; Tan, Z.; Zhou, Y.; Tang, H.; Xu, P.; Liu, H.; Zhu, L. Biocatalytic CO2 Fixation Initiates Selective Oxidative Cracking of 1-Naphthol under Ambient Conditions. Green Chem. 2022, 24, 4766–4771. [Google Scholar] [CrossRef]
  20. Kirimura, K.; Yanaso, S.; Kosaka, S.; Koyama, K.; Hattori, T.; Ishii, Y. Production of P-Aminosalicylic Acid through Enzymatic Kolbe–Schmitt Reaction Catalyzed by Reversible Salicylic Acid Decarboxylase. Chem. Lett. 2011, 40, 206–208. [Google Scholar] [CrossRef]
  21. Kirimura, K.; Araki, M.; Ishihara, M.; Ishii, Y. Expanding Substrate Specificity of Salicylate Decarboxylase by Site-Directed Mutagenesis for Expansion of the Entrance Region Connecting to the Substrate Access Tunnel. Chem. Lett. 2019, 48, 58–61. [Google Scholar] [CrossRef]
  22. Ienaga, S.; Kosaka, S.; Honda, Y.; Ishii, Y.; Kirimura, K. P-Aminosalicylic Acid Production by Enzymatic Kolbe–Schmitt Reaction Using Salicylic Acid Decarboxylases Improved through Site-Directed Mutagenesis. Bull. Chem. Soc. Jpn. 2013, 86, 628–634. [Google Scholar] [CrossRef]
  23. Gao, X.; Wu, M.; Zhang, W.; Li, C.; Guo, R.-T.; Dai, Y.; Liu, W.; Mao, S.; Lu, F.; Qin, H.-M. Structural Basis of Salicylic Acid Decarboxylase Reveals a Unique Substrate Recognition Mode and Access Channel. J. Agric. Food Chem. 2021, 69, 11616–11625. [Google Scholar] [CrossRef] [PubMed]
  24. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.; et al. Gaussian 16, Revision C.01; Gaussian, Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  25. Grimme, S. Density Functional Theory with London Dispersion Corrections. WIREs Comput. Mol. Sci. 2011, 1, 211–228. [Google Scholar] [CrossRef]
  26. Becke, A.D. Density-functional Thermochemistry. III. The Role of Exact Exchange. J. Chem. Phys. 1993, 98, 5648–5652. [Google Scholar] [CrossRef] [Green Version]
  27. Lee, C.; Yang, W.; Parr, R.G. Development of the Colle-Salvetti Correlation-Energy Formula into a Functional of the Electron Density. Phys. Rev. B Condens. Matter 1988, 37, 785–789. [Google Scholar] [CrossRef] [Green Version]
  28. Bursch, M.; Caldeweyher, E.; Hansen, A.; Neugebauer, H.; Ehlert, S.; Grimme, S. Understanding and Quantifying London Dispersion Effects in Organometallic Complexes. Acc. Chem. Res. 2019, 52, 258–266. [Google Scholar] [CrossRef]
  29. Hay, P.J.; Wadt, W.R. Ab Initio Effective Core Potentials for Molecular Calculations. Potentials for K to Au Including the Outermost Core Orbitals. J. Chem. Phys. 1985, 82, 299–310. [Google Scholar] [CrossRef]
  30. Marenich, A.V.; Cramer, C.J.; Truhlar, D.G. Universal Solvation Model Based on Solute Electron Density and on a Continuum Model of the Solvent Defined by the Bulk Dielectric Constant and Atomic Surface Tensions. J. Phys. Chem. B 2009, 113, 6378–6396. [Google Scholar] [CrossRef]
  31. Lind, M.E.S.; Himo, F. Theoretical Study of Reaction Mechanism and Stereoselectivity of Arylmalonate Decarboxylase. ACS Catal. 2014, 4, 4153–4160. [Google Scholar] [CrossRef]
  32. Sheng, X.; Lind, M.E.S.; Himo, F. Theoretical Study of the Reaction Mechanism of Phenolic Acid Decarboxylase. FEBS J. 2015, 282, 4703–4713. [Google Scholar] [CrossRef]
  33. Sheng, X.; Himo, F. Mechanism of 3-Methylglutaconyl CoA Decarboxylase AibA/AibB: Pericyclic Reaction versus Direct Decarboxylation. Angew. Chem. Int. Ed. 2020, 59, 22973–22977. [Google Scholar] [CrossRef] [PubMed]
  34. Planas, F.; Sheng, X.; McLeish, M.J.; Himo, F. A Theoretical Study of the Benzoylformate Decarboxylase Reaction Mechanism. Front. Chem. 2018, 6, 205. [Google Scholar] [CrossRef] [PubMed]
  35. Himo, F.; de Visser, S.P. Status Report on the Quantum Chemical Cluster Approach for Modeling Enzyme Reactions. Commun. Chem. 2022, 5, 29. [Google Scholar] [CrossRef]
  36. Siegbahn, P.E.M. A Quantum Chemical Approach for the Mechanisms of Redox-Active Metalloenzymes. RSC Adv. 2021, 11, 3495–3508. [Google Scholar] [CrossRef]
  37. Himo, F. Recent Trends in Quantum Chemical Modeling of Enzymatic Reactions. J. Am. Chem. Soc. 2017, 139, 6780–6786. [Google Scholar] [CrossRef] [Green Version]
  38. Blomberg, M.R.A.; Borowski, T.; Himo, F.; Liao, R.-Z.; Siegbahn, P.E.M. Quantum Chemical Studies of Mechanisms for Metalloenzymes. Chem. Rev. 2014, 114, 3601–3658. [Google Scholar] [CrossRef]
  39. Sheng, X.; Kazemi, M.; Planas, F.; Himo, F. Modeling Enzymatic Enantioselectivity using Quantum Chemical Methodology. ACS Catal. 2020, 10, 6430–6449. [Google Scholar] [CrossRef]
  40. Otwinowski, Z.; Minor, W. Processing of X-Ray Diffraction Data Collected in Oscillation Mode. Methods Enzymol. 1997, 276, 307–326. [Google Scholar] [CrossRef]
  41. McCoy, A.J.; Grosse-Kunstleve, R.W.; Adams, P.D.; Winn, M.D.; Storoni, L.C.; Read, R.J. Phaser Crystallographic Software. J. Appl. Crystallogr. 2007, 40, 658–674. [Google Scholar] [CrossRef] [Green Version]
  42. Adams, P.D.; Afonine, P.V.; Bunkóczi, G.; Chen, V.B.; Davis, I.W.; Echols, N.; Headd, J.J.; Hung, L.-W.; Kapral, G.J.; Grosse-Kunstleve, R.W.; et al. PHENIX: A Comprehensive Python-Based System for Macromolecular Structure Solution. Acta Crystallogr. D Biol. Crystallogr. 2010, 66, 213–221. [Google Scholar] [CrossRef]
  43. Grosse-Kunstleve, R.W.; Afonine, P.V.; Moriarty, N.W.; Zwart, P.H.; Hung, L.-W.; Read, R.J. Iterative Model Building, Structure Refinement and Density Modification with the PHENIX AutoBuild Wizard. Acta Crystallogr. D Biol. Crystallogr. 2008, 64, 61–69. [Google Scholar] [CrossRef] [Green Version]
  44. Emsley, P.; Cowtan, K. Coot: Model-Building Tools for Molecular Graphics. Acta Crystallogr. D Biol. Crystallogr. 2004, 60, 2126–2132. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Brünger, A.T. Assessment of Phase Accuracy by Cross Validation: The Free R Value. Methods and Applications. Acta Crystallogr. Sect. D 1993, 49, 24–36. [Google Scholar] [CrossRef] [PubMed]
Scheme 1. Reaction catalyzed by salicylic acid decarboxylase (SDC).
Scheme 1. Reaction catalyzed by salicylic acid decarboxylase (SDC).
Catalysts 12 01577 sch001
Figure 1. (A) Crystal structure and (B) active site of salicylic acid decarboxylase (PDB ID: 6JQX).
Figure 1. (A) Crystal structure and (B) active site of salicylic acid decarboxylase (PDB ID: 6JQX).
Catalysts 12 01577 g001
Figure 2. Optimized structures of enzyme–substrate complexes with (A) bidentate binding mode and (B) monodentate binding mode.
Figure 2. Optimized structures of enzyme–substrate complexes with (A) bidentate binding mode and (B) monodentate binding mode.
Catalysts 12 01577 g002
Scheme 2. (A) The proposed mechanism of SDC reaction on the basis of calculations in the present study. (B) The previously proposed mechanism [19].
Scheme 2. (A) The proposed mechanism of SDC reaction on the basis of calculations in the present study. (B) The previously proposed mechanism [19].
Catalysts 12 01577 sch002
Figure 3. Optimized structures for the transition states (TS1-AMg and TS2-AMg) and intermediates (Int-AMg and E:P-AMg) in the reaction pathway of SDC. Only a part of the active site model is shown here. See Supplementary Materials for the structures with full models (Figures S1–S4).
Figure 3. Optimized structures for the transition states (TS1-AMg and TS2-AMg) and intermediates (Int-AMg and E:P-AMg) in the reaction pathway of SDC. Only a part of the active site model is shown here. See Supplementary Materials for the structures with full models (Figures S1–S4).
Catalysts 12 01577 g003
Figure 4. Calculated energy profiles for the decarboxylation reactions catalyzed by SDC with Mg2+, Mn2+, and Zn2+.
Figure 4. Calculated energy profiles for the decarboxylation reactions catalyzed by SDC with Mg2+, Mn2+, and Zn2+.
Catalysts 12 01577 g004
Figure 5. Electron density map (2Fobs-Fcalc) of models with (A) Zn2+, (B) Mn2+, or (C) Mg2+ in blue, contoured at 1.5 sigma, and the difference map (Fobs-Fcalc) is at 3.0 sigma. The B factors at full occupancy are 44.05 (Chain A) and 43.32 (Chain B) for Zn2+, 36.64 (Chain A) and 36.69 (Chain B) for Mn2+, and 16.79 (Chain A) and 16.97 (Chain B) for Mg2+. The average B factor for all atoms of the structure is 18.53, and the average B factor for all atoms of the ligand is 18.94. (D) Overall structure and (E) active site structure of SDC in complex with 2-nitrophenol (PDB ID: 8H41).
Figure 5. Electron density map (2Fobs-Fcalc) of models with (A) Zn2+, (B) Mn2+, or (C) Mg2+ in blue, contoured at 1.5 sigma, and the difference map (Fobs-Fcalc) is at 3.0 sigma. The B factors at full occupancy are 44.05 (Chain A) and 43.32 (Chain B) for Zn2+, 36.64 (Chain A) and 36.69 (Chain B) for Mn2+, and 16.79 (Chain A) and 16.97 (Chain B) for Mg2+. The average B factor for all atoms of the structure is 18.53, and the average B factor for all atoms of the ligand is 18.94. (D) Overall structure and (E) active site structure of SDC in complex with 2-nitrophenol (PDB ID: 8H41).
Catalysts 12 01577 g005
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Chen, F.; Zhao, Y.; Zhang, C.; Wang, W.; Gao, J.; Li, Q.; Qin, H.; Dai, Y.; Liu, W.; Liu, F.; et al. A Combined Computational–Experimental Study on the Substrate Binding and Reaction Mechanism of Salicylic Acid Decarboxylase. Catalysts 2022, 12, 1577. https://doi.org/10.3390/catal12121577

AMA Style

Chen F, Zhao Y, Zhang C, Wang W, Gao J, Li Q, Qin H, Dai Y, Liu W, Liu F, et al. A Combined Computational–Experimental Study on the Substrate Binding and Reaction Mechanism of Salicylic Acid Decarboxylase. Catalysts. 2022; 12(12):1577. https://doi.org/10.3390/catal12121577

Chicago/Turabian Style

Chen, Fuqiang, Yipei Zhao, Chenghua Zhang, Wei Wang, Jian Gao, Qian Li, Huimin Qin, Yujie Dai, Weidong Liu, Fufeng Liu, and et al. 2022. "A Combined Computational–Experimental Study on the Substrate Binding and Reaction Mechanism of Salicylic Acid Decarboxylase" Catalysts 12, no. 12: 1577. https://doi.org/10.3390/catal12121577

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop