QM/MM Molecular Dynamics Simulations Revealed Catalytic Mechanism of UreaseClick to copy article linkArticle link copied!
- Toru Saito*Toru Saito*E-mail: [email protected]Department of Biomedical Information Sciences, Graduate School of Information Sciences, Hiroshima City University, 3-4-1 Ozuka-Higashi, Asa-Minami-Ku, Hiroshima 731-3194 JapanMore by Toru Saito
- Yu TakanoYu TakanoDepartment of Biomedical Information Sciences, Graduate School of Information Sciences, Hiroshima City University, 3-4-1 Ozuka-Higashi, Asa-Minami-Ku, Hiroshima 731-3194 JapanMore by Yu Takano
Abstract
Urease catalyzes the hydrolysis of urea to form ammonia and carbamate, inducing an overall pH increase that affects both human health and agriculture. Inhibition, mutagenesis, and kinetic studies have provided insights into its enzymatic role, but there have been debates on the substrate binding mode as well as the reaction mechanism. In the present study, we report quatum mechanics-only (QM-only) and quantum mechanics/molecular mechanics molecular dynamics (QM/MM MD) calculations on urease that mainly investigate the binding mode of urea and the mechanism of the urease-catalyzed hydrolysis reaction. Comparison between the experimental data and our QM(GFN2-xTB)/MM metadynamics results demonstrates that urea hydrolysis via a complex with bidentate-bound urea is much more favorable than via that with monodentate-bound urea for both nucleophilic attack and the subsequent proton transfer steps. We also indicate that the bidentate coordination of urea fits the active site with a closed conformation of the mobile flap and can facilitate the stabilization of transition states and intermediates by forming multiple hydrogen bonds with certain active site residues.
This publication is licensed under
License Summary*
You are free to share(copy and redistribute) this article in any medium or format and to adapt(remix, transform, and build upon) the material for any purpose, even commercially within the parameters below:
Creative Commons (CC): This is a Creative Commons license.
Attribution (BY): Credit must be given to the creator.
*Disclaimer
This summary highlights only some of the key features and terms of the actual license. It is not a license and has no legal value. Carefully review the actual license before using these materials.
License Summary*
You are free to share(copy and redistribute) this article in any medium or format and to adapt(remix, transform, and build upon) the material for any purpose, even commercially within the parameters below:
Creative Commons (CC): This is a Creative Commons license.
Attribution (BY): Credit must be given to the creator.
*Disclaimer
This summary highlights only some of the key features and terms of the actual license. It is not a license and has no legal value. Carefully review the actual license before using these materials.
License Summary*
You are free to share(copy and redistribute) this article in any medium or format and to adapt(remix, transform, and build upon) the material for any purpose, even commercially within the parameters below:
Creative Commons (CC): This is a Creative Commons license.
Attribution (BY): Credit must be given to the creator.
*Disclaimer
This summary highlights only some of the key features and terms of the actual license. It is not a license and has no legal value. Carefully review the actual license before using these materials.
1. Introduction
Scheme 1
Figure 1
Figure 1. Active site of Sporosarcina pasteurii urease (SPU) and two possible substrate binding modes, monodentate 1 and bidentate 2.
Scheme 2
2. Computational Method
2.1. System Setup
2.2. Classical MD and QM/MM MD Equilibration
2.3. QM-Only Cluster Calculations
2.4. GFN2-xTB/MM Metadynamics Simulations
3. Results and Discussion
3.1. Possible Mechanisms Starting with 1 and 2
Figure 2
Figure 2. QM regions of 1 and 2 after the QM(GFN2-xTB)/MM MD equilibrations (top). Key bond distances are also presented in Å (bottom).
3.2. Potential Energy Profiles Obtained from QM-Only Cluster Calculations
Figure 3
Figure 3. Potential energy profiles (in kcal mol–1) and illustrations of the transition state and product structures corresponding to (a) mechanism (v) starting from 1′ and (b) mechanism (ii) starting from 2′ obtained at GFN2-xTB and UB3LYP/def2-SVP levels.
3.3. Free Energy Profiles Obtained from GFN2-xTB/MM Metadynamics Calculations
3.3.1. Complex with Monodentate-Bound Urea (1)
Figure 4
Figure 4. (a) One-dimensional potentials of mean force (1D-PMF) of the nucleophilic attack reaction and representative snapshots of the active site in (b) 1R, (c) 1TS1A, and (d) 1IntA, with key distances in Å.
Figure 5
Figure 5. (a) One-dimensional potentials of mean force (1D-PMF) of the proton transfer reaction and representative snapshots of the active site in (b) 1TS2A and (c) 1PA, with key distances in Å.
Figure 6
Figure 6. (a) One-dimensional potentials of mean force (1D-PMF) of the W2 dissociation and representative snapshots of the active site in (b) 1TSB and (c) 1PB, with key distances in Å.
3.3.2. Complex with Bidentate-Bound Urea (2)
Figure 7
Figure 7. (a) One-dimensional potentials of mean force (1D-PMF) of the nucleophilic attack reaction and representative snapshots of the active site in (b) 2R, (c) 2TS1A, and (d) 2IntA, with key distances in Å.
Figure 8
Figure 8. (a) One-dimensional potentials of mean force (1D-PMF) of the proton transfer reaction and representative snapshots of the active site in (b) 2TS2A and (c) 2PA, with key distances in Å.
4. Conclusions
Supporting Information
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.jpcb.1c10200.
Figure S1, active site models for 1 and 2 excluding H222, H323, and A366 with key distances and spin density; Cartesian coordinates of these models obtained with the UB3LYP/def2-SVP level; the created topology and parameter files; Figure S2, RMSDs with respect to the initial conformations during a 10 ns classical MD simulation for 1 and 2; Figure S3, overlay of smaller models (1, 2, and 3) optimized in the singlet and quintet states at GFN2-xTB with respect to the structure obtained with the broken-symmetry DFT method; Figure S4, structure and Cartesian coordinates of the QM region for 1 optimized in the open-shell singlet state at the UB3LYP/def2-SV(P) level; Figure S5, potential energy profile of the dissociation of W2; Cartesian coordinates for all stationary points of the QM-only cluster calculations optimized at the GFN2-xTB and UB3LYP levels; Tables S1–S5, key bond distance for all stationary points obtained from QM-only cluster and QM/MM metadynamics calculations; full citation for ref (59) (PDF)
Terms & Conditions
Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.
Acknowledgments
T.S. acknowledges the fund for the Promotion of Joint International Research (Fostering Joint International Research (B)) and for the Grants-in-Aid for Scientific Research (C) from the Japan Society for the Promotion of Science (Nos. 18KK0194 and 21K04985). Y.T acknowledges the fund for the Grants-in-Aid for Scientific Research (C) from the JSPS (No. 19K06589). Y.T. is also grateful to the Ministry of Education, Culture, Sports, Science and Technology (MEXT) for a Grant-in-Aid for Scientific Research on Innovative Areas “3D active-site science” (No. 26105012) and Transformative Research Areas (A) “Progressive condensed matter physics inspired by hyper-ordered structures” (No. 20H05883). Some computations were performed using Research Center for Computational Science, Okazaki, Japan.
References
This article references 64 other publications.
- 1Dixon, N. E.; Gazzola, C.; Blakeley, R. L.; Zerner, B. Jack Bean Urease (EC 3.5.1.5). Metalloenzyme. Simple Biological Role for Nickel. J. Am. Chem. Soc. 1975, 97, 4131– 4133, DOI: 10.1021/ja00847a045Google Scholar1Jack bean urease (EC 3.5.1.5). Metalloenzyme. Simple biological role for nickelDixon, Nicholas E.; Gazzola, Carlo; Blakeley, Robert L.; Zerner, BurtJournal of the American Chemical Society (1975), 97 (14), 4131-3CODEN: JACSAT; ISSN:0002-7863.Jack bean (Canavalia enisformis) urease in phosphate buffer, pH 7.0 (1 mM each in EDTA and β-mercaptoethanol) exhibited electronic and (or) vibrational transitions with broad absorption maxima at 1060 and 725 nm and shoulders at ∼425, ∼376, and ∼316 nm. The spectrum was reversibly modified by β-mercaptoethanol and by acetohydroxamic acid. Chem. and at. absorption anal. showed the presence of 2.0 g atoms of Ni/105,000 g of enzyme. The residual sp. activity was a linear function of the Ni content after denaturation at pH 3.7 in the presence of EDTA. Urease appears to be the 1st example of a Ni metalloenzyme. Ni may well be an essential trace element in jack beans. Published spectral characteristics of other enzymes are consistent with the presence of tightly-bound transition metal ions.
- 2Blakeley, R. L.; Zerner, B. Jack Bean Urease: the First Nickel Enzyme. J. Mol. Catal. 1984, 23, 263– 292, DOI: 10.1016/0304-5102(84)80014-0Google Scholar2Jack bean urease: the first nickel enzymeBlakeley, Robert L.; Zerner, BurtJournal of Molecular Catalysis (1984), 23 (2-3), 263-92CODEN: JMCADS; ISSN:0304-5102.A review and discussion with 98 refs.
- 3Hausinger, R. P. Nickel Utilization by Microorganisms. Microbiol. Rev. 1987, 51, 22– 42, DOI: 10.1128/mr.51.1.22-42.1987Google Scholar3Nickel utilization by microorganismsHausinger, Robert P.Microbiological Reviews (1987), 51 (1), 22-42CODEN: MBRED3; ISSN:0146-0749.A review with 232 refs. on biol. functions of Ni, transport of Ni, and Ni-contg. enzymes.
- 4Mobley, H. L.; Hausinger, R. P. Microbial Ureases: Significance, Regulation, and Molecular Characterization. Microbiol. Rev. 1989, 53, 85– 108, DOI: 10.1128/mr.53.1.85-108.1989Google Scholar4Microbial ureases: significance, regulation, and molecular characterizationMobley, Harry L. T.; Hausinger, Robert P.Microbiological Reviews (1989), 53 (1), 85-108CODEN: MBRED3; ISSN:0146-0749.A review, with 249 refs., of urease in pathogenesis, in the rumen and gastrointestinal tract, in soil and aq. environments, and in microbe physiol. Methods for microbial urease detn. and purifn., characterization of microbial ureases, and the mol. biol. of ureolysis (e.g. plasmid-encoded ureases, cloning of urease genes, bacterial urease operon, fungal and plant urease mol. biol.) are discussed.
- 5Blakeley, R. L.; Webb, E. C.; Zerner, B. Jack Bean Urease (EC 3.5.1.5). A New Purification and Reliable Rate Assay. Biochemistry 1969, 8, 1984– 1990, DOI: 10.1021/bi00833a031Google Scholar5Jack bean urease (EC 3.5.1.5). A new purification and reliable rate assayBlakeley, Robert L.; Webb, Edwin C.; Zerner, BurtBiochemistry (1969), 8 (5), 1984-90CODEN: BICHAW; ISSN:0006-2960.Jack bean urease (EC 3.5.1.5) was reproducibly purified to a hitherto unobtained and const. specific activity. The procedure is capable of handling "poor" jack bean meal. A CHCl3-Me2CO powder of jack bean meal is extd. with 30% Me2CO contg. 1% mercaptoethanol at 39° for 5 min. The filtrate from this step is allowed to stand for 48 hrs. at 4°, and the crystals of the enzyme are harvested by centrifugation. After recycling gel filtration in the presence of 1 mM mercaptoethanol and 1 mM EDTA (Sephadex G-200), and concn. by (NH4)2SO4 pptn., the enzyme reached max. specific activity. A reliable pH-stat assay in the presence of 2 μM dithiothreitol was devised.
- 6Carter, E. L.; Flugga, N.; Boer, J. L.; Mulrooney, S. B.; Hausinger, R. P. Interplay of Metal Ions and Urease. Metallomics 2009, 1, 207– 221, DOI: 10.1039/b903311dGoogle Scholar6Interplay of metal ions and ureaseCarter, Eric L.; Flugga, Nicholas; Boer, Jodi L.; Mulrooney, Scott B.; Hausinger, Robert P.Metallomics (2009), 1 (3), 207-221CODEN: METAJS; ISSN:1756-591X. (Royal Society of Chemistry)A review. Urease, the first enzyme to be crystd., contains a dinuclear nickel metallocenter that catalyzes the decompn. of urea to produce ammonia, a reaction of great agricultural and medical importance. Several mechanisms of urease catalysis have been proposed on the basis of enzyme crystal structures, model complexes, and computational efforts, but the precise steps in catalysis and the requirement of nickel vs. other metals remain unclear. Purified bacterial urease is partially activated via incubation with carbon dioxide plus nickel ions; however, in vitro activation also has been achieved with manganese and cobalt. In vivo activation of most ureases requires accessory proteins that function as nickel metallochaperones and GTP-dependent mol. chaperones or play other roles in the maturation process. In addn., some microorganisms control their levels of urease by metal ion-dependent regulatory mechanisms.
- 7Zambelli, B.; Musiani, F.; Benini, S.; Ciurli, S. Chemistry of Ni2+ in Urease: Sensing, Trafficking, and Catalysis. Acc. Chem. Res. 2011, 44, 520– 530, DOI: 10.1021/ar200041kGoogle Scholar7Chemistry of Ni2+ in urease: Sensing, trafficking, and catalysisZambelli, Barbara; Musiani, Francesco; Benini, Stefano; Ciurli, StefanoAccounts of Chemical Research (2011), 44 (7), 520-530CODEN: ACHRE4; ISSN:0001-4842. (American Chemical Society)A review. Transition metals are both essential to enzymic catalysis and limited in environmental availability. These 2 biol. facts have together driven organisms to evolve mechanisms for selective metal ion sensing and utilization. Changes in metal ion concns. are perceived by metal-dependent transcription factors and transduced into appropriate cellular responses, which regulate the machineries of competitive metal ion homeostasis and metalloenzyme activation. The intrinsic toxicity of the majority of metal ions further creates a need for regulated intracellular trafficking, which is carried out by specific chaperones. The Ni2+-dependent urease system serves as a paradigm for studying the strategies that cells use to handle an essential, yet toxic, metal ion. Although the discovery of urease as the 1st biol. system for which Ni2+ is essential for activity dates to 1975, the rationale for Ni2+ selection, as well as the cascade of events involving metal-dependent gene regulation and protein-protein interactions leading to enzyme activation, have yet to be fully unraveled. The past 14 years since the paper by P. A. Karplus et al. (1997) have witnessed impressive achievements in the understanding of the biochem. of Ni2+ in the urease system. Here, the authors discuss more recent advances in the comprehension of the specific role of Ni2+ in catalysis and in the interplay between Ni2+ and other metal ions, such as Zn2+ and Fe2+, in metal-dependent enzyme activity. The discussion focuses on work carried out in the authors' lab. In particular, the structural features of the enzyme bound to inhibitors, substrate analogs, and transition state or intermediate analogs have shed light on the catalytic mechanism. Structural and functional information has been correlated to understand the Ni2+ sensing effected by NikR protein, a Ni2+-dependent transcription factor. The urease activation process, involving the insertion of Ni2+ into the urease active site, has been in part dissected and analyzed through the investigation of the mol. properties of accessory proteins UreD, UreF, and UreG. The intracellular trafficking of Ni2+ has been rationalized through a deeper understanding of the structural and metal-binding properties of metallochaperone UreE. All the while, a no. of key general concepts have been revealed and developed. These include an understanding of the following: (1) the overall ancillary role of Zn2+ in Ni metab.; (2) the intrinsically disordered nature of the GTPase responsible for coupling the energy consumption to the CO2 requirement for the urease activation process; and (3) the role of the accessory proteins regulating this GTPase activity.
- 8Maroney, M. J.; Ciurli, S. Nonredox Nickel Enzymes. Chem. Rev. 2014, 114, 4206– 4228, DOI: 10.1021/cr4004488Google Scholar8Nonredox nickel enzymesMaroney, Michael J.; Ciurli, StefanoChemical Reviews (Washington, DC, United States) (2014), 114 (8), 4206-4228CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)A review. Current knowledge of the biochem., structure, and catalytic mechanism of enzymes, including urease, glyoxalase I, and acireductone dioxygenase, whose active sites require Ni and utilize it in a non-redox role, are discussed.
- 9Mazzei, L.; Musiani, F.; Ciurli, S. The Structure-based Reaction Mechanism of Urease, a Nickel Dependent Enzyme: Tale of a Long Debate. J. Biol. Inorg. Chem. 2020, 25, 829, DOI: 10.1007/s00775-020-01808-wGoogle Scholar9The structure-based reaction mechanism of urease, a nickel dependent enzyme: tale of a long debateMazzei, Luca; Musiani, Francesco; Ciurli, StefanoJBIC, Journal of Biological Inorganic Chemistry (2020), 25 (6), 829-845CODEN: JJBCFA; ISSN:0949-8257. (Springer)Abstr.: This review is an attempt to retrace the chronicle that starts from the discovery of the role of nickel as the essential metal ion in urease for the enzymic catalysis of urea, a key step in the biogeochem. cycle of nitrogen on Earth, to the most recent progress in understanding the chem. of this historical enzyme. Data and facts are presented through the magnifying lenses of the authors, using their best judgment to filter and elaborate on the many facets of the research carried out on this metalloenzyme over the years. The tale is divided in chapters that discuss and describe the results obtained in the subsequent leaps in the knowledge that led from the discovery of a biol. role for Ni to the most recent advancements in the comprehension of the relationship between the structure and function of urease. This review is intended not only to focus on the bioinorg. chem. of this beautiful metal-based catalysis, but also, and maybe primarily, to evoke inspiration and motivation to further explore the realm of bio-based coordination chem.
- 10Konieczna, I.; Zarnowiec, P.; Kwinkowski, M.; Kolesinska, B.; Fraczyk, J.; Kaminski, Z.; Kaca, W. Bacterial urease and its role in long-lasting human diseases. Curr. Protein Pept. Sci. 2012, 13, 789– 806, DOI: 10.2174/138920312804871094Google Scholar10Bacterial urease and its role in long-lasting human diseasesKonieczna, Iwona; Zarnowiec, Paulina; Kwinkowski, Marek; Kolesinska, Beata; Fraczyk, Justyna; Kaminski, Zbigniew; Kaca, WieslawCurrent Protein and Peptide Science (2012), 13 (8), 789-806CODEN: CPPSCM; ISSN:1389-2037. (Bentham Science Publishers Ltd.)A review. Urease is a virulence factor found in various pathogenic bacteria. It is essential in colonization of a host organism and in maintenance of bacterial cells in tissues. Due to its enzymic activity, urease has a toxic effect on human cells. The presence of ureolytic activity is an important marker of a no. of bacterial infections. Urease is also an immunogenic protein and is recognized by antibodies present in human sera. The presence of such antibodies is connected with progress of several long-lasting diseases, like rheumatoid arthritis, atherosclerosis or urinary tract infections. In bacterial ureases, motives with a sequence and/or structure similar to human proteins may occur. This phenomenon, known as mol. mimicry, leads to the appearance of autoantibodies, which take part in host mols. destruction. Detection of antibodies-binding motives (epitopes) in bacterial proteins is a complex process. However, org. chem. tools, such as synthetic peptide libraries, are helpful in both, epitope mapping as well as in serol. investigations. In this review, we present a synthetic report on a mol. organization of bacterial ureases - genetic as well as structural. We characterize methods used in detecting urease and ureolytic activity, including techniques applied in disease diagnostic processes and in chem. synthesis of urease epitopes. The review also provides a summary of knowledge about a toxic effect of bacterial ureases on human body and about occurrence of anti-urease antibodies in long-lasting diseases.
- 11Rajakovich, L. J.; Balskus, E. P. Metabolic Functions of the Human Gut Microbiota: the Role of Metalloenzymes. Nat. Prod. Rep. 2019, 36, 593– 625, DOI: 10.1039/C8NP00074CGoogle Scholar11Metabolic functions of the human gut microbiota: the role of metalloenzymesRajakovich, Lauren J.; Balskus, Emily P.Natural Product Reports (2019), 36 (4), 593-625CODEN: NPRRDF; ISSN:0265-0568. (Royal Society of Chemistry)Covering: up to the end of 2017The human body is composed of an equal no. of human and microbial cells. While the microbial community inhabiting the human gastrointestinal tract plays an essential role in host health, these organisms have also been connected to various diseases. Yet, the gut microbial functions that modulate host biol. are not well established. In this review, we describe metabolic functions of the human gut microbiota that involve metalloenzymes. These activities enable gut microbial colonization, mediate interactions with the host, and impact human health and disease. We highlight cases in which enzyme characterization has advanced our understanding of the gut microbiota and examples that illustrate the diverse ways in which metalloenzymes facilitate both essential and unique functions of this community. Finally, we analyze Human Microbiome Project sequencing datasets to assess the distribution of a prominent family of metalloenzymes in human-assocd. microbial communities, guiding future enzyme characterization efforts.
- 12Bremner, J. M.; Krogmeier, M. J. Evidence that the Adverse Effect of Urea Fertilizer on Seed Germination in Soil is due to Ammonia Formed through Hydrolysis of Urea by Soil Urease. Proc. Natl. Acad. Sci. U. S. A. 1989, 86, 8185– 8188, DOI: 10.1073/pnas.86.21.8185Google Scholar12Evidence that the adverse effect of urea fertilizer on seed germination in soil is due to ammonia formed through hydrolysis of urea by soil ureaseBremner, John M.; Krogmeier, Michael J.Proceedings of the National Academy of Sciences of the United States of America (1989), 86 (21), 8185-8CODEN: PNASA6; ISSN:0027-8424.Studies using seeds of wheat (Triticum aestivum), rye (Secale cereale), barley (Hordeum vulgare) and corn (Zea mays) indicated that the adverse effect of urea fertilizer on seed germination in soil is due to ammonia formed through hydrolysis of urea by soil urease and is not due to urea itself, to urea fertilizer impurities such as biuret, or to nitrite formed by nitrification of urea nitrogen. Support for this conclusion was obtained from (i) comparison of the effects on seed germination in soil of purified urea, urea fertilizers, urea fertilizer impurities, and compds. formed by enzymic and microbial transformations of urea in soil; (ii) studies showing that ammonia volatilized from soils treated with urea completely inhibited germination of seeds close to, but not in contact with, these soils; and (iii) expts. showing that the adverse effect of urea fertilizer on seed germination in soil was completely eliminated when the soil was autoclaved to destroy urease or was treated with phenylphosphorodiamidate to inhibit soil urease activity before treatment with urea fertilizer.
- 13Gioacchini, P.; Nastri, A.; Marzadori, C.; Giovannini, C.; Antisari, L. V.; Gessa, C. Influence of Urease and Nitrification Inhibitors on N Losses from Soils Fertilized with Urea. Biol. Fertil. 2002, 36, 129– 135, DOI: 10.1007/s00374-002-0521-1Google Scholar13Influence of urease and nitrification inhibitors on N losses from soils fertilized with ureaGioacchini, Paola; Nastri, Anna; Marzadori, Claudio; Giovannini, Camilla; Vittori Antisari, Livia; Gessa, CarloBiology and Fertility of Soils (2002), 36 (2), 129-135CODEN: BFSOEE; ISSN:0178-2762. (Springer-Verlag)The aim was to evaluate how the N losses through volatilization and leaching from soils fertilized with urea can be affected by the application of a urease inhibitor or a urease plus a nitrification inhibitor. The expt. was carried out using lysimeters with 15N-labeled urea and N-butylthiophosphoric triamide (NBPT) as urease inhibitor and dicyandiamide (DCD) as nitrification inhibitor, comparing 3 treatments: urea alone (U), urea + NBPT (UN) and urea + NBPT + DCD (UND). Both volatilization and leaching were significantly different in the soils used, according to their physicochem. characteristics. However, the pattern of the loss was similar: the volatilization was significantly reduced by NBPT (UN), but the presence of DCD (UND) significantly increased the loss, with respect to UN. Considering leaching, the highest amt. of NO3- was lost with UND, the lowest with U. The greatest amt. of N lost by leaching was soil-derived N produced by the N mineralization-immobilization turnover. We suggest that, by maintaining the NH4+ in the soils, the inhibitors, in particular DCD, caused a priming effect with a subsequent increase in the rate of soil org. matter mineralization and an extra release of soil org. N. The priming effect was real in the sandy loam (SL) soil where a net N release was obsd., whereas in the clay loam (CL) soil the effect of the inhibitors was less pronounced and an apparent priming effect was obsd.; however, a real priming effect also cannot be excluded in this soil.
- 14Jabri, E.; Carr, M. B.; Hausinger, R. P.; Karplus, P. A. The Crystal Structure of Urease from Klebsiella aerogenes. Science 1995, 268, 998– 1004, DOI: 10.1126/science.7754395Google Scholar14The crystal structure of urease from Klebsiella aerogenesJabri, Evelyn; Carr, Mary Beth; Hausinger, Robert P.; Karplus, P. AndrewScience (Washington, D. C.) (1995), 268 (5213), 998-1004CODEN: SCIEAS; ISSN:0036-8075. (American Association for the Advancement of Science)The crystal structure of urease from K. aerogenes was detd. at 2.2 Å resoln. and refined to an R factor of 18.2%. The enzyme contained 4 structural domains: 3 with novel folds playing structural roles, and an (αβ)8 barrel domain, which contained the bi-Ni center. The 2 active site Ni ions were 3.5 Å apart. One Ni ion was coordinated by 3 ligands (with low occupancy of a 4th ligand) and the 2nd was coordinated by 5 ligands. A carbamylated Lys residue provided an O ligand to each Ni, explaining why CO2 is required for the activation of urease apoenzyme. The structure was compatible with a catalytic mechanism whereby urea ligates Ni-1 to complete its tetrahedral coordination and a hydroxide ligand of Ni-2 attacks the carbonyl C atom. A surprisingly high structural similarity between the urease catalytic domain and that of the Zn-dependent adenosine deaminase revealed a remarkable example of active site divergence.
- 15Benini, S.; Rypniewski, W. R.; Wilson, K. S.; Miletti, S.; Ciurli, S.; Mangani, S. A New Proposal for Urease Mechanism Based on the Crystal Structures of the Native and Inhibited Enzyme from Bacillus pasteurii: Why Urea Hydrolysis Costs Two Nickels. Structure 1999, 7, 205– 216, DOI: 10.1016/S0969-2126(99)80026-4Google Scholar15A new proposal for urease mechanism based on the crystal structures of the native and inhibited enzyme from Bacillus pasteurii: why urea hydrolysis costs two nickelsBenini, Stefano; Rypniewski, Wojciech R.; Wilson, Keith S.; Miletti, Silvia; Ciurli, Stefano; Mangani, StefanoStructure (London) (1999), 7 (2), 205-216CODEN: STRUE6; ISSN:0969-2126. (Current Biology Publications)Urease catalyzes the hydrolysis of urea, the final step of org. nitrogen mineralization, using a bimetallic nickel center. The role of the active site metal ions and amino acid residues has not been elucidated to date. Many pathologies are assocd. with the activity of ureolytic bacteria, and the efficiency of soil nitrogen fertilization with urea is severely decreased by urease activity. Therefore, the development of urease inhibitors would lead to a redn. of environmental pollution, to enhanced efficiency of nitrogen uptake by plants, and to improved therapeutic strategies for treatment of infections due to ureolytic bacteria. Structure-based design of urease inhibitors would require knowledge of the enzyme mechanism at the mol. level. The structures of native and inhibited urease from Bacillus pasteurii have been detd. at a resoln. of 2.0 Å by synchroton X-ray cryogenic crystallog. In the native enzyme, the coordination sphere of each of the two nickel ions is completed by a water mol. and a bridging hydroxide. A fourth water mol. completes a tetrahedral cluster of solvent mols. The enzyme crystd. in the presence of phenylphosphorodiamidate contains the tetrahedral transition-state analog diamidophosphoric acid, bound to the two nickel ions in an unprecedented mode. Comparison of the native and inhibited structures reveals two distinct conformations of the flap lining the active-site cavity. The mode of binding of the inhibitor, and a comparison between the native and inhibited urease structures, indicate a novel mechanism for enzymic urea hydrolysis which reconciles the available structural and biochem. data.
- 16Ha, N.-C.; Oh, S.-T.; Sung, J. Y.; Cha, K. A.; Lee, M. H.; Oh, B.-H. Supramolecular Assembly and Acid Resistance of Helicobacter pylori Urease. Nat. Struct. Biol. 2001, 8, 505– 509, DOI: 10.1038/88563Google Scholar16Supramolecular assembly and acid resistance of Helicobacter pylori ureaseHa, Nam-Chul; Oh, Sang-Taek; Sung, Jae Young; Cha, Kyeung Ah; Lee, Mann Hyung; Oh, Byung-HaNature Structural Biology (2001), 8 (6), 505-509CODEN: NSBIEW; ISSN:1072-8368. (Nature America Inc.)H. pylori, an etiol. agent in a variety of gastro-duodenal diseases, produces a large amt. of urease (I), which is believed to neutralize gastric acid by producing NH3 for the survival of the bacteria. Up to 30% of I assocs. with the surface of intact cells upon lysis of neighboring bacteria. The role of I at the extracellular location has been a subject of controversy because the purified enzyme is irreversibly inactivated below pH 5. Here, the authors detd. the crystal structure of H. pylori I, which had a 1.1 MDa spherical assembly of 12 catalytic units with an outer diam. of ∼160 Å. Under physiol. relevant conditions, the activity of the enzyme remains unaffected down to pH 3. Activity assays under different conditions indicated that the duster of the 12 active sites on the supramol. assembly may be crit. for the survival of the enzyme at low pH. The structure provides a novel example of a mol. assembly adapted for acid resistance that, together with the low Km of the I, is likely to enable the organism to inhabit the hostile niche.
- 17Phillips, K.; Munster, D. J.; Allardyce, R. A.; Bagshaw, P. F. Antibacterial Action of the Urease Inhibitor Acetohydroxamic acid on Helicobacter pylori. J. Clin. Pathol. 1993, 46, 372– 373, DOI: 10.1136/jcp.46.4.372Google Scholar17Antibacterial action of the urease inhibitor acetohydroxamic acid on Helicobacter pyloriPhillips K; Munster D J; Allardyce R A; Bagshaw P FJournal of clinical pathology (1993), 46 (4), 372-3 ISSN:0021-9746.The urease inhibitor acetohydroxamic acid (AHA) was assessed for its bacteriostatic and bactericidal effects on Helicobacter pylori. For eight isolates of H pylori, the minimum inhibitory concentration (MIC) was either 200 mg/l or 400 mg/l. Interactions between AHA and antimicrobial drugs used to treat H pylori were also determined. For most isolates AHA reduced the MIC for colloidal bismuth subcitrate (CBS), tetracycline, metronidazole, and amoxicillin. In a few isolates, however, AHA increased the minimum bactericidal concentration (MBC) for these antimicrobial treatments. In vitro AHA is active against H pylori and it interacts with other agents directed against H pylori.
- 18Benini, S.; Rypniewski, W. R.; Wilson, K. S.; Miletti, S.; Ciurli, S.; Mangani, S. The Complex of Bacillus pasteurii Urease with Acetohydroxamate Anion from X-ray Data at 1.55 Å Resolution. J. Biol. Inorg. Chem. 2000, 5, 110– 118, DOI: 10.1007/s007750050014Google Scholar18The complex of Bacillus pasteurii urease with acetohydroxamate anion from x-ray data at 1.55 Å resolutionBenini, Stefano; Rypniewski, Wojciech R.; Wilson, Keith S.; Miletti, Silvia; Ciurli, Stefano; Mangani, StefanoJBIC, Journal of Biological Inorganic Chemistry (2000), 5 (1), 110-118CODEN: JJBCFA; ISSN:0949-8257. (Springer-Verlag)The crystal structure of B. pasteurii urease inhibited by acetohydroxamic acid was solved and refined anisotropically using synchrotron x-ray cryogenic diffraction data (1.55 Å resoln.; 99.5% completeness; data redundancy = 26; R-factor = 15.1%, PDB code 4UBP). The 2 Ni2+ ions in the active site were sepd. by a distance of 3.53 Å. The structure clearly showed the binding mode of the inhibitor anion, sym. bridging the 2 Ni2+ ions in the active site through the hydroxamate O atom and chelating 1 Ni2+ ion through the carbonyl O atom. The flexible flap flanking the active site cavity was in the open conformation. The possible implications of the results on structure-based mol. design of new urease inhibitors were discussed.
- 19Benini, S.; Rypniewski, W.; Wilson, K.; Ciurli, S.; Mangani, S. Structure-based Rationalization of Urease Inhibition by Phosphate: Novel Insights into the Enzyme Mechanism. J. Biol. Inorg. Chem. 2001, 6, 778– 790, DOI: 10.1007/s007750100254Google Scholar19Structure-based rationalization of urease inhibition by phosphate: novel insights into the enzyme mechanismBenini, S.; Rypniewski, W. R.; Wilson, K. S.; Ciurli, S.; Mangani, S.JBIC, Journal of Biological Inorganic Chemistry (2001), 6 (8), 778-790CODEN: JJBCFA; ISSN:0949-8257. (Springer-Verlag)The structure of Bacillus pasteurii urease (BPU) inhibited with phosphate was solved and refined using synchrotron x-ray diffraction data from a vitrified crystal (1.85 Å resoln., 99.3% completeness, data redundancy 4.6, R-factor 17.3%, PDB code 6UBP). A distance of 3.5 Å separates the two Ni ions in the active site. The binding mode of the inhibitor involves the formation of four coordination bonds with the two Ni ions: one phosphate oxygen atom sym. bridges the two metal ions (1.9-2.0 Å), while two of the remaining phosphate oxygen atoms bind to the Ni atoms at 2.4 Å. The fourth phosphate oxygen is directed into the active site channel. Anal. of the H-bonding network around the bound inhibitor indicates that phosphate is bound as the H2PO4- anion, and that an addnl. proton is present on the Oδ2 atom of Aspα363, an active site residue involved in Ni coordination through Oδ1. The flexible flap flanking the active site cavity is in the open conformation. Anal. of the complex reveals why phosphate is a relatively weak inhibitor and why sulfate does not bind to the nickels in the active site. The implications of the results for the understanding of the urease catalytic mechanism are reviewed. A novel alternative for the proton donor is presented.
- 20Benini, S.; Rypniewski, W. R.; Wilson, K. S.; Mangani, S.; Ciurli, S. Molecular Details of Urease Inhibition by Boric Acid: Insights into the Catalytic Mechanism. J. Am. Chem. Soc. 2004, 126, 3714– 3715, DOI: 10.1021/ja049618pGoogle Scholar20Molecular Details of Urease Inhibition by Boric Acid: Insights into the Catalytic MechanismBenini, Stefano; Rypniewski, Wojciech R.; Wilson, Keith S.; Mangani, Stefano; Ciurli, StefanoJournal of the American Chemical Society (2004), 126 (12), 3714-3715CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)The structure of the complex of urease, a Ni-contg. metalloenzyme, with boric acid was detd. at 2.10 Å resoln. The complex shows the unprecedented binding mode of the competitive inhibitor to the dinuclear metal center, with the B(OH)3 mol. bridging the Ni ions and leaving in place the bridging hydroxide. Boric acid can be considered a substrate analog of urea, and the structure supports the proposal that the Ni-bridging hydroxide acts as the nucleophile in the enzymic process of urea hydrolysis.
- 21Benini, S.; Cianci, M.; Mazzei, L.; Ciurli, S. Fluoride inhibition of Sporosarcina pasteurii urease: structure and thermodynamics. J. Biol. Inorg. Chem. 2014, 19, 1243– 1261, DOI: 10.1007/s00775-014-1182-xGoogle Scholar21Fluoride inhibition of Sporosarcina pasteurii urease: structure and thermodynamicsBenini, Stefano; Cianci, Michele; Mazzei, Luca; Ciurli, StefanoJBIC, Journal of Biological Inorganic Chemistry (2014), 19 (8), 1243-1261CODEN: JJBCFA; ISSN:0949-8257. (Springer)Urease is a Ni-dependent enzyme and a virulence factor for ureolytic bacterial human pathogens, but it is also necessary to convert urea, the most worldwide-used fertilizer, into forms of nitrogen that can be taken up by crop plants. A strategy to control the activity of urease for medical and agricultural applications is to use enzyme inhibitors. F- is a known urease inhibitor, but the structural basis of its mode of inhibition is still undetd. Here, kinetic studies on the F--induced inhibition of urease from S. pasteurii, a widespread and highly ureolytic soil bacterium, were performed using isothermal titrn. calorimetry and revealed a mixed competitive and uncompetitive mechanism. The pH dependence of the Ki values, investigated in the pH 6.5-8.0 range, revealed a predominant uncompetitive mechanism that increased by increasing the pH, and a lesser competitive inhibition that increased by lowering the pH. Ten crystal structures of the enzyme were independently detd. using 5 crystals of the native form and 5 crystals of the protein crystd. in the presence of F-. The anal. of these structures revealed the presence of 2 F- anions coordinated to the Ni2+ ions in the active site, in terminal and bridging positions. The present study consistently supported an interaction of F- with the Ni centers in the urease active site in which one F- anion competitively bound to Ni2+ proposed to coordinate urea in the initial step of the catalytic mechanism, while another F- anion uncompetitively substituted the Ni2+-bridging hydroxide, blocking its nucleophilic attack on urea.
- 22Mazzei, L.; Cianci, M.; Musiani, F.; Lente, G.; Palombo, M.; Ciurli, S. Inactivation of Urease by Catechol: Kinetics and Structure. J. inorg. Biochem. 2017, 166, 182– 189, DOI: 10.1016/j.jinorgbio.2016.11.016Google Scholar22Inactivation of urease by catechol: Kinetics and structureMazzei, Luca; Cianci, Michele; Musiani, Francesco; Lente, Gabor; Palombo, Marta; Ciurli, StefanoJournal of Inorganic Biochemistry (2017), 166 (), 182-189CODEN: JIBIDJ; ISSN:0162-0134. (Elsevier)Urease is a Ni(II)-contg. enzyme that catalyzes the hydrolysis of urea to yield NH3 and carbamate at a rate 1015-fold higher than the uncatalyzed reaction. Urease is a virulence factor of several human pathogens, in addn. to decreasing the efficiency of soil org. nitrogen fertilization. Therefore, efficient urease inhibitors are actively sought. Here, the authors describe the characterization of the interaction between urease from Sporosarcina pasteurii (SPU) and Canavalia ensiformis (jack bean, JBU) with catechol, a model polyphenol. In particular, catechol irreversibly inactivated both SPU and JBU with a complex radical-based autocatalytic multistep mechanism. The crystal structure of the SPU-catechol complex, detd. at 1.50 Å resoln., revealed the structural details of the enzyme inhibition.
- 23Mazzei, L.; Cianci, M.; Contaldo, U.; Ciurli, S. Insights into Urease Inhibition by N-(n-Butyl) Phosphoric Triamide through an Integrated Structural and Kinetic Approach. J. Agric. Food Chem. 2019, 67, 2127– 2138, DOI: 10.1021/acs.jafc.8b04791Google Scholar23Insights into urease inhibition by N-(n-butyl) phosphoric triamide through an integrated structural and kinetic approachMazzei, Luca; Cianci, Michele; Contaldo, Umberto; Ciurli, StefanoJournal of Agricultural and Food Chemistry (2019), 67 (8), 2127-2138CODEN: JAFCAU; ISSN:0021-8561. (American Chemical Society)The nickel-dependent enzyme urease represents a neg. element for the efficiency of soil nitrogen fertilization as well as a virulence factor for a large no. of pathogenic and antibiotic-resistant bacteria. The development of ever more efficient urease inhibitors demands knowledge of their modes of action at the mol. level. N-(n-Butyl)-phosphoric triamide (NBPTO) is the oxo-deriv. of N-(n-butyl)-thiophosphoric triamide (NBPT), which is extensively employed in agriculture to increase the efficiency of urea-based fertilizers. The 1.45 Å resoln. structure of the enzyme-inhibitor complex obtained upon incubation of Sporosarcina pasteurii urease (SPU) with NBPTO shows the presence of diamido phosphoric acid (DAP), generated upon enzymic hydrolysis of NBPTO with the release of Bu amine. DAP is bound in a tridentate binding mode to the two Ni(II) ions in the active site of urease via two O atoms and an amide NH2 group, whereas the second amide group of DAP points away from the metal center into the active-site channel. The mobile flap modulating the size of the active-site cavity is found in a disordered closed-open conformation. A kinetic characterization of the NBPTO-based inhibition of both bacterial (SPU) and plant (Canavalia ensiformis or jack bean, JBU) ureases, carried out by calorimetric measurements, indicates the occurrence of a reversible slow-inhibition mode of action. The latter is characterized by a very small value of the equil. dissocn. const. of the urease-DAP complex caused, in turn, by the large rate const. for the formation of the enzyme-inhibitor complex. The much greater capability of NBPTO to inhibit urease, as compared with that of NBPT, is thus not caused by the presence of a P=O moiety vs. a P=S moiety, as previously suggested, but rather by the readiness of NBPTO to react with urease without the need to convert one of the P-NH2 amide moieties to its P-OH acid deriv., as in the case of NBPT. The latter process is indeed characterized by a very small equil. const. that reduces drastically the concn. of the active form of the inhibitor in the case of NBPT. This indicates that high-efficiency phosphoramide-based urease inhibitors must have at least one O atom bound to the central P atom in order for the mol. to efficiently and rapidly bind to the dinickel center of the enzyme.
- 24Mazzei, L.; Cianci, M.; Benini, S.; Ciurli, S. The Structure of the Elusive Urease-Urea Complex Unveils the Mechanism of a Paradigmatic Nickel-Dependent Enzyme. Angew. Chem., Int. Ed. 2019, 58, 7415– 7419, DOI: 10.1002/anie.201903565Google Scholar24The structure of the elusive urease-urea complex unveils the mechanism of a paradigmatic nickel-dependent enzymeMazzei, Luca; Cianci, Michele; Benini, Stefano; Ciurli, StefanoAngewandte Chemie, International Edition (2019), 58 (22), 7415-7419CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)Urease, the most efficient enzyme known, contains an essential dinuclear NiII cluster in the active site. It catalyzes the hydrolysis of urea, inducing a rapid pH increase that has neg. effects on human health and agriculture. Thus, the control of urease activity is of utmost importance in medical, pharmaceutical, and agro-environmental applications. All known urease inhibitors are either toxic or inefficient. The development of new and efficient chems. able to inhibit urease relies on the knowledge of all steps of the catalytic mechanism. The short (microseconds) lifetime of the urease-urea complex has hampered the detn. of its structure. The present study uses fluoride to substitute the hydroxide acting as the co-substrate in the reaction, preventing the occurrence of the catalytic steps that follow substrate binding. The 1.42 Å crystal structure of the urease-urea complex, reported here, resolves the enduring debate on the mechanism of this metalloenzyme.
- 25Karplus, P. A.; Pearson, M. A.; Hausinger, R. P. 70 Years of Crystalline Urease: What Have We Learned?. Acc. Chem. Res. 1997, 30, 330– 337, DOI: 10.1021/ar960022jGoogle Scholar2570 Years of crystalline urease: What have we learned?Karplus, P. Andrew; Pearson, Matthew; Hausinger, Robert P.Accounts of Chemical Research (1997), 30 (8), 330-337CODEN: ACHRE4; ISSN:0001-4842. (American Chemical Society)A review with 60 refs. on the functional and structural characteristics of Klebsiella aerogenes urease. Crystallog. and functional studies of wild-type urease and some site-directed variants suggest a detailed catalytic mechanism which accounts for its kinetic properties and its structure. This mechanism assigns the limiting pKa's of 6.4 and 9.0 to the general acid His-320 and the hydrolytic water, resp. This assignment is such that only a small fraction of the urease mols. are present in the correct protonation state for activity. Obsd. disorder of the solvation in the active site and the comparison of multiple structures provide evidence for suboptimal interactions in the water-filled active site. These suboptimal interactions may lead to an enthalpy-driven free energy gain when this water is released into the bulk solvent upon urea binding. As such, the relaxation of the high-energy state could provide a large amt. of binding energy required for catalysis.
- 26Pearson, M. A.; Park, I.-S.; Schaller, R. A.; Michel, L. O.; Karplus, P. A.; Hausinger, R. P. Kinetic and Structural Characterization of Urease Active Site Variants. Biochemistry 2000, 39, 8575– 8584, DOI: 10.1021/bi000613oGoogle Scholar26Kinetic and Structural Characterization of Urease Active Site VariantsPearson, Matthew A.; Park, Il-Seon; Schaller, Ruth A.; Michel, Linda O.; Karplus, P. Andrew; Hausinger, Robert P.Biochemistry (2000), 39 (29), 8575-8584CODEN: BICHAW; ISSN:0006-2960. (American Chemical Society)Klebsiella aerogenes urease uses a dinuclear nickel active site to catalyze urea hydrolysis at >1014-fold the spontaneous rate. To better define the enzyme mechanism, we examd. the kinetics and structures for a suite of site-directed variants involving four residues at the active site: His320, His219, Asp221, and Arg336. Compared to wild-type urease, the H320A, H320N, and H320Q variants exhibit similar ∼10-5-fold deficiencies in rates, modest Km changes, and disorders in the peptide flap covering their active sites. H219A urease exhibits 103-fold increased Km over that of native enzyme, whereas the increase is less marked (∼102-fold) in the H219N and H219Q variants that retain hydrogen bonding capability. Structures for these variants show clearly resolved active site water mols. covered by well-ordered peptide flaps. Whereas the D221N variant is only moderately affected compared to wild-type enzyme, D221A urease possesses low activity (∼10-3 that of native enzyme), a small increase in Km, and a pH 5 optimum. The crystal structure for D221A urease is reminiscent of the His320 variants. The R336Q enzyme has a ∼10-4-fold decreased catalytic rate with near-normal pH dependence and an unaffected Km. Phenylglyoxal inactivates the R336Q variant at over half the rate obsd. for native enzyme, demonstrating that modification of non-active-site arginines can eliminate activity, perhaps by affecting the peptide flap. Our data favor a mechanism in which His219 helps to polarize the substrate carbonyl group, a metal-bound terminal hydroxide or bridging oxo-dianion attacks urea to form a tetrahedral intermediate, and protonation occurs via the general acid His320 with Asp221 and Arg336 orienting and influencing the acidity of this residue. Furthermore, we conclude that the simple bell-shaped pH dependence of kcat and kcat/Km for the native enzyme masks a more complex underlying pH dependence involving at least four pKas.
- 27Dixon, N. E.; Blakeley, R. L.; Zerner, B. Jack Bean Urease (EC 3.5.1.5). III. The Involvement of Active-site Nickel Ion in Inhibition by b-Mercaptoethanol, Phosphoramidate, and Fluoride. Can. J. Biochem. 1980, 58, 481– 488, DOI: 10.1139/o80-064Google Scholar27Jack bean urease (EC 3.5.1.5). III. The involvement of active-site nickel ion in inhibition by β-mercaptoethanol, phosphoramidate, and fluorideDixon, Nicholas E.; Blakeley, Robert L.; Zerner, BurtCanadian Journal of Biochemistry (1980), 58 (6), 481-8CODEN: CJBIAE; ISSN:0008-4018.Interaction of β-mercaptoethanol with urease produces large, rapid, and fully reversible spectral changes in that part of the electronic absorption spectrum which is assocd. with the tightly bound Ni2+. The spectrophotometrically detd. value of the dissocn. const. of the β-mercaptoethanol-urease complex (0.95 mM at pH 7.12 and 25°) is in agreement with the Ki (0.72 mM) for β-mercaptoethanol acting as a competitive inhibitor in the hydrolysis of urea. This constitutes direct evidence that the Ni2+ in jack bean urease is at the active site. Inhibition of urease by phosphoramidate is slowly achieved and slowly reversed, and upon reactivation of the isolated phosphoramidate-urease complex, phosphoramidate is regenerated in good yield. Spectrophotometric expts. indicate that phosphoramidate binds to Ni2+ in urease. Competition with β-mercaptoethanol was used to det. a dissocn. const. (1.23 mM at pH 7.12 and 25°) for a F--urease complex in which F- also coordinates with an active-site Ni2+. Kinetic evidence is presented which indicates that in the presence of urea, a ternary complex (F-urea-urease) is formed.
- 28Pelmenschikov, V.; Siegbahn, P. E. Nickel Superoxide Dismutase Reaction Mechanism Studied by Hybrid Density Functional Methods. J. Am. Chem. Soc. 2006, 128, 7466– 7475, DOI: 10.1021/ja053665fGoogle Scholar28Nickel Superoxide Dismutase Reaction Mechanism Studied by Hybrid Density Functional MethodsPelmenschikov, Vladimir; Siegbahn, Per E. M.Journal of the American Chemical Society (2006), 128 (23), 7466-7475CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)The reaction mechanism for the disproportionation of the toxic superoxide radical to mol. oxygen and hydrogen peroxide by the nickel-dependent superoxide dismutase (NiSOD) has been studied using the B3LYP hybrid DFT method. Based on the recent x-ray structures of the enzyme in the resting oxidized Ni(III) and x-ray-reduced Ni(II) states, the model investigated includes the backbone spacer of six residues (sequence nos. 1-6) as a structural framework. The side chains of residues His-1, Cys-2, and Cys-6, which are essential for nickel binding and catalysis, were modeled explicitly. The catalytic cycle consists of two half-reactions, each initiated by the successive substrate approach to the metal center. The two protons necessary for the dismutation are postulated to be delivered concertedly with the superoxide radical anions. The first (reductive) phase involves Ni(III) redn. to Ni(II), and the second (oxidative) phase involves the metal reoxidn. back to its resting state. The Cys-2 thiolate sulfur serves as a transient protonation site in the interim between the two half-reactions, allowing for the dioxygen and hydrogen peroxide mols. to be released in the reductive and oxidative phases, resp. The His-1 side chain nitrogen and backbone amides of the active site channel are shown to be less favorable transient proton locations, as compared to the Cys-2 sulfur. Comparisons are made to the Cu- and Zn-dependent SOD, studied previously using similar models.
- 29Sparta, M.; Valdez, C. E.; Alexandrova, A. N. Metal-dependent Activity of Fe and Ni Acireductone Dioxygenases: How Two Electrons Reroute the Catalytic Pathway. J. Mol. Biol. 2013, 425, 3007– 3018, DOI: 10.1016/j.jmb.2013.05.001Google Scholar29Metal-Dependent Activity of Fe and Ni Acireductone Dioxygenases: How Two Electrons Reroute the Catalytic PathwaySparta, Manuel; Valdez, Crystal E.; Alexandrova, Anastassia N.Journal of Molecular Biology (2013), 425 (16), 3007-3018CODEN: JMOBAK; ISSN:0022-2836. (Elsevier Ltd.)Two virtually identical acireductone dioxygenases, ARD and ARD', catalyze completely different oxidn. reactions of the same substrate, 1,2-dihydroxy-3-keto-5-(methylthio)pentene, depending exclusively on the nature of the bound metal. Fe2+-dependent ARD' produces the α-keto acid precursor of methionine and formate and allows for the recycling of methionine in cells. Ni2+-dependent ARD instead produces methylthiopropionate, CO, and formate, and exits the methionine salvage cycle. This mechanistic difference has not been understood to date but has been speculated to be due to the difference in coordination of the substrate to Fe2+ vs. Ni2+: forming a five-membered ring vs. a six-membered ring, resp., thus exposing different carbon atoms for the attack by O2. Here, using mixed quantum-classical mol. dynamics simulations followed by the d. functional theory mechanistic investigation, we show that contrary to the old hypothesis, both metals preferentially bind the substrate as a six-membered ring, exposing the exact same sites to the attack by O2. It is the electronic properties of the metals that are then responsible for the system following different reaction paths, to yield the resp. products. We fully explain the puzzling metal-induced difference in functionality between ARD and ARD' and, in particular, propose a new mechanism for ARD'. All results are in agreement with available isotopic substitution and other exptl. data.
- 30Wang, W.-J.; Wei, W.-J.; Liao, R.-Z. Deciphering the Chemoselectivity of Nickel-dependent Quercetin 2, 4-Dioxygenase. Phys. Chem. Chem. Phys. 2018, 20, 15784– 15794, DOI: 10.1039/C8CP02683AGoogle Scholar30Deciphering the chemoselectivity of nickel-dependent quercetin 2,4-dioxygenaseWang, Wen-Juan; Wei, Wen-Jie; Liao, Rong-ZhenPhysical Chemistry Chemical Physics (2018), 20 (23), 15784-15794CODEN: PPCPFQ; ISSN:1463-9076. (Royal Society of Chemistry)The reaction mechanism and chemoselectivity of nickel-dependent quercetin 2,4-dioxygenase (2,4-QueD) have been investigated using the QM/MM approach. The protonation state of the Glu74 residue, a first-shell ligand of Ni, has been considered to be either neutral or deprotonated. QM/MM calcns. predict that Glu74 must be deprotonated to rationalize the chemoselectivity and steer the 2,4-dioxygenolytic cleavage of quercetin, which harvests the exptl.-obsd. product, 2-protocatechuoylphloroglucinol carboxylic acid, coupled with the release of carbon monoxide. If the enzyme has a neutral Glu74 residue, the undesired 2,3-dioxygenolytic cleavage of quercetin becomes the dominant pathway, leading to the formation of α-keto acid. The calcns. suggest that the reaction takes place via three major steps: (1) attack of the superoxide on the C2 of the substrate pyrone ring to generate a NiII-peroxide intermediate; (2) formation of the second C-O bond between C4 and the peroxide to produce a peroxide bridge; (3) simultaneous cleavage of the C2-C3, C3-C4, and O1-O2 bonds with the formation of 2-protocatechuoylphloroglucinol carboxylic acid and carbon monoxide. The third step was found to be rate-limiting, with a barrier of 17.4 kcal mol-1, which is in very good agreement with the exptl. kinetic data. For the second C-O bond formation, an alternative pathway is that the peroxide attacks the C3 of the substrate pyrone ring, leading to the formation of a four-membered ring intermediate, which then undergoes concerted C2-C3 and O1-O2 bond cleavages to produce an α-keto acid. This pathway is assocd. with a barrier of 30.6 kcal mol-1, which is much higher than the major pathway. When Glu74 is protonated, the 2,3-dioxygenolytic pathway, however, has a lower barrier (21.8 kcal mol-1) than the 2,4-dioxygenolytic pathway.
- 31Siegbahn, P. E.; Chen, S.-L.; Liao, R.-Z. Theoretical Studies of Nickel-dependent Enzymes. Inorganics 2019, 7, 95, DOI: 10.3390/inorganics7080095Google Scholar31Theoretical studies of nickel-dependent enzymesSiegbahn, Per E. M.; Chen, Shi-Lu; Liao, Rong-ZhenInorganics (2019), 7 (8), 95CODEN: INORCW; ISSN:2304-6740. (MDPI AG)The advancements of quantum chem. methods and computer power allow detailed mechanistic investigations of metalloenzymes. In particular, both quantum chem. cluster and combined QM/MM approaches have been used, which have been proven to successfully complement exptl. studies. This review starts with a brief introduction of nickel-dependent enzymes and then summarizes theor. studies on the reaction mechanisms of these enzymes, including NiFe hydrogenase, methyl-coenzyme M reductase, nickel CO dehydrogenase, acetyl CoA synthase, acireductone dioxygenase, quercetin 2,4-dioxygenase, urease, lactate racemase, and superoxide dismutase.
- 32Suárez, D.; Díaz, N.; Merz, K. M. Ureases: Quantum Chemical Calculations on Cluster Models. J. Am. Chem. Soc. 2003, 125, 15324– 15337, DOI: 10.1021/ja030145gGoogle Scholar32Ureases: Quantum Chemical Calculations on Cluster ModelsSuarez, Dimas; Diaz, Natalia; Merz, Kenneth M., Jr.Journal of the American Chemical Society (2003), 125 (50), 15324-15337CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)Herein, the authors present results from a computational study of dinickel complexes that are relevant to the catalytic hydrolysis of urea exerted by the urease enzymes. The B3LYP d. functional is used to characterize the equil. geometry, electronic and magnetic properties, and energies for a series of realistic complexes modeling the active site of ureases. The anal. of the theor. results gives new insight into the structure, substrate binding, and catalytic mechanism. The water bridge between the two Ni(II) ions obsd. in the crystallog. structures of the ureases was assigned to a hydroxide bridge in agreement with the obsd. small antiferromagnetic coupling. Both monodentate and bidentate urea-bound complexes, in which urea had favorable orientations for catalysis, were characterized. Finally, two reaction mechanisms were investigated starting from the monodentate and bidentate urea-bound complexes, resp. Both a Ni1···Ni2 bridging hydroxide and a Ni2-bound water mol. play crucial roles in the two mechanisms.
- 33Estiu, G.; Merz, K. M. Competitive Hydrolytic and Elimination Mechanisms in the Urease Catalyzed Decomposition of Urea. J. Phys. Chem. B 2007, 111, 10263– 10274, DOI: 10.1021/jp072323oGoogle Scholar33Competitive Hydrolytic and Elimination Mechanisms in the Urease Catalyzed Decomposition of UreaEstiu, Guillermina; Merz, Kenneth M., Jr.Journal of Physical Chemistry B (2007), 111 (34), 10263-10274CODEN: JPCBFK; ISSN:1520-6106. (American Chemical Society)We present a high-level quantum chem. study of possible elimination reaction mechanisms assocd. with the catalytic decompn. of urea at the binuclear nickel active site cluster of urease. Stable intermediates and transition state structures have been identified along several possible reaction pathways. The computed results are compared with those reported by Suarez et al. for the hydrolytic catalyzed decompn. On the basis of these comparative studies, we propose a monodentate coordination of urea in the active site from which both the elimination and hydrolytic pathways can decomp. urea into CO2 and NH3. This observation is counter to what has been exptl. suggested based on the exogenous observation of carbamic acid (the reaction product from the hydrolysis pathway). However, this does not address what has occurred at the active site of urease prior to product release. On the basis of our computed results, the observation that urea prefers the elimination channel in aq. soln. and on the observation of Lippard and co-workers of an elimination reaction channel in a urease biomimetic model, we propose that the elimination channel needs to be re-examd. as a viable reaction channel in urease.
- 34Carlsson, H.; Nordlander, E. Computational Modeling of the Mechanism of Urease. Bioinorg. Chem. Appl. 2010, 2010, 364891, DOI: 10.1155/2010/364891Google ScholarThere is no corresponding record for this reference.
- 35Barrios, A. M.; Lippard, S. J. Interaction of Urea with a Hydroxide-Bridged Dinuclear Nickel Center: An Alternative Model for the Mechanism of Urease. J. Am. Chem. Soc. 2000, 122, 9172– 9177, DOI: 10.1021/ja000202vGoogle Scholar35Interaction of Urea with a Hydroxide-Bridged Dinuclear Nickel Center: An Alternative Model for the Mechanism of UreaseBarrios, Amy M.; Lippard, Stephen J.Journal of the American Chemical Society (2000), 122 (38), 9172-9177CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)A hydroxide-bridged dinuclear nickel complex with a urea mol. linking the two metal ions through its carbonyl oxygen atom has been prepd. as a model for the metalloenzyme urease. This complex, [Ni2(μ-OH)(μ-urea)(bdptz)(urea)(CH3CN)](ClO4)3, where bdptz is the dinucleating ligand 1,4-bis(2,2'-dipyridylmethyl)phthalazine, effects the hydrolysis of urea upon heating in a two-step reaction. In the first step, a mol. of ammonia is eliminated from urea with concomitant prodn. of cyanate, the first-order rate const. in acetonitrile being (7.7 ± 0.5) × 10-4 h-1. This reaction is at least 500 times faster than the spontaneous decompn. of urea under the same conditions. When the cyanate-contg. product is further heated in the presence of water, the cyanate is hydrolyzed with a second-order rate const. of (9.5 ± 1) × 10-4 M-1 h-1. Reaction of [Ni2(μ-OH)(μ-urea)(bdptz)(urea)(CH3CN)](ClO4)3 in 50% aq. acetonitrile afforded ammonia with no appreciable buildup of the cyanate-contg. species. A possible analog of the cyanate-contg. product, [Ni2(μ-OH)(μ-H2O)(bdptz)(μ-OCN)]2(OTs)4, was independently synthesized and structurally characterized. These results establish the precedence for hydrolysis of urea via a cyanate intermediate as an alternative mechanism for the urease-catalyzed hydrolysis of urea.
- 36Senn, H. M.; Thiel, W. QM/MM Methods for Biomolecular Systems. Angew. Chem., Int. Ed. 2009, 48, 1198– 1229, DOI: 10.1002/anie.200802019Google Scholar36QM/MM methods for biomolecular systemsSenn, Hans Martin; Thiel, WalterAngewandte Chemie, International Edition (2009), 48 (7), 1198-1229CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)A review. Combined quantum-mechanics/mol.-mechanics (QM/MM) approaches have become the method of choice for modeling reactions in biomol. systems. Quantum-mech. (QM) methods are required for describing chem. reactions and other electronic processes, such as charge transfer or electronic excitation. However, QM methods are restricted to systems of up to a few hundred atoms. However, the size and conformational complexity of biopolymers calls for methods capable of treating up to several 100,000 atoms and allowing for simulations over time scales of tens of nanoseconds. This is achieved by highly efficient, force-field-based mol. mechanics (MM) methods. Thus to model large biomols. the logical approach is to combine the two techniques and, to use a QM method for the chem. active region (e.g., substrates and co-factors in an enzymic reaction) and an MM treatment for the surroundings (e.g., protein and solvent). The resulting schemes are commonly referred to as combined or hybrid QM/MM methods. They enable the modeling of reactive biomol. systems at a reasonable computational effort while providing the necessary accuracy.
- 37Ahmadi, S.; Barrios Herrera, L.; Chehelamirani, M.; Hostaš, J.; Jalife, S.; Salahub, D. R. Multiscale Modeling of Enzymes: QM-cluster, QM/MM, and QM/MM/MD: A Tutorial Review. Int. J. Quantum Chem. 2018, 118, e25558, DOI: 10.1002/qua.25558Google ScholarThere is no corresponding record for this reference.
- 38Neese, F. The ORCA Program System. WIREs Comput. Mol. Sci. 2012, 2, 73– 78, DOI: 10.1002/wcms.81Google Scholar38The ORCA program systemNeese, FrankWiley Interdisciplinary Reviews: Computational Molecular Science (2012), 2 (1), 73-78CODEN: WIRCAH; ISSN:1759-0884. (Wiley-Blackwell)A review. ORCA is a general-purpose quantum chem. program package that features virtually all modern electronic structure methods (d. functional theory, many-body perturbation and coupled cluster theories, and multireference and semiempirical methods). It is designed with the aim of generality, extendibility, efficiency, and user friendliness. Its main field of application is larger mols., transition metal complexes, and their spectroscopic properties. ORCA uses std. Gaussian basis functions and is fully parallelized. The article provides an overview of its current possibilities and documents its efficiency.
- 39Becke, A. D. Density-functional Exchange-energy Approximation with Correct Asymptotic Behavior. Phys. Rev. A 1988, 38, 3098– 3100, DOI: 10.1103/PhysRevA.38.3098Google Scholar39Density-functional exchange-energy approximation with correct asymptotic behaviorBecke, A. D.Physical Review A: Atomic, Molecular, and Optical Physics (1988), 38 (6), 3098-100CODEN: PLRAAN; ISSN:0556-2791.Current gradient-cor. d.-functional approxns. for the exchange energies of at. and mol. systems fail to reproduce the correct 1/r asymptotic behavior of the exchange-energy d. A gradient-cor. exchange-energy functional is given with the proper asymptotic limit. This functional, contg. only one parameter, fits the exact Hartree-Fock exchange energies of a wide variety of at. systems with remarkable accuracy, surpassing the performance of previous functionals contg. two parameters or more.
- 40Lee, C.; Yang, W.; Parr, R. G. Development of the Colle-Salvetti Correlation-energy Formula into a Functional of the Electron Density. Phys. Rev. B 1988, 37, 785– 789, DOI: 10.1103/PhysRevB.37.785Google Scholar40Development of the Colle-Salvetti correlation-energy formula into a functional of the electron densityLee, Chengteh; Yang, Weitao; Parr, Robert G.Physical Review B: Condensed Matter and Materials Physics (1988), 37 (2), 785-9CODEN: PRBMDO; ISSN:0163-1829.A correlation-energy formula due to R. Colle and D. Salvetti (1975), in which the correlation energy d. is expressed in terms of the electron d. and a Laplacian of the 2nd-order Hartree-Fock d. matrix, is restated as a formula involving the d. and local kinetic-energy d. On insertion of gradient expansions for the local kinetic-energy d., d.-functional formulas for the correlation energy and correlation potential are then obtained. Through numerical calcns. on a no. of atoms, pos. ions, and mols., of both open- and closed-shell type, it is demonstrated that these formulas, like the original Colle-Salvetti formulas, give correlation energies within a few percent.
- 41Becke, A. D. Density-functional Thermochemistry. III. The Role of Exact Exchange. J. Chem. phys. 1993, 98, 5648– 5652, DOI: 10.1063/1.464913Google Scholar41Density-functional thermochemistry. III. The role of exact exchangeBecke, Axel D.Journal of Chemical Physics (1993), 98 (7), 5648-52CODEN: JCPSA6; ISSN:0021-9606.Despite the remarkable thermochem. accuracy of Kohn-Sham d.-functional theories with gradient corrections for exchange-correlation, the author believes that further improvements are unlikely unless exact-exchange information is considered. Arguments to support this view are presented, and a semiempirical exchange-correlation functional (contg. local-spin-d., gradient, and exact-exchange terms) is tested for 56 atomization energies, 42 ionization potentials, 8 proton affinities, and 10 total at. energies of first- and second-row systems. This functional performs better than previous functionals with gradient corrections only, and fits expt. atomization energies with an impressively small av. abs. deviation of 2.4 kcal/mol.
- 42Stephens, P. J.; Devlin, F. J.; Chabalowski, C. F.; Frisch, M. J. Ab Initio Calculation of Vibrational Absorption and Circular Dichroism Spectra Using Density Functional Force Fields. J. Phys. Chem. 1994, 98, 11623– 11627, DOI: 10.1021/j100096a001Google Scholar42Ab Initio Calculation of Vibrational Absorption and Circular Dichroism Spectra Using Density Functional Force FieldsStephens, P. J.; Devlin, F. J.; Chabalowski, C. F.; Frisch, M. J.Journal of Physical Chemistry (1994), 98 (45), 11623-7CODEN: JPCHAX; ISSN:0022-3654.The unpolarized absorption and CD spectra of the fundamental vibrational transitions of the chiral mol. 4-methyl-2-oxetanone are calcd. ab initio. Harmonic force fields are obtained using d. functional theory (DFT), MP2 and SCF methodologies, and a [5s4p2d/3s2p] (TZ2P) basis set. DFT calcns. use the LSDA, BLYP, and Becke3LYP (B3LYP) d. functionals. Mid-IR spectra predicted using LSDA, BLYP, and B3LYP force fields are of significantly different quality, the B3LYP force field yielding spectra in clearly superior, and overall excellent, agreement with expt. The MP2 force field yields spectra in slightly worse agreement with expt. than the B3LYP force field. The SCF force field yields spectra in poor agreement with expt. The basis set dependence of B3LYP force fields is also explored: the 6-31G* and TZ2P basis sets give very similar results while the 3-21G basis set yields spectra in substantially worse agreement with expt.
- 43Neese, F.; Wennmohs, F.; Hansen, A.; Becker, U. Efficient, Approximate and Parallel Hartree-Fock and Hybrid DFT Calculations. A ‘Chain-of-Spheres’ Algorithm for the Hartree-Fock Exchange. Chem. Phys. 2009, 356, 98– 109, DOI: 10.1016/j.chemphys.2008.10.036Google Scholar43Efficient, approximate and parallel Hartree-Fock and hybrid DFT calculations. A 'chain-of-spheres' algorithm for the Hartree-Fock exchangeNeese, Frank; Wennmohs, Frank; Hansen, Andreas; Becker, UteChemical Physics (2009), 356 (1-3), 98-109CODEN: CMPHC2; ISSN:0301-0104. (Elsevier B.V.)In this paper, the possibility is explored to speed up Hartree-Fock and hybrid d. functional calcns. by forming the Coulomb and exchange parts of the Fock matrix by different approxns. For the Coulomb part the previously introduced Split-RI-J variant of the well-known d. fitting' approxn. is used. The exchange part is formed by semi-numerical integration techniques that are closely related to Friesner's pioneering pseudo-spectral approach. Our potentially linear scaling realization of this algorithm is called the 'chain-of-spheres exchange' (COSX). A combination of semi-numerical integration and d. fitting is also proposed. Both Split-RI-J and COSX scale very well with the highest angular momentum in the basis sets. It is shown that for extended basis sets speed-ups of up to two orders of magnitude compared to traditional implementations can be obtained in this way. Total energies are reproduced with an av. error of <0.3 kcal/mol as detd. from extended test calcns. with various basis sets on a set of 26 mols. with 20-200 atoms and up to 2000 basis functions. Reaction energies agree to within 0.2 kcal/mol (Hartree-Fock) or 0.05 kcal/mol (hybrid DFT) with the canonical values. The COSX algorithm parallelizes with a speedup of 8.6 obsd. for 10 processes. Min. energy geometries differ by less than 0.3 pm in the bond distances and 0.5° in the bond angles from their canonical values. These developments enable highly efficient and accurate SCF calcns. including nonlocal Hartree-Fock exchange for large mols. In combination with the RI-MP2 method and large basis sets, second-order many body perturbation energies can be obtained for medium sized mols. with unprecedented efficiency. The algorithms are implemented into the ORCA electronic structure system.
Moving Frontiers in Quantum Chemistry.
- 44Weigend, F.; Ahlrichs, R. Balanced Basis Sets of Split Valence, Triple Zeta Valence and Quadruple Zeta Valence Quality for H to Rn: Design and Assessment of Accuracy. Phys. Chem. Chem. Phys. 2005, 7, 3297– 3305, DOI: 10.1039/b508541aGoogle Scholar44Balanced basis sets of split valence, triple zeta valence and quadruple zeta valence quality for H to Rn: Design and assessment of accuracyWeigend, Florian; Ahlrichs, ReinhartPhysical Chemistry Chemical Physics (2005), 7 (18), 3297-3305CODEN: PPCPFQ; ISSN:1463-9076. (Royal Society of Chemistry)Gaussian basis sets of quadruple zeta valence quality for Rb-Rn are presented, as well as bases of split valence and triple zeta valence quality for H-Rn. The latter were obtained by (partly) modifying bases developed previously. A large set of more than 300 mols. representing (nearly) all elements-except lanthanides-in their common oxidn. states was used to assess the quality of the bases all across the periodic table. Quantities investigated were atomization energies, dipole moments and structure parameters for Hartree-Fock, d. functional theory and correlated methods, for which we had chosen Moller-Plesset perturbation theory as an example. Finally recommendations are given which type of basis set is used best for a certain level of theory and a desired quality of results.
- 45Breneman, C. M.; Wiberg, K. B. Determining Atom-centered Monopoles from Molecular Electrostatic Potentials. The Need for High Sampling Density in Formamide Conformational Analysis. J. Comput. Chem. 1990, 11, 361– 373, DOI: 10.1002/jcc.540110311Google Scholar45Determining atom-centered monopoles from molecular electrostatic potentials. The need for high sampling density in formamide conformational analysisBreneman, Curt M.; Wiberg, Kenneth B.Journal of Computational Chemistry (1990), 11 (3), 361-73CODEN: JCCHDD; ISSN:0192-8651.An improved method for computing potential-derived charges is described which is based upon the CHELP program available from QCPE. This approach (CHELPG) is shown to be considerably less dependent upon mol. orientation than the original CHELP program. In the second part of this work, the CHELPG point selection algorithm was used to analyze the changes in the potential-derived charges in formamide during rotation about the C-N bond. In order to achieve a level of rotational invariance less than 10% of the magnitude of the electronic effects studied, an equally-spaced array of points 0.3 Å apart was required. Points found to be greater than 2.8 Å from any nucleus were eliminated, along with all points contained within the defined VDW distances from each of the atoms. The results are compared to those obtained by using CHELP. Even when large nos. of points (ca. 3000) were sampled using the CHELP selection routine, the results did not indicate a satisfactory level of rotational invariance. On the basis of these results, the original CHELP program was inadequate for analyzing internal rotations.
- 46Vanommeslaeghe, K.; Hatcher, E.; Acharya, C.; Kundu, S.; Zhong, S.; Shim, J.; Darian, E.; Guvench, O.; Lopes, P.; Vorobyov, I.; Mackerell, A. D., Jr. CHARMM General Force Field: A Force Field for Drug-like Molecules Compatible with the CHARMM All-atom Additive Biological Force Fields. J. Comput. Chem. 2010, 31, 671– 690, DOI: 10.1002/jcc.21367Google Scholar46CHARMM general force field: A force field for drug-like molecules compatible with the CHARMM all-atom additive biological force fieldsVanommeslaeghe, K.; Hatcher, E.; Acharya, C.; Kundu, S.; Zhong, S.; Shim, J.; Darian, E.; Guvench, O.; Lopes, P.; Vorobyov, I.; Mackerell, A. D., Jr.Journal of Computational Chemistry (2010), 31 (4), 671-690CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)The widely used CHARMM additive all-atom force field includes parameters for proteins, nucleic acids, lipids, and carbohydrates. In the present article, an extension of the CHARMM force field to drug-like mols. is presented. The resulting CHARMM General Force Field (CGenFF) covers a wide range of chem. groups present in biomols. and drug-like mols., including a large no. of heterocyclic scaffolds. The parametrization philosophy behind the force field focuses on quality at the expense of transferability, with the implementation concg. on an extensible force field. Statistics related to the quality of the parametrization with a focus on exptl. validation are presented. Addnl., the parametrization procedure, described fully in the present article in the context of the model systems, pyrrolidine, and 3-phenoxymethyl-pyrrolidine will allow users to readily extend the force field to chem. groups that are not explicitly covered in the force field as well as add functional groups to and link together mols. already available in the force field. CGenFF thus makes it possible to perform "all-CHARMM" simulations on drug-target interactions thereby extending the utility of CHARMM force fields to medicinally relevant systems. © 2009 Wiley Periodicals, Inc.J Comput Chem, 2010.
- 47Li, P.; Roberts, B. P.; Chakravorty, D. K.; Merz, K. M., Jr. Rational Design of Particle Mesh Ewald Compatible Lennard-Jones Parameters for + 2 Metal Cations in Explicit Solvent. J. Chem. Theory Comput. 2013, 9, 2733– 2748, DOI: 10.1021/ct400146wGoogle Scholar47Rational Design of Particle Mesh Ewald Compatible Lennard-Jones Parameters for +2 Metal Cations in Explicit SolventLi, Pengfei; Roberts, Benjamin P.; Chakravorty, Dhruva K.; Merz, Kenneth M.Journal of Chemical Theory and Computation (2013), 9 (6), 2733-2748CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)Metal ions play significant roles in biol. systems. Accurate mol. dynamics (MD) simulations on these systems require a validated set of parameters. Although there are more detailed ways to model metal ions, the nonbonded model, which employs a 12-6 Lennard-Jones (LJ) term plus an electrostatic potential, is still widely used in MD simulations today due to its simple form. However, LJ parameters have limited transferability due to different combining rules, various water models, and diverse simulation methods. Recently, simulations employing a Particle Mesh Ewald (PME) treatment for long-range electrostatics have become more and more popular owing to their speed and accuracy. In the present work, we have systematically designed LJ parameters for 24 +2 metal (M(II)) cations to reproduce different exptl. properties appropriate for the Lorentz-Berthelot combining rules and PME simulations. We began by testing the transferability of currently available M(II) ion LJ parameters. The results showed that there are differences between simulations employing Ewald summation with other simulation methods and that it was necessary to design new parameters specific for PME based simulations. Employing the thermodn. integration (TI) method and performing periodic boundary MD simulations employing PME, allowed for a systematic investigation of the LJ parameter space. Hydration free energies (HFEs), the ion-oxygen distance in the first solvation shell (IOD), and coordination nos. (CNs) were obtained for various combinations of the parameters of the LJ potential for four widely used water models (TIP3P, SPC/E, TIP4P, and TIP4PEW). Results showed that the three simulated properties were highly correlated. Meanwhile, M(II) ions with the same parameters in different water models produce remarkably different HFEs but similar structural properties. It is difficult to reproduce various exptl. values simultaneously because the nonbonded model underestimates the interaction between the metal ions and water mols. at short-range. Moreover, the extent of underestimation increases successively for the TIP3P, SPC/E, TIP4PEW, and TIP4P water models. Nonetheless, we fitted a curve to describe the relationship between ε (the well depth) and radius (Rmin/2) from exptl. data on noble gases to facilitate the generation of the best possible compromise models. Hence, by targeting different exptl. values, we developed three sets of parameters for M(II) cations for three different water models (TIP3P, SPC/E, and TIP4PEW). These parameters we feel represent the best possible compromise that can be achieved using the nonbonded model for the ions in combination with simple water models. From a computational uncertainty anal. we est. that the uncertainty in our computed HFEs is on the order of ±1 kcal/mol. Further improvements will require more advanced nonbonded models likely with inclusion of polarization.
- 48Lee, J.; Cheng, X.; Swails, J. M.; Yeom, M. S.; Eastman, P. K.; Lemkul, J. A.; Wei, S.; Buckner, J.; Jeong, J. C.; Qi, Y. CHARMM-GUI Input Generator for NAMD, GROMACS, AMBER, OpenMM, and CHARMM/OpenMM Simulations Using the CHARMM36 Additive Force Field. J. Chem. Theory Comput. 2016, 12, 405– 413, DOI: 10.1021/acs.jctc.5b00935Google Scholar48CHARMM-GUI Input Generator for NAMD, GROMACS, AMBER, OpenMM, and CHARMM/OpenMM Simulations Using the CHARMM36 Additive Force FieldLee, Jumin; Cheng, Xi; Swails, Jason M.; Yeom, Min Sun; Eastman, Peter K.; Lemkul, Justin A.; Wei, Shuai; Buckner, Joshua; Jeong, Jong Cheol; Qi, Yifei; Jo, Sunhwan; Pande, Vijay S.; Case, David A.; Brooks, Charles L.; MacKerell, Alexander D.; Klauda, Jeffery B.; Im, WonpilJournal of Chemical Theory and Computation (2016), 12 (1), 405-413CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)Proper treatment of nonbonded interactions is essential for the accuracy of mol. dynamics (MD) simulations, esp. in studies of lipid bilayers. The use of the CHARMM36 force field (C36 FF) in different MD simulation programs can result in disagreements with published simulations performed with CHARMM due to differences in the protocols used to treat the long-range and 1-4 nonbonded interactions. In this study, we systematically test the use of the C36 lipid FF in NAMD, GROMACS, AMBER, OpenMM, and CHARMM/OpenMM. A wide range of Lennard-Jones (LJ) cutoff schemes and integrator algorithms were tested to find the optimal simulation protocol to best match bilayer properties of six lipids with varying acyl chain satn. and head groups. MD simulations of a 1,2-dipalmitoyl-sn-phosphatidylcholine (DPPC) bilayer were used to obtain the optimal protocol for each program. MD simulations with all programs were found to reasonably match the DPPC bilayer properties (surface area per lipid, chain order parameters, and area compressibility modulus) obtained using the std. protocol used in CHARMM as well as from expts. The optimal simulation protocol was then applied to the other five lipid simulations and resulted in excellent agreement between results from most simulation programs as well as with exptl. data. AMBER compared least favorably with the expected membrane properties, which appears to be due to its use of the hard-truncation in the LJ potential vs. a force-based switching function used to smooth the LJ potential as it approaches the cutoff distance. The optimal simulation protocol for each program has been implemented in CHARMM-GUI. This protocol is expected to be applicable to the remainder of the additive C36 FF including the proteins, nucleic acids, carbohydrates, and small mols.
- 49Humphrey, W.; Dalke, A.; Schulten, K. VMD: Visual Molecular Dynamics. J. Mol. Graph. 1996, 14, 33– 38, DOI: 10.1016/0263-7855(96)00018-5Google Scholar49VDM: visual molecular dynamicsHumphrey, William; Dalke, Andrew; Schulten, KlausJournal of Molecular Graphics (1996), 14 (1), 33-8, plates, 27-28CODEN: JMGRDV; ISSN:0263-7855. (Elsevier)VMD is a mol. graphics program designed for the display and anal. of mol. assemblies, in particular, biopolymers such as proteins and nucleic acids. VMD can simultaneously display any no. of structures using a wide variety of rendering styles and coloring methods. Mols. are displayed as one or more "representations," in which each representation embodies a particular rendering method and coloring scheme for a selected subset of atoms. The atoms displayed in each representation are chosen using an extensive atom selection syntax, which includes Boolean operators and regular expressions. VMD provides a complete graphical user interface for program control, as well as a text interface using the Tcl embeddable parser to allow for complex scripts with variable substitution, control loops, and function calls. Full session logging is supported, which produces a VMD command script for later playback. High-resoln. raster images of displayed mols. may be produced by generating input scripts for use by a no. of photorealistic image-rendering applications. VMD has also been expressly designed with the ability to animate mol. dynamics (MD) simulation trajectories, imported either from files or from a direct connection to a running MD simulation. VMD is the visualization component of MDScope, a set of tools for interactive problem solving in structural biol., which also includes the parallel MD program NAMD, and the MDCOMM software used to connect the visualization and simulation programs, VMD is written in C++, using an object-oriented design; the program, including source code and extensive documentation, is freely available via anonymous ftp and through the World Wide Web.
- 50Park, I.-S.; Hausinger, R. P. Site-Directed Mutagenesis of Klebsiella aerogenes Urease: Identification of Histidine Residues that Appear to Function in Nickel Ligation, Substrate Binding, and Catalysis. Protein Sci. 1993, 2, 1034– 1041, DOI: 10.1002/pro.5560020616Google Scholar50Site-directed mutagenesis of Klebsiella aerogenes urease: Identification of histidine residues that appear to function in nickel ligation, substrate binding, and catalysisPark, Il Seon; Hausinger, Robert P.Protein Science (1993), 2 (6), 1034-41CODEN: PRCIEI; ISSN:0961-8368.Comparison of six urease sequences revealed the presence of 10 conserved histidine residues (H96 in the γ subunit, H39 and H41 in β, and H134, H136, H219, H246, H312, H320, and H321 in the α subunit of the K. aerogenes enzyme). Each of these residues in K. aerogenes urease was substituted with alanine by site-directed mutagenesis, and the mutant proteins were purified and characterized in order to identify essential histidine residues and assign their roles. The γH96A, βH39A, βH41A, αH312A, and αH321A mutant proteins possess activities and nickel contents similar to wild-type enzyme, suggesting that these residues are not essential for substrate binding, catalysis, or metal binding. In contrast, the αH134A, αH136A, and αH246A proteins exhibit no detectable activity and possess 53%, 6%, and 21% of the nickel content of wild-type enzyme. These results are consistent with αH134, αH136, and αH246 functioning as nickel ligands. The αH219A protein is active and has nickel (∼1.9% and ∼80%, resp., when compared to wild-type protein) but exhibits a very high Km value (1,100 mM compared to 2.3 mM for the wild-type enzyme). These results are compatible with αH219 having some role in facilitating substrate binding. Finally, the αH320A protein (Km = 8.3 mM) only displays ∼0.003% of the wild-type enzyme activity, despite having a normal nickel content. Unlike the wild-type and αH219A ureases, this mutant protein was not inactivated by diethylpyrocarbonate (DEP), consistent with αH320 being the DEP-reactive general base that facilitates catalysis.
- 51Nelson, M. T.; Humphrey, W.; Gursoy, A.; Dalke, A.; Kalé, L. V.; Skeel, R. D.; Schulten, K. NAMD: A Parallel, Object-Oriented Molecular Dynamics Program. Int. J. High Perform. Comput. Appl. 1996, 10, 251– 268, DOI: 10.1177/109434209601000401Google ScholarThere is no corresponding record for this reference.
- 52Darden, T.; York, D.; Pedersen, L. Particle Mesh Ewald: An N·log (N) Method for Ewald Sums in Large Systems. J. Chem. Phys. 1993, 98, 10089– 10092, DOI: 10.1063/1.464397Google Scholar52Particle mesh Ewald: an N·log(N) method for Ewald sums in large systemsDarden, Tom; York, Darrin; Pedersen, LeeJournal of Chemical Physics (1993), 98 (12), 10089-92CODEN: JCPSA6; ISSN:0021-9606.An N·log(N) method for evaluating electrostatic energies and forces of large periodic systems is presented. The method is based on interpolation of the reciprocal space Ewald sums and evaluation of the resulting convolution using fast Fourier transforms. Timings and accuracies are presented for three large cryst. ionic systems.
- 53Huang, J.; MacKerell, A. D., Jr CHARMM36 All-atom Additive Protein Force Field: Validation Based on Comparison to NMR Data. J. Comput. Chem. 2013, 34, 2135– 2145, DOI: 10.1002/jcc.23354Google Scholar53CHARMM36 all-atom additive protein force field: Validation based on comparison to NMR dataHuang, Jing; MacKerell, Alexander D. JrJournal of Computational Chemistry (2013), 34 (25), 2135-2145CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)Protein structure and dynamics can be characterized on the atomistic level with both NMR expts. and mol. dynamics (MD) simulations. Here, the authors quantify the ability of the recently presented CHARMM36 (C36) force field (FF) to reproduce various NMR observables using MD simulations. The studied NMR properties include backbone scalar couplings across hydrogen bonds, residual dipolar couplings (RDCs) and relaxation order parameter, as well as scalar couplings, RDCs, and order parameters for side-chain amino- and methyl-contg. groups. The C36 FF leads to better correlation with exptl. data compared to the CHARMM22/CMAP FF and suggest using C36 in protein simulations. Although both CHARMM FFs contains the same nonbond parameters, the authors' results show how the changes in the internal parameters assocd. with the peptide backbone via CMAP and the χ1 and χ2 dihedral parameters leads to improved treatment of the analyzed nonbond interactions. This highlights the importance of proper treatment of the internal covalent components in modeling nonbond interactions with mol. mechanics FFs.
- 54Bannwarth, C.; Ehlert, S.; Grimme, S. GFN2-xTB─An Accurate and Broadly Parametrized Self-Consistent Tight-Binding Quantum Chemical Method with Multipole Electrostatics and Density-Dependent Dispersion Contributions. J. Chem. Theory Comput. 2019, 15, 1652– 1671, DOI: 10.1021/acs.jctc.8b01176Google Scholar54GFN2-xTB-An Accurate and Broadly Parametrized Self-Consistent Tight-Binding Quantum Chemical Method with Multipole Electrostatics and Density-Dependent Dispersion ContributionsBannwarth, Christoph; Ehlert, Sebastian; Grimme, StefanJournal of Chemical Theory and Computation (2019), 15 (3), 1652-1671CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)An extended semiempirical tight-binding model is presented, which is primarily designed for the fast calcn. of structures and noncovalent interactions energies for mol. systems with roughly 1000 atoms. The essential novelty in this so-called GFN2-xTB method is the inclusion of anisotropic second order d. fluctuation effects via short-range damped interactions of cumulative at. multipole moments. Without noticeable increase in the computational demands, this results in a less empirical and overall more phys. sound method, which does not require any classical halogen or hydrogen bonding corrections and which relies solely on global and element-specific parameters (available up to radon, Z = 86). Moreover, the at. partial charge dependent D4 London dispersion model is incorporated self-consistently, which can be naturally obtained in a tight-binding picture from second order d. fluctuations. Fully anal. and numerically precise gradients (nuclear forces) are implemented. The accuracy of the method is benchmarked for a wide variety of systems and compared with other semiempirical methods. Along with excellent performance for the "target" properties, we also find lower errors for "off-target" properties such as barrier heights and mol. dipole moments. High computational efficiency along with the improved physics compared to it precursor GFN-xTB makes this method well-suited to explore the conformational space of mol. systems. Significant improvements are furthermore obsd. for various benchmark sets, which are prototypical for biomol. systems in aq. soln.
- 55Melo, M. C.; Bernardi, R. C.; Rudack, T.; Scheurer, M.; Riplinger, C.; Phillips, J. C.; Maia, J. D.; Rocha, G. B.; Ribeiro, J. V.; Stone, J. E. NAMD Goes Quantum: An Integrative Suite for Hybrid Simulations. Nat. Methods 2018, 15, 351– 354, DOI: 10.1038/nmeth.4638Google Scholar55NAMD goes quantum: an integrative suite for hybrid simulationsMelo, Marcelo C. R.; Bernardi, Rafael C.; Rudack, Till; Scheurer, Maximilian; Riplinger, Christoph; Phillips, James C.; Maia, Julio D. C.; Rocha, Gerd B.; Ribeiro, Joao V.; Stone, John E.; Neese, Frank; Schulten, Klaus; Luthey-Schulten, ZaidaNature Methods (2018), 15 (5), 351-354CODEN: NMAEA3; ISSN:1548-7091. (Nature Research)Hybrid methods that combine quantum mechanics (QM) and mol. mechanics (MM) can be applied to studies of reaction mechanisms in locations ranging from active sites of small enzymes to multiple sites in large bioenergetic complexes. By combining the widely used mol. dynamics and visualization programs NAMD and VMD with the quantum chem. packages ORCA and MOPAC, we created an integrated, comprehensive, customizable, and easy-to-use suite (http://www.ks.uiuc.edu/Research/qmmm). Through the QwikMD interface, setup, execution, visualization, and anal. are streamlined for all levels of expertise.
- 56Bursch, M.; Neugebauer, H.; Grimme, S. Structure Optimisation of Large Transition-Metal Complexes with Extended Tight-Binding Methods. Angew. Chem., Int. Ed. 2019, 58, 11078– 11087, DOI: 10.1002/anie.201904021Google Scholar56Structure Optimisation of Large Transition-Metal Complexes with Extended Tight-Binding MethodsBursch, Markus; Neugebauer, Hagen; Grimme, StefanAngewandte Chemie, International Edition (2019), 58 (32), 11078-11087CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)Large transition-metal complexes are used in numerous areas of chem. Computer-aided theor. investigations of such complexes are limited by the sheer size of real systems often consisting of hundreds to thousands of atoms. Accordingly, the development and thorough evaluation of fast semiempirical quantum chem. methods that are universally applicable to a large part of the periodic table is indispensable. Herein, we report on the capability of the recently developed GFNn-xTB method family for full quantum-mech. geometry optimization of medium to very large transition-metal complexes and organometallic supramol. structures. The results for a specially compiled benchmark set of 145 diverse closed-shell transition-metal complex structures for all metals up to Hg are presented. Further the GFNn-xTB methods are tested on three established benchmark sets regarding reaction energies and barrier heights of organometallic reactions.
- 57Maruno, Y.; Shoji, M.; Koizumi, K.; Nishiyama, Y.; Kitagawa, Y.; Kawakami, T.; Okumura, M.; Yamaguchi, K. Theoretical Studies on Magnetic Interactions between Ni (II) Ions in Urease. Polyhedron 2005, 24, 2778– 2783, DOI: 10.1016/j.poly.2005.03.161Google Scholar57Theoretical studies on magnetic interactions between Ni(II) ions in ureaseMaruno, Yusuke; Shoji, Mitsuo; Koizumi, Kenichi; Nishiyama, Yusuke; Kitagawa, Yasutaka; Kawakami, Takashi; Okumura, Mitsutaka; Yamaguchi, KizashiPolyhedron (2005), 24 (16-17), 2778-2783CODEN: PLYHDE; ISSN:0277-5387. (Elsevier B.V.)The chem. species of bridge oxygen (WB) in active site of urease was examd. in terms of a magnetic interaction between Ni(II) ions. The effective exchange integrals (Jab) values between Ni(II) ions were calcd. by using three candidates of WB, i.e., H2O, OH- and O2-. The unrestricted hybrid d. functional theory (UHDFT) calcns. were carried out using the smallest models and realistic models. In the case of the smallest models, the exptl. Jab value (-6.3 cm-1) was a middle value of WB = H2O (-2.6 cm-1) and WB = OH- (-11.4 cm-1), while we could not obtain an appropriate electronic structure using the O2--bridge model. The Jab became -16.7 cm-1 by using more realistic OH--bridge model, while one spin d. on the Ni2 ion was delocalized to Asp360 in case of H2O-bridge model. Natural orbital analyses revealed that the magnetic orbital through Lys-bridge delocalized to Asp360 and it seemed to assoc. with the spin transfer. From the results, the OH--bridge seemed to be the most likely candidate for the structure of active site in urease.
- 58Kitagawa, Y.; Saito, T.; Ito, M.; Nakanishi, Y.; Shoji, M.; Koizumi, K.; Yamanaka, S.; Kawakami, T.; Okumura, M.; Yamaguchi, K. Geometry Optimization Method Based on Approximate Spin Projection and its Application to F2, CH2, CH2OO, and Active Site of Urease. Int. J. Quantum Chem. 2007, 107, 3094– 3102, DOI: 10.1002/qua.21456Google Scholar58Geometry optimization method based on approximate spin projection and its application to F2, CH2, CH2OO, and active site of ureaseKitagawa, Yasutaka; Saito, Toru; Ito, Masahide; Nakanishi, Yasuyuki; Shoji, Mitsuo; Koizumi, Kenichi; Yamanaka, Shusuke; Kawakami, Takashi; Okumura, Mitsutaka; Yamaguchi, KizashiInternational Journal of Quantum Chemistry (2007), 107 (15), 3094-3102CODEN: IJQCB2; ISSN:0020-7608. (John Wiley & Sons, Inc.)A new geometry optimization method based on an approx. spin projection (AP) procedure is proposed to eliminate a spin contamination effect in an optimized structure on a low spin (LS) state of a broken symmetry (BS) method. First, an energy gradient with the AP (AP gradient) is derived and it is applied to the geometry optimization of F2, CH2, and CH2OO in order to obtain their structures without the spin contamination. The optimization method corrects H-C-H angle of the CH2 about 10° in comparison with the BS method, and indicates that the spin contamination in optimized geometry of the BS LS states is considerably large. Next, an active site of Urease that consists of two Ni(II) ions is optimized by the AP method. The result indicates that structure of the active site is sensitive to a water mol. called W3.
- 59Frisch, M. J.; Gaussian 16, Revision B.01; Gaussian Inc.: Wallingford, CT, 2016.Google ScholarThere is no corresponding record for this reference.
- 60Bussi, G.; Laio, A.; Parrinello, M. Equilibrium Free Energies from Nonequilibrium Metadynamics. Physi. Rev. Lett. 2006, 96, 090601, DOI: 10.1103/PhysRevLett.96.090601Google Scholar60Equilibrium Free Energies from Nonequilibrium MetadynamicsBussi, Giovanni; Laio, Alessandro; Parrinello, MichelePhysical Review Letters (2006), 96 (9), 090601/1-090601/4CODEN: PRLTAO; ISSN:0031-9007. (American Physical Society)We propose a new formalism to map history-dependent metadynamics in a Markovian process. We apply this formalism to model Langevin dynamics and det. the equil. distribution of a collection of simulations. We demonstrate that the reconstructed free energy is an unbiased est. of the underlying free energy and anal. derive an expression for the error. The present results can be applied to other history-dependent stochastic processes, such as Wang-Landau sampling.
- 61Castro, C. B.; Ferreira, M. P.; Caterina, G. M. N. Metalloenzyme Mechanisms Correlated to Their Turnover Number and Metal Lability. Curr. Res. Chem. Biol. 2021, 1, 100004, DOI: 10.1016/j.crchbi.2021.100004Google ScholarThere is no corresponding record for this reference.
- 62Krajewska, B.; van Eldik, R.; Brindell, M. Temperature- and Pressure-Dependent Stopped-Flow Kinetic Studies of Jack Bean Urease. Implications for the Catalytic Mechanism. J. Biol. Inorg. Chem. 2012, 17, 1123– 1134, DOI: 10.1007/s00775-012-0926-8Google Scholar62Temperature- and pressure-dependent stopped-flow kinetic studies of jack bean urease. Implications for the catalytic mechanismKrajewska, Barbara; van Eldik, Rudi; Brindell, MalgorzataJBIC, Journal of Biological Inorganic Chemistry (2012), 17 (7), 1123-1134CODEN: JJBCFA; ISSN:0949-8257. (Springer)Urease, a Ni-contg. metalloenzyme, features an activity that has profound medical and agricultural implications. The mechanism of this activity, however, has not been as yet thoroughly established. Accordingly, to improve its understanding, in this study we analyzed the steady-state kinetic parameters of the enzyme (jack bean), Km and kcat, measured at different temps. and pressures. Such an anal. is useful as it provides information on the mol. nature of the intermediate and transition states of the catalytic reaction. We measured the parameters in a noninteracting buffer using a stopped-flow technique at 15-35° and in the pressure range 5-132 MPa, the pressure-dependent measurements being the first of their kind performed for urease. While temp. enhanced the activity of urease, pressure inhibited the enzyme; the inhibition was biphasic. Analyzing Km provided the characteristics of the formation of the ES complex, and analyzing kcat, the characteristics of the activation of ES. From the temp.-dependent measurements, the energetic parameters were derived, i.e. thermodn. ΔHo and ΔSo for ES formation, and kinetic ΔH≠ and ΔS≠ for ES activation, while from the pressure-dependent measurements, the binding ΔVb and activation vols. were detd. The thermodn. and activation parameters obtained are discussed in terms of the current proposals for the mechanism of the urease reaction, and they are found to support the mechanism proposed by Benini et al., in which the Ni-Ni bridging hydroxide-not the terminal hydroxide-is the nucleophile in the catalytic reaction.
- 63El-Hefnawy, M. E.; Sakran, M.; Ismail, A. I.; Aboelfetoh, E. F. Extraction, Purification, Kinetic and Thermodynamic Properties of Urease from Germinating Pisum Sativum L. seeds. BMC Biochem. 2014, 15, 15, DOI: 10.1186/1471-2091-15-15Google Scholar63Extraction, purification, kinetic and thermodynamic properties of urease from germinating Pisum sativum L. seedsEl-Hefnawy, Mohamed E.; Sakran, Mohamed; Ismail, Ali I.; Aboelfetoh, Eman FahmyBMC Biochemistry (2014), 15 (), 15/1-15/8, 8 pp.CODEN: BBMIB3 ISSN:. (BioMed Central Ltd.)Background: Urease, one of the highly efficient known enzymes, catalyzes the hydrolysis of urea into ammonia and carbon dioxide. The present study aimed to ext. urease from pea seeds (Pisum sativum L). The enzyme was then purified in three consequence steps: acetone pptn., DEAE-cellulose ion-exchange chromatog. and gel filtration chromatog. (Sephacryl S-200 column). Results: The purifn. fold was 12.85 with a yield of 40%. The mol. wt. of the isolated urease was estd. by chromatog. to be 269,000 Daltons. Maximum urease activity (190 U/g) was achieved at the optimum conditions of 40°C and pH of 7.5 after 5 min of incubation. The kinetic parameters, Km and Vmax, were estd. by Lineweaver-Burk fits and found to be 500 mM and 333.3 U/g, resp. The thermodn. consts. of activation, ΔH, Ea, and ΔS, were detd. using Arrhenius plot and found to be 21.20 kJ/mol, 23.7 kJ/mol and 1.18 kJ/mol/K, resp. Conclusions: Urease was purified from germinating Pisum sativum L. seeds. The purifn. fold, yield and mol. wt. were detd. The effects of pH, concn. of enzyme, temp., concn. of substrate and storage period on urease activity were examd. This may provide an insight on the various aspects of the property of the enzyme. The significance of extg. urease from different sources could play a good role in understanding the metab. of urea in plants.
- 64Musiani, F.; Arnofi, E.; Casadio, R.; Ciurli, S. Structure-Based Computational Study of the Catalytic and Inhibition Mechanisms of Urease. J. Biol. Inorg. Chem. 2001, 6, 300– 314, DOI: 10.1007/s007750000204Google Scholar64Structure-based computational study of the catalytic and inhibition mechanisms of ureaseMusiani, Francesco; Arnofi, Elisa; Casadio, Rita; Ciurli, StefanoJBIC, Journal of Biological Inorganic Chemistry (2001), 6 (3), 300-314CODEN: JJBCFA; ISSN:0949-8257. (Springer-Verlag)The viability of different mechanisms of catalysis and inhibition of the nickel-contg. enzyme urease was explored using the available high-resoln. structures of the enzyme isolated from Bacillus pasteurii in the native form and inhibited with several substrates. The structures and charge distribution of urea, its catalytic transition state, and three enzyme inhibitors were calcd. using ab initio and d. functional theory methods. The DOCK program suite was employed to det. families of structures of urease complexes characterized by docking energy scores indicative of their relative stability according to steric and electrostatic criteria. Adjustment of the parameters used by DOCK, in order to account for the presence of the metal ion in the active site, resulted in the calcn. of best energy structures for the nickel-bound inhibitors β-mercaptoethanol, acetohydroxamic acid, and diamidophosphoric acid. These calcd. structures are in good agreement with the exptl. detd. structures, and provide hints on the reactivity and mobility of the inhibitors in the active site. The same docking protocol was applied to the substrate urea and its catalytic transition state, in order to shed light onto the possible catalytic steps occurring at the binuclear nickel active site. These calcns. suggest that the most viable pathway for urea hydrolysis involve a nucleophilic attack by the bridging, and not the terminal, nickel-bound hydroxide onto a urea mol., with active site residues playing important roles in orienting and activating the substrate, and stabilizing the catalytic transition state.
Cited By
Smart citations by scite.ai include citation statements extracted from the full text of the citing article. The number of the statements may be higher than the number of citations provided by ACS Publications if one paper cites another multiple times or lower if scite has not yet processed some of the citing articles.
This article is cited by 24 publications.
- Tatiana I. Mulashkina, Anna M. Kulakova, Maria G. Khrenova. Molecular Basis of the Substrate Specificity of Phosphotriesterase from Pseudomonas diminuta: A Combined QM/MM MD and Electron Density Study. Journal of Chemical Information and Modeling 2024, 64
(18)
, 7035-7045. https://doi.org/10.1021/acs.jcim.4c00425
- Yuzhuang Fu, Jun Yu, Fangfang Fan, Binju Wang, Zexing Cao. Elucidating the Enzymatic Mechanism of Dihydrocoumarin Degradation: Insight into the Functional Evolution of Methyl-Parathion Hydrolase from QM/MM and MM MD Simulations. The Journal of Physical Chemistry B 2024, 128
(23)
, 5567-5575. https://doi.org/10.1021/acs.jpcb.4c00970
- Qian-Qian Wang, Jinshuai Song, Donghui Wei. Origin of Chemoselectivity of Halohydrin Dehalogenase-Catalyzed Epoxide Ring-Opening Reactions. Journal of Chemical Information and Modeling 2024, 64
(11)
, 4530-4541. https://doi.org/10.1021/acs.jcim.4c00640
- Mark A. Hix, Alice R. Walker. AutoParams: An Automated Web-Based Tool To Generate Force Field Parameters for Molecular Dynamics Simulations. Journal of Chemical Information and Modeling 2023, 63
(20)
, 6293-6301. https://doi.org/10.1021/acs.jcim.3c01049
- Alexander Zlobin, Julia Belyaeva, Andrey Golovin. Challenges in Protein QM/MM Simulations with Intra-Backbone Link Atoms. Journal of Chemical Information and Modeling 2023, 63
(2)
, 546-560. https://doi.org/10.1021/acs.jcim.2c01071
- Kai Linghu, Kangjie Xu, Xinyi Zhao, Jingwen Zhou, Xinglong Wang. Modulating phosphate transfer process for promoting phosphorylation activity of acid phosphatase. Bioresource Technology 2025, 427 , 132348. https://doi.org/10.1016/j.biortech.2025.132348
- Yifei Zheng, Lingling Wu, Qiucai Zhang, Lin Hu, Yakun Tian, Min Wang, Huaimiao Zheng, Zhijun Zhang. A constant pH molecular dynamics and experimental study on the effect of different pH on the structure of urease from Sporosarcina pasteurii. Journal of Molecular Modeling 2025, 31
(6)
https://doi.org/10.1007/s00894-025-06369-w
- Mark A. Hix, Alice R. Walker. Adaptive Restraints to Accelerate Geometry Optimizations of Large Biomolecular Systems. Journal of Computational Chemistry 2025, 46
(12)
https://doi.org/10.1002/jcc.70127
- Helder M. Marques. The Bioinorganic Chemistry of the First Row d-Block Metal Ions—An Introduction. Inorganics 2025, 13
(5)
, 137. https://doi.org/10.3390/inorganics13050137
- Shengwei Sun, Per-Olof Syrén. Enzymatic degradation of phthalate esters in the environment: Advances, challenges and opportunities. Chemical Engineering Journal 2025, 508 , 160640. https://doi.org/10.1016/j.cej.2025.160640
- Sudheesh Devadas, Midhun George Thomas, Simahudeen Bathir Jaber Sathik Rifayee, Bhargav Varada, Walter White, Ethan Sommer, Kylin Campbell, Christopher J. Schofield, Christo Z. Christov. Origins of Catalysis in Non‐Heme Fe(II)/2‐Oxoglutarate‐Dependent Histone Lysine Demethylase KDM4A with Differently Methylated Histone H3 Peptides. Chemistry – A European Journal 2025, 31
(3)
https://doi.org/10.1002/chem.202403989
- Zhiwei Ma, Mengyao Chen, Juncheng Lu, Shichuang Liu, Yanling Ma, . Exploration of urease-aided calcium carbonate mineralization by enzyme analyses of
Neobacillus mesonae
strain NS-6. Microbiology Spectrum 2025, 13
(1)
https://doi.org/10.1128/spectrum.01891-24
- Shuqi Zhang, Shichuang Liu, Mengyao Chen, Juncheng Lu, Yanling Ma. Characterization of urease active calcite-producing strain YX-3 combined with the whole genome. Environmental Research 2024, 262 , 119855. https://doi.org/10.1016/j.envres.2024.119855
- Syamili Sarma, Anil Kumar Mishra. Microbial-Induced Calcium Carbonate Precipitation – A Potentially Sustainable Approach for Geo-environmental Challenges: A Retrospection into the Mechanism, Influencing Factors, Characterization, and Applications. Geomicrobiology Journal 2024, 41
(9)
, 921-938. https://doi.org/10.1080/01490451.2024.2401887
- Paulina Kaziukonytė, Vilius Petraška, Visvaldas Kairys, Algirdas Brukštus, Ieva Žutautė. Broadening the Scope of van Leusen Reaction: Investigation of Benzofuro[2,3‐
d
]imidazole Formation. Asian Journal of Organic Chemistry 2024, 13
(8)
https://doi.org/10.1002/ajoc.202400170
- John P. Pederson, Jesse G. McDaniel. PyDFT-QMMM: A modular, extensible software framework for DFT-based QM/MM molecular dynamics. The Journal of Chemical Physics 2024, 161
(3)
https://doi.org/10.1063/5.0219851
- Samia S. Alkhalil. Fluoroquinolone and enoxacin molecules are potential urease inhibitors for treating ureolytic bacterial infections. Materials Express 2024, 14
(4)
, 558-571. https://doi.org/10.1166/mex.2024.2650
- Javad Mohammadi, Mehdi Pordel, Mohammad Reza Bozorgmehr. Imidazo[4,5-a]acridines and Pyrazolo[4,3-a]acridines as a New Class of Urease Inhibitors: Synthesis, In Vitro Interactions, and Molecular Docking Studies. Pharmaceutical Chemistry Journal 2024, 57
(12)
, 1918-1928. https://doi.org/10.1007/s11094-024-03097-y
- Stéphane Réty. The Structure and Properties of Microbial Enzymes. Microorganisms 2024, 12
(1)
, 45. https://doi.org/10.3390/microorganisms12010045
- Jingya Fan, Guanyi Qu, Datao Wang, Jian Chen, Guocheng Du, Fang Fang. Synergistic Fermentation with Functional Microorganisms Improves Safety and Quality of Traditional Chinese Fermented Foods. Foods 2023, 12
(15)
, 2892. https://doi.org/10.3390/foods12152892
- Hinako Fujikawa, Hiroko X. Kondo, Yusuke Kanematsu, Yu Takano. Random Forest Classifier of Heme Proteins Using Porphyrin Distortions and Axial Ligands of Heme. Chemistry Letters 2023, 52
(6)
, 503-507. https://doi.org/10.1246/cl.230150
- Shunya Suenaga, Yu Takano, Toru Saito. Unraveling Binding Mechanism and Stability of Urease Inhibitors: A QM/MM MD Study. Molecules 2023, 28
(6)
, 2697. https://doi.org/10.3390/molecules28062697
- Hiroko X. Kondo, Hiroyuki Iizuka, Gen Masumoto, Yuichi Kabaya, Yusuke Kanematsu, Yu Takano. Prediction of Protein Function from Tertiary Structure of the Active Site in Heme Proteins by Convolutional Neural Network. Biomolecules 2023, 13
(1)
, 137. https://doi.org/10.3390/biom13010137
- Zahra Najafi rad, Farzaneh Farzad, Leila Razavi. Surface functionalization of graphene nanosheet with poly (l-histidine) and its application in drug delivery: covalent vs non-covalent approaches. Scientific Reports 2022, 12
(1)
https://doi.org/10.1038/s41598-022-21619-0
Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.
Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.
The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.
Recommended Articles
Abstract
Scheme 1
Scheme 1. Urea Hydrolysis ReactionFigure 1
Figure 1. Active site of Sporosarcina pasteurii urease (SPU) and two possible substrate binding modes, monodentate 1 and bidentate 2.
Scheme 2
Scheme 2. Previously Proposed Urease Reaction Mechanisms (i)–(iv) and Our Proposed One (v), the Details of Which Are Described in the Text, with Amino Acid Numbering for SPU/KAUFigure 2
Figure 2. QM regions of 1 and 2 after the QM(GFN2-xTB)/MM MD equilibrations (top). Key bond distances are also presented in Å (bottom).
Figure 3
Figure 3. Potential energy profiles (in kcal mol–1) and illustrations of the transition state and product structures corresponding to (a) mechanism (v) starting from 1′ and (b) mechanism (ii) starting from 2′ obtained at GFN2-xTB and UB3LYP/def2-SVP levels.
Figure 4
Figure 4. (a) One-dimensional potentials of mean force (1D-PMF) of the nucleophilic attack reaction and representative snapshots of the active site in (b) 1R, (c) 1TS1A, and (d) 1IntA, with key distances in Å.
Figure 5
Figure 5. (a) One-dimensional potentials of mean force (1D-PMF) of the proton transfer reaction and representative snapshots of the active site in (b) 1TS2A and (c) 1PA, with key distances in Å.
Figure 6
Figure 6. (a) One-dimensional potentials of mean force (1D-PMF) of the W2 dissociation and representative snapshots of the active site in (b) 1TSB and (c) 1PB, with key distances in Å.
Figure 7
Figure 7. (a) One-dimensional potentials of mean force (1D-PMF) of the nucleophilic attack reaction and representative snapshots of the active site in (b) 2R, (c) 2TS1A, and (d) 2IntA, with key distances in Å.
Figure 8
Figure 8. (a) One-dimensional potentials of mean force (1D-PMF) of the proton transfer reaction and representative snapshots of the active site in (b) 2TS2A and (c) 2PA, with key distances in Å.
References
This article references 64 other publications.
- 1Dixon, N. E.; Gazzola, C.; Blakeley, R. L.; Zerner, B. Jack Bean Urease (EC 3.5.1.5). Metalloenzyme. Simple Biological Role for Nickel. J. Am. Chem. Soc. 1975, 97, 4131– 4133, DOI: 10.1021/ja00847a0451Jack bean urease (EC 3.5.1.5). Metalloenzyme. Simple biological role for nickelDixon, Nicholas E.; Gazzola, Carlo; Blakeley, Robert L.; Zerner, BurtJournal of the American Chemical Society (1975), 97 (14), 4131-3CODEN: JACSAT; ISSN:0002-7863.Jack bean (Canavalia enisformis) urease in phosphate buffer, pH 7.0 (1 mM each in EDTA and β-mercaptoethanol) exhibited electronic and (or) vibrational transitions with broad absorption maxima at 1060 and 725 nm and shoulders at ∼425, ∼376, and ∼316 nm. The spectrum was reversibly modified by β-mercaptoethanol and by acetohydroxamic acid. Chem. and at. absorption anal. showed the presence of 2.0 g atoms of Ni/105,000 g of enzyme. The residual sp. activity was a linear function of the Ni content after denaturation at pH 3.7 in the presence of EDTA. Urease appears to be the 1st example of a Ni metalloenzyme. Ni may well be an essential trace element in jack beans. Published spectral characteristics of other enzymes are consistent with the presence of tightly-bound transition metal ions.
- 2Blakeley, R. L.; Zerner, B. Jack Bean Urease: the First Nickel Enzyme. J. Mol. Catal. 1984, 23, 263– 292, DOI: 10.1016/0304-5102(84)80014-02Jack bean urease: the first nickel enzymeBlakeley, Robert L.; Zerner, BurtJournal of Molecular Catalysis (1984), 23 (2-3), 263-92CODEN: JMCADS; ISSN:0304-5102.A review and discussion with 98 refs.
- 3Hausinger, R. P. Nickel Utilization by Microorganisms. Microbiol. Rev. 1987, 51, 22– 42, DOI: 10.1128/mr.51.1.22-42.19873Nickel utilization by microorganismsHausinger, Robert P.Microbiological Reviews (1987), 51 (1), 22-42CODEN: MBRED3; ISSN:0146-0749.A review with 232 refs. on biol. functions of Ni, transport of Ni, and Ni-contg. enzymes.
- 4Mobley, H. L.; Hausinger, R. P. Microbial Ureases: Significance, Regulation, and Molecular Characterization. Microbiol. Rev. 1989, 53, 85– 108, DOI: 10.1128/mr.53.1.85-108.19894Microbial ureases: significance, regulation, and molecular characterizationMobley, Harry L. T.; Hausinger, Robert P.Microbiological Reviews (1989), 53 (1), 85-108CODEN: MBRED3; ISSN:0146-0749.A review, with 249 refs., of urease in pathogenesis, in the rumen and gastrointestinal tract, in soil and aq. environments, and in microbe physiol. Methods for microbial urease detn. and purifn., characterization of microbial ureases, and the mol. biol. of ureolysis (e.g. plasmid-encoded ureases, cloning of urease genes, bacterial urease operon, fungal and plant urease mol. biol.) are discussed.
- 5Blakeley, R. L.; Webb, E. C.; Zerner, B. Jack Bean Urease (EC 3.5.1.5). A New Purification and Reliable Rate Assay. Biochemistry 1969, 8, 1984– 1990, DOI: 10.1021/bi00833a0315Jack bean urease (EC 3.5.1.5). A new purification and reliable rate assayBlakeley, Robert L.; Webb, Edwin C.; Zerner, BurtBiochemistry (1969), 8 (5), 1984-90CODEN: BICHAW; ISSN:0006-2960.Jack bean urease (EC 3.5.1.5) was reproducibly purified to a hitherto unobtained and const. specific activity. The procedure is capable of handling "poor" jack bean meal. A CHCl3-Me2CO powder of jack bean meal is extd. with 30% Me2CO contg. 1% mercaptoethanol at 39° for 5 min. The filtrate from this step is allowed to stand for 48 hrs. at 4°, and the crystals of the enzyme are harvested by centrifugation. After recycling gel filtration in the presence of 1 mM mercaptoethanol and 1 mM EDTA (Sephadex G-200), and concn. by (NH4)2SO4 pptn., the enzyme reached max. specific activity. A reliable pH-stat assay in the presence of 2 μM dithiothreitol was devised.
- 6Carter, E. L.; Flugga, N.; Boer, J. L.; Mulrooney, S. B.; Hausinger, R. P. Interplay of Metal Ions and Urease. Metallomics 2009, 1, 207– 221, DOI: 10.1039/b903311d6Interplay of metal ions and ureaseCarter, Eric L.; Flugga, Nicholas; Boer, Jodi L.; Mulrooney, Scott B.; Hausinger, Robert P.Metallomics (2009), 1 (3), 207-221CODEN: METAJS; ISSN:1756-591X. (Royal Society of Chemistry)A review. Urease, the first enzyme to be crystd., contains a dinuclear nickel metallocenter that catalyzes the decompn. of urea to produce ammonia, a reaction of great agricultural and medical importance. Several mechanisms of urease catalysis have been proposed on the basis of enzyme crystal structures, model complexes, and computational efforts, but the precise steps in catalysis and the requirement of nickel vs. other metals remain unclear. Purified bacterial urease is partially activated via incubation with carbon dioxide plus nickel ions; however, in vitro activation also has been achieved with manganese and cobalt. In vivo activation of most ureases requires accessory proteins that function as nickel metallochaperones and GTP-dependent mol. chaperones or play other roles in the maturation process. In addn., some microorganisms control their levels of urease by metal ion-dependent regulatory mechanisms.
- 7Zambelli, B.; Musiani, F.; Benini, S.; Ciurli, S. Chemistry of Ni2+ in Urease: Sensing, Trafficking, and Catalysis. Acc. Chem. Res. 2011, 44, 520– 530, DOI: 10.1021/ar200041k7Chemistry of Ni2+ in urease: Sensing, trafficking, and catalysisZambelli, Barbara; Musiani, Francesco; Benini, Stefano; Ciurli, StefanoAccounts of Chemical Research (2011), 44 (7), 520-530CODEN: ACHRE4; ISSN:0001-4842. (American Chemical Society)A review. Transition metals are both essential to enzymic catalysis and limited in environmental availability. These 2 biol. facts have together driven organisms to evolve mechanisms for selective metal ion sensing and utilization. Changes in metal ion concns. are perceived by metal-dependent transcription factors and transduced into appropriate cellular responses, which regulate the machineries of competitive metal ion homeostasis and metalloenzyme activation. The intrinsic toxicity of the majority of metal ions further creates a need for regulated intracellular trafficking, which is carried out by specific chaperones. The Ni2+-dependent urease system serves as a paradigm for studying the strategies that cells use to handle an essential, yet toxic, metal ion. Although the discovery of urease as the 1st biol. system for which Ni2+ is essential for activity dates to 1975, the rationale for Ni2+ selection, as well as the cascade of events involving metal-dependent gene regulation and protein-protein interactions leading to enzyme activation, have yet to be fully unraveled. The past 14 years since the paper by P. A. Karplus et al. (1997) have witnessed impressive achievements in the understanding of the biochem. of Ni2+ in the urease system. Here, the authors discuss more recent advances in the comprehension of the specific role of Ni2+ in catalysis and in the interplay between Ni2+ and other metal ions, such as Zn2+ and Fe2+, in metal-dependent enzyme activity. The discussion focuses on work carried out in the authors' lab. In particular, the structural features of the enzyme bound to inhibitors, substrate analogs, and transition state or intermediate analogs have shed light on the catalytic mechanism. Structural and functional information has been correlated to understand the Ni2+ sensing effected by NikR protein, a Ni2+-dependent transcription factor. The urease activation process, involving the insertion of Ni2+ into the urease active site, has been in part dissected and analyzed through the investigation of the mol. properties of accessory proteins UreD, UreF, and UreG. The intracellular trafficking of Ni2+ has been rationalized through a deeper understanding of the structural and metal-binding properties of metallochaperone UreE. All the while, a no. of key general concepts have been revealed and developed. These include an understanding of the following: (1) the overall ancillary role of Zn2+ in Ni metab.; (2) the intrinsically disordered nature of the GTPase responsible for coupling the energy consumption to the CO2 requirement for the urease activation process; and (3) the role of the accessory proteins regulating this GTPase activity.
- 8Maroney, M. J.; Ciurli, S. Nonredox Nickel Enzymes. Chem. Rev. 2014, 114, 4206– 4228, DOI: 10.1021/cr40044888Nonredox nickel enzymesMaroney, Michael J.; Ciurli, StefanoChemical Reviews (Washington, DC, United States) (2014), 114 (8), 4206-4228CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)A review. Current knowledge of the biochem., structure, and catalytic mechanism of enzymes, including urease, glyoxalase I, and acireductone dioxygenase, whose active sites require Ni and utilize it in a non-redox role, are discussed.
- 9Mazzei, L.; Musiani, F.; Ciurli, S. The Structure-based Reaction Mechanism of Urease, a Nickel Dependent Enzyme: Tale of a Long Debate. J. Biol. Inorg. Chem. 2020, 25, 829, DOI: 10.1007/s00775-020-01808-w9The structure-based reaction mechanism of urease, a nickel dependent enzyme: tale of a long debateMazzei, Luca; Musiani, Francesco; Ciurli, StefanoJBIC, Journal of Biological Inorganic Chemistry (2020), 25 (6), 829-845CODEN: JJBCFA; ISSN:0949-8257. (Springer)Abstr.: This review is an attempt to retrace the chronicle that starts from the discovery of the role of nickel as the essential metal ion in urease for the enzymic catalysis of urea, a key step in the biogeochem. cycle of nitrogen on Earth, to the most recent progress in understanding the chem. of this historical enzyme. Data and facts are presented through the magnifying lenses of the authors, using their best judgment to filter and elaborate on the many facets of the research carried out on this metalloenzyme over the years. The tale is divided in chapters that discuss and describe the results obtained in the subsequent leaps in the knowledge that led from the discovery of a biol. role for Ni to the most recent advancements in the comprehension of the relationship between the structure and function of urease. This review is intended not only to focus on the bioinorg. chem. of this beautiful metal-based catalysis, but also, and maybe primarily, to evoke inspiration and motivation to further explore the realm of bio-based coordination chem.
- 10Konieczna, I.; Zarnowiec, P.; Kwinkowski, M.; Kolesinska, B.; Fraczyk, J.; Kaminski, Z.; Kaca, W. Bacterial urease and its role in long-lasting human diseases. Curr. Protein Pept. Sci. 2012, 13, 789– 806, DOI: 10.2174/13892031280487109410Bacterial urease and its role in long-lasting human diseasesKonieczna, Iwona; Zarnowiec, Paulina; Kwinkowski, Marek; Kolesinska, Beata; Fraczyk, Justyna; Kaminski, Zbigniew; Kaca, WieslawCurrent Protein and Peptide Science (2012), 13 (8), 789-806CODEN: CPPSCM; ISSN:1389-2037. (Bentham Science Publishers Ltd.)A review. Urease is a virulence factor found in various pathogenic bacteria. It is essential in colonization of a host organism and in maintenance of bacterial cells in tissues. Due to its enzymic activity, urease has a toxic effect on human cells. The presence of ureolytic activity is an important marker of a no. of bacterial infections. Urease is also an immunogenic protein and is recognized by antibodies present in human sera. The presence of such antibodies is connected with progress of several long-lasting diseases, like rheumatoid arthritis, atherosclerosis or urinary tract infections. In bacterial ureases, motives with a sequence and/or structure similar to human proteins may occur. This phenomenon, known as mol. mimicry, leads to the appearance of autoantibodies, which take part in host mols. destruction. Detection of antibodies-binding motives (epitopes) in bacterial proteins is a complex process. However, org. chem. tools, such as synthetic peptide libraries, are helpful in both, epitope mapping as well as in serol. investigations. In this review, we present a synthetic report on a mol. organization of bacterial ureases - genetic as well as structural. We characterize methods used in detecting urease and ureolytic activity, including techniques applied in disease diagnostic processes and in chem. synthesis of urease epitopes. The review also provides a summary of knowledge about a toxic effect of bacterial ureases on human body and about occurrence of anti-urease antibodies in long-lasting diseases.
- 11Rajakovich, L. J.; Balskus, E. P. Metabolic Functions of the Human Gut Microbiota: the Role of Metalloenzymes. Nat. Prod. Rep. 2019, 36, 593– 625, DOI: 10.1039/C8NP00074C11Metabolic functions of the human gut microbiota: the role of metalloenzymesRajakovich, Lauren J.; Balskus, Emily P.Natural Product Reports (2019), 36 (4), 593-625CODEN: NPRRDF; ISSN:0265-0568. (Royal Society of Chemistry)Covering: up to the end of 2017The human body is composed of an equal no. of human and microbial cells. While the microbial community inhabiting the human gastrointestinal tract plays an essential role in host health, these organisms have also been connected to various diseases. Yet, the gut microbial functions that modulate host biol. are not well established. In this review, we describe metabolic functions of the human gut microbiota that involve metalloenzymes. These activities enable gut microbial colonization, mediate interactions with the host, and impact human health and disease. We highlight cases in which enzyme characterization has advanced our understanding of the gut microbiota and examples that illustrate the diverse ways in which metalloenzymes facilitate both essential and unique functions of this community. Finally, we analyze Human Microbiome Project sequencing datasets to assess the distribution of a prominent family of metalloenzymes in human-assocd. microbial communities, guiding future enzyme characterization efforts.
- 12Bremner, J. M.; Krogmeier, M. J. Evidence that the Adverse Effect of Urea Fertilizer on Seed Germination in Soil is due to Ammonia Formed through Hydrolysis of Urea by Soil Urease. Proc. Natl. Acad. Sci. U. S. A. 1989, 86, 8185– 8188, DOI: 10.1073/pnas.86.21.818512Evidence that the adverse effect of urea fertilizer on seed germination in soil is due to ammonia formed through hydrolysis of urea by soil ureaseBremner, John M.; Krogmeier, Michael J.Proceedings of the National Academy of Sciences of the United States of America (1989), 86 (21), 8185-8CODEN: PNASA6; ISSN:0027-8424.Studies using seeds of wheat (Triticum aestivum), rye (Secale cereale), barley (Hordeum vulgare) and corn (Zea mays) indicated that the adverse effect of urea fertilizer on seed germination in soil is due to ammonia formed through hydrolysis of urea by soil urease and is not due to urea itself, to urea fertilizer impurities such as biuret, or to nitrite formed by nitrification of urea nitrogen. Support for this conclusion was obtained from (i) comparison of the effects on seed germination in soil of purified urea, urea fertilizers, urea fertilizer impurities, and compds. formed by enzymic and microbial transformations of urea in soil; (ii) studies showing that ammonia volatilized from soils treated with urea completely inhibited germination of seeds close to, but not in contact with, these soils; and (iii) expts. showing that the adverse effect of urea fertilizer on seed germination in soil was completely eliminated when the soil was autoclaved to destroy urease or was treated with phenylphosphorodiamidate to inhibit soil urease activity before treatment with urea fertilizer.
- 13Gioacchini, P.; Nastri, A.; Marzadori, C.; Giovannini, C.; Antisari, L. V.; Gessa, C. Influence of Urease and Nitrification Inhibitors on N Losses from Soils Fertilized with Urea. Biol. Fertil. 2002, 36, 129– 135, DOI: 10.1007/s00374-002-0521-113Influence of urease and nitrification inhibitors on N losses from soils fertilized with ureaGioacchini, Paola; Nastri, Anna; Marzadori, Claudio; Giovannini, Camilla; Vittori Antisari, Livia; Gessa, CarloBiology and Fertility of Soils (2002), 36 (2), 129-135CODEN: BFSOEE; ISSN:0178-2762. (Springer-Verlag)The aim was to evaluate how the N losses through volatilization and leaching from soils fertilized with urea can be affected by the application of a urease inhibitor or a urease plus a nitrification inhibitor. The expt. was carried out using lysimeters with 15N-labeled urea and N-butylthiophosphoric triamide (NBPT) as urease inhibitor and dicyandiamide (DCD) as nitrification inhibitor, comparing 3 treatments: urea alone (U), urea + NBPT (UN) and urea + NBPT + DCD (UND). Both volatilization and leaching were significantly different in the soils used, according to their physicochem. characteristics. However, the pattern of the loss was similar: the volatilization was significantly reduced by NBPT (UN), but the presence of DCD (UND) significantly increased the loss, with respect to UN. Considering leaching, the highest amt. of NO3- was lost with UND, the lowest with U. The greatest amt. of N lost by leaching was soil-derived N produced by the N mineralization-immobilization turnover. We suggest that, by maintaining the NH4+ in the soils, the inhibitors, in particular DCD, caused a priming effect with a subsequent increase in the rate of soil org. matter mineralization and an extra release of soil org. N. The priming effect was real in the sandy loam (SL) soil where a net N release was obsd., whereas in the clay loam (CL) soil the effect of the inhibitors was less pronounced and an apparent priming effect was obsd.; however, a real priming effect also cannot be excluded in this soil.
- 14Jabri, E.; Carr, M. B.; Hausinger, R. P.; Karplus, P. A. The Crystal Structure of Urease from Klebsiella aerogenes. Science 1995, 268, 998– 1004, DOI: 10.1126/science.775439514The crystal structure of urease from Klebsiella aerogenesJabri, Evelyn; Carr, Mary Beth; Hausinger, Robert P.; Karplus, P. AndrewScience (Washington, D. C.) (1995), 268 (5213), 998-1004CODEN: SCIEAS; ISSN:0036-8075. (American Association for the Advancement of Science)The crystal structure of urease from K. aerogenes was detd. at 2.2 Å resoln. and refined to an R factor of 18.2%. The enzyme contained 4 structural domains: 3 with novel folds playing structural roles, and an (αβ)8 barrel domain, which contained the bi-Ni center. The 2 active site Ni ions were 3.5 Å apart. One Ni ion was coordinated by 3 ligands (with low occupancy of a 4th ligand) and the 2nd was coordinated by 5 ligands. A carbamylated Lys residue provided an O ligand to each Ni, explaining why CO2 is required for the activation of urease apoenzyme. The structure was compatible with a catalytic mechanism whereby urea ligates Ni-1 to complete its tetrahedral coordination and a hydroxide ligand of Ni-2 attacks the carbonyl C atom. A surprisingly high structural similarity between the urease catalytic domain and that of the Zn-dependent adenosine deaminase revealed a remarkable example of active site divergence.
- 15Benini, S.; Rypniewski, W. R.; Wilson, K. S.; Miletti, S.; Ciurli, S.; Mangani, S. A New Proposal for Urease Mechanism Based on the Crystal Structures of the Native and Inhibited Enzyme from Bacillus pasteurii: Why Urea Hydrolysis Costs Two Nickels. Structure 1999, 7, 205– 216, DOI: 10.1016/S0969-2126(99)80026-415A new proposal for urease mechanism based on the crystal structures of the native and inhibited enzyme from Bacillus pasteurii: why urea hydrolysis costs two nickelsBenini, Stefano; Rypniewski, Wojciech R.; Wilson, Keith S.; Miletti, Silvia; Ciurli, Stefano; Mangani, StefanoStructure (London) (1999), 7 (2), 205-216CODEN: STRUE6; ISSN:0969-2126. (Current Biology Publications)Urease catalyzes the hydrolysis of urea, the final step of org. nitrogen mineralization, using a bimetallic nickel center. The role of the active site metal ions and amino acid residues has not been elucidated to date. Many pathologies are assocd. with the activity of ureolytic bacteria, and the efficiency of soil nitrogen fertilization with urea is severely decreased by urease activity. Therefore, the development of urease inhibitors would lead to a redn. of environmental pollution, to enhanced efficiency of nitrogen uptake by plants, and to improved therapeutic strategies for treatment of infections due to ureolytic bacteria. Structure-based design of urease inhibitors would require knowledge of the enzyme mechanism at the mol. level. The structures of native and inhibited urease from Bacillus pasteurii have been detd. at a resoln. of 2.0 Å by synchroton X-ray cryogenic crystallog. In the native enzyme, the coordination sphere of each of the two nickel ions is completed by a water mol. and a bridging hydroxide. A fourth water mol. completes a tetrahedral cluster of solvent mols. The enzyme crystd. in the presence of phenylphosphorodiamidate contains the tetrahedral transition-state analog diamidophosphoric acid, bound to the two nickel ions in an unprecedented mode. Comparison of the native and inhibited structures reveals two distinct conformations of the flap lining the active-site cavity. The mode of binding of the inhibitor, and a comparison between the native and inhibited urease structures, indicate a novel mechanism for enzymic urea hydrolysis which reconciles the available structural and biochem. data.
- 16Ha, N.-C.; Oh, S.-T.; Sung, J. Y.; Cha, K. A.; Lee, M. H.; Oh, B.-H. Supramolecular Assembly and Acid Resistance of Helicobacter pylori Urease. Nat. Struct. Biol. 2001, 8, 505– 509, DOI: 10.1038/8856316Supramolecular assembly and acid resistance of Helicobacter pylori ureaseHa, Nam-Chul; Oh, Sang-Taek; Sung, Jae Young; Cha, Kyeung Ah; Lee, Mann Hyung; Oh, Byung-HaNature Structural Biology (2001), 8 (6), 505-509CODEN: NSBIEW; ISSN:1072-8368. (Nature America Inc.)H. pylori, an etiol. agent in a variety of gastro-duodenal diseases, produces a large amt. of urease (I), which is believed to neutralize gastric acid by producing NH3 for the survival of the bacteria. Up to 30% of I assocs. with the surface of intact cells upon lysis of neighboring bacteria. The role of I at the extracellular location has been a subject of controversy because the purified enzyme is irreversibly inactivated below pH 5. Here, the authors detd. the crystal structure of H. pylori I, which had a 1.1 MDa spherical assembly of 12 catalytic units with an outer diam. of ∼160 Å. Under physiol. relevant conditions, the activity of the enzyme remains unaffected down to pH 3. Activity assays under different conditions indicated that the duster of the 12 active sites on the supramol. assembly may be crit. for the survival of the enzyme at low pH. The structure provides a novel example of a mol. assembly adapted for acid resistance that, together with the low Km of the I, is likely to enable the organism to inhabit the hostile niche.
- 17Phillips, K.; Munster, D. J.; Allardyce, R. A.; Bagshaw, P. F. Antibacterial Action of the Urease Inhibitor Acetohydroxamic acid on Helicobacter pylori. J. Clin. Pathol. 1993, 46, 372– 373, DOI: 10.1136/jcp.46.4.37217Antibacterial action of the urease inhibitor acetohydroxamic acid on Helicobacter pyloriPhillips K; Munster D J; Allardyce R A; Bagshaw P FJournal of clinical pathology (1993), 46 (4), 372-3 ISSN:0021-9746.The urease inhibitor acetohydroxamic acid (AHA) was assessed for its bacteriostatic and bactericidal effects on Helicobacter pylori. For eight isolates of H pylori, the minimum inhibitory concentration (MIC) was either 200 mg/l or 400 mg/l. Interactions between AHA and antimicrobial drugs used to treat H pylori were also determined. For most isolates AHA reduced the MIC for colloidal bismuth subcitrate (CBS), tetracycline, metronidazole, and amoxicillin. In a few isolates, however, AHA increased the minimum bactericidal concentration (MBC) for these antimicrobial treatments. In vitro AHA is active against H pylori and it interacts with other agents directed against H pylori.
- 18Benini, S.; Rypniewski, W. R.; Wilson, K. S.; Miletti, S.; Ciurli, S.; Mangani, S. The Complex of Bacillus pasteurii Urease with Acetohydroxamate Anion from X-ray Data at 1.55 Å Resolution. J. Biol. Inorg. Chem. 2000, 5, 110– 118, DOI: 10.1007/s00775005001418The complex of Bacillus pasteurii urease with acetohydroxamate anion from x-ray data at 1.55 Å resolutionBenini, Stefano; Rypniewski, Wojciech R.; Wilson, Keith S.; Miletti, Silvia; Ciurli, Stefano; Mangani, StefanoJBIC, Journal of Biological Inorganic Chemistry (2000), 5 (1), 110-118CODEN: JJBCFA; ISSN:0949-8257. (Springer-Verlag)The crystal structure of B. pasteurii urease inhibited by acetohydroxamic acid was solved and refined anisotropically using synchrotron x-ray cryogenic diffraction data (1.55 Å resoln.; 99.5% completeness; data redundancy = 26; R-factor = 15.1%, PDB code 4UBP). The 2 Ni2+ ions in the active site were sepd. by a distance of 3.53 Å. The structure clearly showed the binding mode of the inhibitor anion, sym. bridging the 2 Ni2+ ions in the active site through the hydroxamate O atom and chelating 1 Ni2+ ion through the carbonyl O atom. The flexible flap flanking the active site cavity was in the open conformation. The possible implications of the results on structure-based mol. design of new urease inhibitors were discussed.
- 19Benini, S.; Rypniewski, W.; Wilson, K.; Ciurli, S.; Mangani, S. Structure-based Rationalization of Urease Inhibition by Phosphate: Novel Insights into the Enzyme Mechanism. J. Biol. Inorg. Chem. 2001, 6, 778– 790, DOI: 10.1007/s00775010025419Structure-based rationalization of urease inhibition by phosphate: novel insights into the enzyme mechanismBenini, S.; Rypniewski, W. R.; Wilson, K. S.; Ciurli, S.; Mangani, S.JBIC, Journal of Biological Inorganic Chemistry (2001), 6 (8), 778-790CODEN: JJBCFA; ISSN:0949-8257. (Springer-Verlag)The structure of Bacillus pasteurii urease (BPU) inhibited with phosphate was solved and refined using synchrotron x-ray diffraction data from a vitrified crystal (1.85 Å resoln., 99.3% completeness, data redundancy 4.6, R-factor 17.3%, PDB code 6UBP). A distance of 3.5 Å separates the two Ni ions in the active site. The binding mode of the inhibitor involves the formation of four coordination bonds with the two Ni ions: one phosphate oxygen atom sym. bridges the two metal ions (1.9-2.0 Å), while two of the remaining phosphate oxygen atoms bind to the Ni atoms at 2.4 Å. The fourth phosphate oxygen is directed into the active site channel. Anal. of the H-bonding network around the bound inhibitor indicates that phosphate is bound as the H2PO4- anion, and that an addnl. proton is present on the Oδ2 atom of Aspα363, an active site residue involved in Ni coordination through Oδ1. The flexible flap flanking the active site cavity is in the open conformation. Anal. of the complex reveals why phosphate is a relatively weak inhibitor and why sulfate does not bind to the nickels in the active site. The implications of the results for the understanding of the urease catalytic mechanism are reviewed. A novel alternative for the proton donor is presented.
- 20Benini, S.; Rypniewski, W. R.; Wilson, K. S.; Mangani, S.; Ciurli, S. Molecular Details of Urease Inhibition by Boric Acid: Insights into the Catalytic Mechanism. J. Am. Chem. Soc. 2004, 126, 3714– 3715, DOI: 10.1021/ja049618p20Molecular Details of Urease Inhibition by Boric Acid: Insights into the Catalytic MechanismBenini, Stefano; Rypniewski, Wojciech R.; Wilson, Keith S.; Mangani, Stefano; Ciurli, StefanoJournal of the American Chemical Society (2004), 126 (12), 3714-3715CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)The structure of the complex of urease, a Ni-contg. metalloenzyme, with boric acid was detd. at 2.10 Å resoln. The complex shows the unprecedented binding mode of the competitive inhibitor to the dinuclear metal center, with the B(OH)3 mol. bridging the Ni ions and leaving in place the bridging hydroxide. Boric acid can be considered a substrate analog of urea, and the structure supports the proposal that the Ni-bridging hydroxide acts as the nucleophile in the enzymic process of urea hydrolysis.
- 21Benini, S.; Cianci, M.; Mazzei, L.; Ciurli, S. Fluoride inhibition of Sporosarcina pasteurii urease: structure and thermodynamics. J. Biol. Inorg. Chem. 2014, 19, 1243– 1261, DOI: 10.1007/s00775-014-1182-x21Fluoride inhibition of Sporosarcina pasteurii urease: structure and thermodynamicsBenini, Stefano; Cianci, Michele; Mazzei, Luca; Ciurli, StefanoJBIC, Journal of Biological Inorganic Chemistry (2014), 19 (8), 1243-1261CODEN: JJBCFA; ISSN:0949-8257. (Springer)Urease is a Ni-dependent enzyme and a virulence factor for ureolytic bacterial human pathogens, but it is also necessary to convert urea, the most worldwide-used fertilizer, into forms of nitrogen that can be taken up by crop plants. A strategy to control the activity of urease for medical and agricultural applications is to use enzyme inhibitors. F- is a known urease inhibitor, but the structural basis of its mode of inhibition is still undetd. Here, kinetic studies on the F--induced inhibition of urease from S. pasteurii, a widespread and highly ureolytic soil bacterium, were performed using isothermal titrn. calorimetry and revealed a mixed competitive and uncompetitive mechanism. The pH dependence of the Ki values, investigated in the pH 6.5-8.0 range, revealed a predominant uncompetitive mechanism that increased by increasing the pH, and a lesser competitive inhibition that increased by lowering the pH. Ten crystal structures of the enzyme were independently detd. using 5 crystals of the native form and 5 crystals of the protein crystd. in the presence of F-. The anal. of these structures revealed the presence of 2 F- anions coordinated to the Ni2+ ions in the active site, in terminal and bridging positions. The present study consistently supported an interaction of F- with the Ni centers in the urease active site in which one F- anion competitively bound to Ni2+ proposed to coordinate urea in the initial step of the catalytic mechanism, while another F- anion uncompetitively substituted the Ni2+-bridging hydroxide, blocking its nucleophilic attack on urea.
- 22Mazzei, L.; Cianci, M.; Musiani, F.; Lente, G.; Palombo, M.; Ciurli, S. Inactivation of Urease by Catechol: Kinetics and Structure. J. inorg. Biochem. 2017, 166, 182– 189, DOI: 10.1016/j.jinorgbio.2016.11.01622Inactivation of urease by catechol: Kinetics and structureMazzei, Luca; Cianci, Michele; Musiani, Francesco; Lente, Gabor; Palombo, Marta; Ciurli, StefanoJournal of Inorganic Biochemistry (2017), 166 (), 182-189CODEN: JIBIDJ; ISSN:0162-0134. (Elsevier)Urease is a Ni(II)-contg. enzyme that catalyzes the hydrolysis of urea to yield NH3 and carbamate at a rate 1015-fold higher than the uncatalyzed reaction. Urease is a virulence factor of several human pathogens, in addn. to decreasing the efficiency of soil org. nitrogen fertilization. Therefore, efficient urease inhibitors are actively sought. Here, the authors describe the characterization of the interaction between urease from Sporosarcina pasteurii (SPU) and Canavalia ensiformis (jack bean, JBU) with catechol, a model polyphenol. In particular, catechol irreversibly inactivated both SPU and JBU with a complex radical-based autocatalytic multistep mechanism. The crystal structure of the SPU-catechol complex, detd. at 1.50 Å resoln., revealed the structural details of the enzyme inhibition.
- 23Mazzei, L.; Cianci, M.; Contaldo, U.; Ciurli, S. Insights into Urease Inhibition by N-(n-Butyl) Phosphoric Triamide through an Integrated Structural and Kinetic Approach. J. Agric. Food Chem. 2019, 67, 2127– 2138, DOI: 10.1021/acs.jafc.8b0479123Insights into urease inhibition by N-(n-butyl) phosphoric triamide through an integrated structural and kinetic approachMazzei, Luca; Cianci, Michele; Contaldo, Umberto; Ciurli, StefanoJournal of Agricultural and Food Chemistry (2019), 67 (8), 2127-2138CODEN: JAFCAU; ISSN:0021-8561. (American Chemical Society)The nickel-dependent enzyme urease represents a neg. element for the efficiency of soil nitrogen fertilization as well as a virulence factor for a large no. of pathogenic and antibiotic-resistant bacteria. The development of ever more efficient urease inhibitors demands knowledge of their modes of action at the mol. level. N-(n-Butyl)-phosphoric triamide (NBPTO) is the oxo-deriv. of N-(n-butyl)-thiophosphoric triamide (NBPT), which is extensively employed in agriculture to increase the efficiency of urea-based fertilizers. The 1.45 Å resoln. structure of the enzyme-inhibitor complex obtained upon incubation of Sporosarcina pasteurii urease (SPU) with NBPTO shows the presence of diamido phosphoric acid (DAP), generated upon enzymic hydrolysis of NBPTO with the release of Bu amine. DAP is bound in a tridentate binding mode to the two Ni(II) ions in the active site of urease via two O atoms and an amide NH2 group, whereas the second amide group of DAP points away from the metal center into the active-site channel. The mobile flap modulating the size of the active-site cavity is found in a disordered closed-open conformation. A kinetic characterization of the NBPTO-based inhibition of both bacterial (SPU) and plant (Canavalia ensiformis or jack bean, JBU) ureases, carried out by calorimetric measurements, indicates the occurrence of a reversible slow-inhibition mode of action. The latter is characterized by a very small value of the equil. dissocn. const. of the urease-DAP complex caused, in turn, by the large rate const. for the formation of the enzyme-inhibitor complex. The much greater capability of NBPTO to inhibit urease, as compared with that of NBPT, is thus not caused by the presence of a P=O moiety vs. a P=S moiety, as previously suggested, but rather by the readiness of NBPTO to react with urease without the need to convert one of the P-NH2 amide moieties to its P-OH acid deriv., as in the case of NBPT. The latter process is indeed characterized by a very small equil. const. that reduces drastically the concn. of the active form of the inhibitor in the case of NBPT. This indicates that high-efficiency phosphoramide-based urease inhibitors must have at least one O atom bound to the central P atom in order for the mol. to efficiently and rapidly bind to the dinickel center of the enzyme.
- 24Mazzei, L.; Cianci, M.; Benini, S.; Ciurli, S. The Structure of the Elusive Urease-Urea Complex Unveils the Mechanism of a Paradigmatic Nickel-Dependent Enzyme. Angew. Chem., Int. Ed. 2019, 58, 7415– 7419, DOI: 10.1002/anie.20190356524The structure of the elusive urease-urea complex unveils the mechanism of a paradigmatic nickel-dependent enzymeMazzei, Luca; Cianci, Michele; Benini, Stefano; Ciurli, StefanoAngewandte Chemie, International Edition (2019), 58 (22), 7415-7419CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)Urease, the most efficient enzyme known, contains an essential dinuclear NiII cluster in the active site. It catalyzes the hydrolysis of urea, inducing a rapid pH increase that has neg. effects on human health and agriculture. Thus, the control of urease activity is of utmost importance in medical, pharmaceutical, and agro-environmental applications. All known urease inhibitors are either toxic or inefficient. The development of new and efficient chems. able to inhibit urease relies on the knowledge of all steps of the catalytic mechanism. The short (microseconds) lifetime of the urease-urea complex has hampered the detn. of its structure. The present study uses fluoride to substitute the hydroxide acting as the co-substrate in the reaction, preventing the occurrence of the catalytic steps that follow substrate binding. The 1.42 Å crystal structure of the urease-urea complex, reported here, resolves the enduring debate on the mechanism of this metalloenzyme.
- 25Karplus, P. A.; Pearson, M. A.; Hausinger, R. P. 70 Years of Crystalline Urease: What Have We Learned?. Acc. Chem. Res. 1997, 30, 330– 337, DOI: 10.1021/ar960022j2570 Years of crystalline urease: What have we learned?Karplus, P. Andrew; Pearson, Matthew; Hausinger, Robert P.Accounts of Chemical Research (1997), 30 (8), 330-337CODEN: ACHRE4; ISSN:0001-4842. (American Chemical Society)A review with 60 refs. on the functional and structural characteristics of Klebsiella aerogenes urease. Crystallog. and functional studies of wild-type urease and some site-directed variants suggest a detailed catalytic mechanism which accounts for its kinetic properties and its structure. This mechanism assigns the limiting pKa's of 6.4 and 9.0 to the general acid His-320 and the hydrolytic water, resp. This assignment is such that only a small fraction of the urease mols. are present in the correct protonation state for activity. Obsd. disorder of the solvation in the active site and the comparison of multiple structures provide evidence for suboptimal interactions in the water-filled active site. These suboptimal interactions may lead to an enthalpy-driven free energy gain when this water is released into the bulk solvent upon urea binding. As such, the relaxation of the high-energy state could provide a large amt. of binding energy required for catalysis.
- 26Pearson, M. A.; Park, I.-S.; Schaller, R. A.; Michel, L. O.; Karplus, P. A.; Hausinger, R. P. Kinetic and Structural Characterization of Urease Active Site Variants. Biochemistry 2000, 39, 8575– 8584, DOI: 10.1021/bi000613o26Kinetic and Structural Characterization of Urease Active Site VariantsPearson, Matthew A.; Park, Il-Seon; Schaller, Ruth A.; Michel, Linda O.; Karplus, P. Andrew; Hausinger, Robert P.Biochemistry (2000), 39 (29), 8575-8584CODEN: BICHAW; ISSN:0006-2960. (American Chemical Society)Klebsiella aerogenes urease uses a dinuclear nickel active site to catalyze urea hydrolysis at >1014-fold the spontaneous rate. To better define the enzyme mechanism, we examd. the kinetics and structures for a suite of site-directed variants involving four residues at the active site: His320, His219, Asp221, and Arg336. Compared to wild-type urease, the H320A, H320N, and H320Q variants exhibit similar ∼10-5-fold deficiencies in rates, modest Km changes, and disorders in the peptide flap covering their active sites. H219A urease exhibits 103-fold increased Km over that of native enzyme, whereas the increase is less marked (∼102-fold) in the H219N and H219Q variants that retain hydrogen bonding capability. Structures for these variants show clearly resolved active site water mols. covered by well-ordered peptide flaps. Whereas the D221N variant is only moderately affected compared to wild-type enzyme, D221A urease possesses low activity (∼10-3 that of native enzyme), a small increase in Km, and a pH 5 optimum. The crystal structure for D221A urease is reminiscent of the His320 variants. The R336Q enzyme has a ∼10-4-fold decreased catalytic rate with near-normal pH dependence and an unaffected Km. Phenylglyoxal inactivates the R336Q variant at over half the rate obsd. for native enzyme, demonstrating that modification of non-active-site arginines can eliminate activity, perhaps by affecting the peptide flap. Our data favor a mechanism in which His219 helps to polarize the substrate carbonyl group, a metal-bound terminal hydroxide or bridging oxo-dianion attacks urea to form a tetrahedral intermediate, and protonation occurs via the general acid His320 with Asp221 and Arg336 orienting and influencing the acidity of this residue. Furthermore, we conclude that the simple bell-shaped pH dependence of kcat and kcat/Km for the native enzyme masks a more complex underlying pH dependence involving at least four pKas.
- 27Dixon, N. E.; Blakeley, R. L.; Zerner, B. Jack Bean Urease (EC 3.5.1.5). III. The Involvement of Active-site Nickel Ion in Inhibition by b-Mercaptoethanol, Phosphoramidate, and Fluoride. Can. J. Biochem. 1980, 58, 481– 488, DOI: 10.1139/o80-06427Jack bean urease (EC 3.5.1.5). III. The involvement of active-site nickel ion in inhibition by β-mercaptoethanol, phosphoramidate, and fluorideDixon, Nicholas E.; Blakeley, Robert L.; Zerner, BurtCanadian Journal of Biochemistry (1980), 58 (6), 481-8CODEN: CJBIAE; ISSN:0008-4018.Interaction of β-mercaptoethanol with urease produces large, rapid, and fully reversible spectral changes in that part of the electronic absorption spectrum which is assocd. with the tightly bound Ni2+. The spectrophotometrically detd. value of the dissocn. const. of the β-mercaptoethanol-urease complex (0.95 mM at pH 7.12 and 25°) is in agreement with the Ki (0.72 mM) for β-mercaptoethanol acting as a competitive inhibitor in the hydrolysis of urea. This constitutes direct evidence that the Ni2+ in jack bean urease is at the active site. Inhibition of urease by phosphoramidate is slowly achieved and slowly reversed, and upon reactivation of the isolated phosphoramidate-urease complex, phosphoramidate is regenerated in good yield. Spectrophotometric expts. indicate that phosphoramidate binds to Ni2+ in urease. Competition with β-mercaptoethanol was used to det. a dissocn. const. (1.23 mM at pH 7.12 and 25°) for a F--urease complex in which F- also coordinates with an active-site Ni2+. Kinetic evidence is presented which indicates that in the presence of urea, a ternary complex (F-urea-urease) is formed.
- 28Pelmenschikov, V.; Siegbahn, P. E. Nickel Superoxide Dismutase Reaction Mechanism Studied by Hybrid Density Functional Methods. J. Am. Chem. Soc. 2006, 128, 7466– 7475, DOI: 10.1021/ja053665f28Nickel Superoxide Dismutase Reaction Mechanism Studied by Hybrid Density Functional MethodsPelmenschikov, Vladimir; Siegbahn, Per E. M.Journal of the American Chemical Society (2006), 128 (23), 7466-7475CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)The reaction mechanism for the disproportionation of the toxic superoxide radical to mol. oxygen and hydrogen peroxide by the nickel-dependent superoxide dismutase (NiSOD) has been studied using the B3LYP hybrid DFT method. Based on the recent x-ray structures of the enzyme in the resting oxidized Ni(III) and x-ray-reduced Ni(II) states, the model investigated includes the backbone spacer of six residues (sequence nos. 1-6) as a structural framework. The side chains of residues His-1, Cys-2, and Cys-6, which are essential for nickel binding and catalysis, were modeled explicitly. The catalytic cycle consists of two half-reactions, each initiated by the successive substrate approach to the metal center. The two protons necessary for the dismutation are postulated to be delivered concertedly with the superoxide radical anions. The first (reductive) phase involves Ni(III) redn. to Ni(II), and the second (oxidative) phase involves the metal reoxidn. back to its resting state. The Cys-2 thiolate sulfur serves as a transient protonation site in the interim between the two half-reactions, allowing for the dioxygen and hydrogen peroxide mols. to be released in the reductive and oxidative phases, resp. The His-1 side chain nitrogen and backbone amides of the active site channel are shown to be less favorable transient proton locations, as compared to the Cys-2 sulfur. Comparisons are made to the Cu- and Zn-dependent SOD, studied previously using similar models.
- 29Sparta, M.; Valdez, C. E.; Alexandrova, A. N. Metal-dependent Activity of Fe and Ni Acireductone Dioxygenases: How Two Electrons Reroute the Catalytic Pathway. J. Mol. Biol. 2013, 425, 3007– 3018, DOI: 10.1016/j.jmb.2013.05.00129Metal-Dependent Activity of Fe and Ni Acireductone Dioxygenases: How Two Electrons Reroute the Catalytic PathwaySparta, Manuel; Valdez, Crystal E.; Alexandrova, Anastassia N.Journal of Molecular Biology (2013), 425 (16), 3007-3018CODEN: JMOBAK; ISSN:0022-2836. (Elsevier Ltd.)Two virtually identical acireductone dioxygenases, ARD and ARD', catalyze completely different oxidn. reactions of the same substrate, 1,2-dihydroxy-3-keto-5-(methylthio)pentene, depending exclusively on the nature of the bound metal. Fe2+-dependent ARD' produces the α-keto acid precursor of methionine and formate and allows for the recycling of methionine in cells. Ni2+-dependent ARD instead produces methylthiopropionate, CO, and formate, and exits the methionine salvage cycle. This mechanistic difference has not been understood to date but has been speculated to be due to the difference in coordination of the substrate to Fe2+ vs. Ni2+: forming a five-membered ring vs. a six-membered ring, resp., thus exposing different carbon atoms for the attack by O2. Here, using mixed quantum-classical mol. dynamics simulations followed by the d. functional theory mechanistic investigation, we show that contrary to the old hypothesis, both metals preferentially bind the substrate as a six-membered ring, exposing the exact same sites to the attack by O2. It is the electronic properties of the metals that are then responsible for the system following different reaction paths, to yield the resp. products. We fully explain the puzzling metal-induced difference in functionality between ARD and ARD' and, in particular, propose a new mechanism for ARD'. All results are in agreement with available isotopic substitution and other exptl. data.
- 30Wang, W.-J.; Wei, W.-J.; Liao, R.-Z. Deciphering the Chemoselectivity of Nickel-dependent Quercetin 2, 4-Dioxygenase. Phys. Chem. Chem. Phys. 2018, 20, 15784– 15794, DOI: 10.1039/C8CP02683A30Deciphering the chemoselectivity of nickel-dependent quercetin 2,4-dioxygenaseWang, Wen-Juan; Wei, Wen-Jie; Liao, Rong-ZhenPhysical Chemistry Chemical Physics (2018), 20 (23), 15784-15794CODEN: PPCPFQ; ISSN:1463-9076. (Royal Society of Chemistry)The reaction mechanism and chemoselectivity of nickel-dependent quercetin 2,4-dioxygenase (2,4-QueD) have been investigated using the QM/MM approach. The protonation state of the Glu74 residue, a first-shell ligand of Ni, has been considered to be either neutral or deprotonated. QM/MM calcns. predict that Glu74 must be deprotonated to rationalize the chemoselectivity and steer the 2,4-dioxygenolytic cleavage of quercetin, which harvests the exptl.-obsd. product, 2-protocatechuoylphloroglucinol carboxylic acid, coupled with the release of carbon monoxide. If the enzyme has a neutral Glu74 residue, the undesired 2,3-dioxygenolytic cleavage of quercetin becomes the dominant pathway, leading to the formation of α-keto acid. The calcns. suggest that the reaction takes place via three major steps: (1) attack of the superoxide on the C2 of the substrate pyrone ring to generate a NiII-peroxide intermediate; (2) formation of the second C-O bond between C4 and the peroxide to produce a peroxide bridge; (3) simultaneous cleavage of the C2-C3, C3-C4, and O1-O2 bonds with the formation of 2-protocatechuoylphloroglucinol carboxylic acid and carbon monoxide. The third step was found to be rate-limiting, with a barrier of 17.4 kcal mol-1, which is in very good agreement with the exptl. kinetic data. For the second C-O bond formation, an alternative pathway is that the peroxide attacks the C3 of the substrate pyrone ring, leading to the formation of a four-membered ring intermediate, which then undergoes concerted C2-C3 and O1-O2 bond cleavages to produce an α-keto acid. This pathway is assocd. with a barrier of 30.6 kcal mol-1, which is much higher than the major pathway. When Glu74 is protonated, the 2,3-dioxygenolytic pathway, however, has a lower barrier (21.8 kcal mol-1) than the 2,4-dioxygenolytic pathway.
- 31Siegbahn, P. E.; Chen, S.-L.; Liao, R.-Z. Theoretical Studies of Nickel-dependent Enzymes. Inorganics 2019, 7, 95, DOI: 10.3390/inorganics708009531Theoretical studies of nickel-dependent enzymesSiegbahn, Per E. M.; Chen, Shi-Lu; Liao, Rong-ZhenInorganics (2019), 7 (8), 95CODEN: INORCW; ISSN:2304-6740. (MDPI AG)The advancements of quantum chem. methods and computer power allow detailed mechanistic investigations of metalloenzymes. In particular, both quantum chem. cluster and combined QM/MM approaches have been used, which have been proven to successfully complement exptl. studies. This review starts with a brief introduction of nickel-dependent enzymes and then summarizes theor. studies on the reaction mechanisms of these enzymes, including NiFe hydrogenase, methyl-coenzyme M reductase, nickel CO dehydrogenase, acetyl CoA synthase, acireductone dioxygenase, quercetin 2,4-dioxygenase, urease, lactate racemase, and superoxide dismutase.
- 32Suárez, D.; Díaz, N.; Merz, K. M. Ureases: Quantum Chemical Calculations on Cluster Models. J. Am. Chem. Soc. 2003, 125, 15324– 15337, DOI: 10.1021/ja030145g32Ureases: Quantum Chemical Calculations on Cluster ModelsSuarez, Dimas; Diaz, Natalia; Merz, Kenneth M., Jr.Journal of the American Chemical Society (2003), 125 (50), 15324-15337CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)Herein, the authors present results from a computational study of dinickel complexes that are relevant to the catalytic hydrolysis of urea exerted by the urease enzymes. The B3LYP d. functional is used to characterize the equil. geometry, electronic and magnetic properties, and energies for a series of realistic complexes modeling the active site of ureases. The anal. of the theor. results gives new insight into the structure, substrate binding, and catalytic mechanism. The water bridge between the two Ni(II) ions obsd. in the crystallog. structures of the ureases was assigned to a hydroxide bridge in agreement with the obsd. small antiferromagnetic coupling. Both monodentate and bidentate urea-bound complexes, in which urea had favorable orientations for catalysis, were characterized. Finally, two reaction mechanisms were investigated starting from the monodentate and bidentate urea-bound complexes, resp. Both a Ni1···Ni2 bridging hydroxide and a Ni2-bound water mol. play crucial roles in the two mechanisms.
- 33Estiu, G.; Merz, K. M. Competitive Hydrolytic and Elimination Mechanisms in the Urease Catalyzed Decomposition of Urea. J. Phys. Chem. B 2007, 111, 10263– 10274, DOI: 10.1021/jp072323o33Competitive Hydrolytic and Elimination Mechanisms in the Urease Catalyzed Decomposition of UreaEstiu, Guillermina; Merz, Kenneth M., Jr.Journal of Physical Chemistry B (2007), 111 (34), 10263-10274CODEN: JPCBFK; ISSN:1520-6106. (American Chemical Society)We present a high-level quantum chem. study of possible elimination reaction mechanisms assocd. with the catalytic decompn. of urea at the binuclear nickel active site cluster of urease. Stable intermediates and transition state structures have been identified along several possible reaction pathways. The computed results are compared with those reported by Suarez et al. for the hydrolytic catalyzed decompn. On the basis of these comparative studies, we propose a monodentate coordination of urea in the active site from which both the elimination and hydrolytic pathways can decomp. urea into CO2 and NH3. This observation is counter to what has been exptl. suggested based on the exogenous observation of carbamic acid (the reaction product from the hydrolysis pathway). However, this does not address what has occurred at the active site of urease prior to product release. On the basis of our computed results, the observation that urea prefers the elimination channel in aq. soln. and on the observation of Lippard and co-workers of an elimination reaction channel in a urease biomimetic model, we propose that the elimination channel needs to be re-examd. as a viable reaction channel in urease.
- 34Carlsson, H.; Nordlander, E. Computational Modeling of the Mechanism of Urease. Bioinorg. Chem. Appl. 2010, 2010, 364891, DOI: 10.1155/2010/364891There is no corresponding record for this reference.
- 35Barrios, A. M.; Lippard, S. J. Interaction of Urea with a Hydroxide-Bridged Dinuclear Nickel Center: An Alternative Model for the Mechanism of Urease. J. Am. Chem. Soc. 2000, 122, 9172– 9177, DOI: 10.1021/ja000202v35Interaction of Urea with a Hydroxide-Bridged Dinuclear Nickel Center: An Alternative Model for the Mechanism of UreaseBarrios, Amy M.; Lippard, Stephen J.Journal of the American Chemical Society (2000), 122 (38), 9172-9177CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)A hydroxide-bridged dinuclear nickel complex with a urea mol. linking the two metal ions through its carbonyl oxygen atom has been prepd. as a model for the metalloenzyme urease. This complex, [Ni2(μ-OH)(μ-urea)(bdptz)(urea)(CH3CN)](ClO4)3, where bdptz is the dinucleating ligand 1,4-bis(2,2'-dipyridylmethyl)phthalazine, effects the hydrolysis of urea upon heating in a two-step reaction. In the first step, a mol. of ammonia is eliminated from urea with concomitant prodn. of cyanate, the first-order rate const. in acetonitrile being (7.7 ± 0.5) × 10-4 h-1. This reaction is at least 500 times faster than the spontaneous decompn. of urea under the same conditions. When the cyanate-contg. product is further heated in the presence of water, the cyanate is hydrolyzed with a second-order rate const. of (9.5 ± 1) × 10-4 M-1 h-1. Reaction of [Ni2(μ-OH)(μ-urea)(bdptz)(urea)(CH3CN)](ClO4)3 in 50% aq. acetonitrile afforded ammonia with no appreciable buildup of the cyanate-contg. species. A possible analog of the cyanate-contg. product, [Ni2(μ-OH)(μ-H2O)(bdptz)(μ-OCN)]2(OTs)4, was independently synthesized and structurally characterized. These results establish the precedence for hydrolysis of urea via a cyanate intermediate as an alternative mechanism for the urease-catalyzed hydrolysis of urea.
- 36Senn, H. M.; Thiel, W. QM/MM Methods for Biomolecular Systems. Angew. Chem., Int. Ed. 2009, 48, 1198– 1229, DOI: 10.1002/anie.20080201936QM/MM methods for biomolecular systemsSenn, Hans Martin; Thiel, WalterAngewandte Chemie, International Edition (2009), 48 (7), 1198-1229CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)A review. Combined quantum-mechanics/mol.-mechanics (QM/MM) approaches have become the method of choice for modeling reactions in biomol. systems. Quantum-mech. (QM) methods are required for describing chem. reactions and other electronic processes, such as charge transfer or electronic excitation. However, QM methods are restricted to systems of up to a few hundred atoms. However, the size and conformational complexity of biopolymers calls for methods capable of treating up to several 100,000 atoms and allowing for simulations over time scales of tens of nanoseconds. This is achieved by highly efficient, force-field-based mol. mechanics (MM) methods. Thus to model large biomols. the logical approach is to combine the two techniques and, to use a QM method for the chem. active region (e.g., substrates and co-factors in an enzymic reaction) and an MM treatment for the surroundings (e.g., protein and solvent). The resulting schemes are commonly referred to as combined or hybrid QM/MM methods. They enable the modeling of reactive biomol. systems at a reasonable computational effort while providing the necessary accuracy.
- 37Ahmadi, S.; Barrios Herrera, L.; Chehelamirani, M.; Hostaš, J.; Jalife, S.; Salahub, D. R. Multiscale Modeling of Enzymes: QM-cluster, QM/MM, and QM/MM/MD: A Tutorial Review. Int. J. Quantum Chem. 2018, 118, e25558, DOI: 10.1002/qua.25558There is no corresponding record for this reference.
- 38Neese, F. The ORCA Program System. WIREs Comput. Mol. Sci. 2012, 2, 73– 78, DOI: 10.1002/wcms.8138The ORCA program systemNeese, FrankWiley Interdisciplinary Reviews: Computational Molecular Science (2012), 2 (1), 73-78CODEN: WIRCAH; ISSN:1759-0884. (Wiley-Blackwell)A review. ORCA is a general-purpose quantum chem. program package that features virtually all modern electronic structure methods (d. functional theory, many-body perturbation and coupled cluster theories, and multireference and semiempirical methods). It is designed with the aim of generality, extendibility, efficiency, and user friendliness. Its main field of application is larger mols., transition metal complexes, and their spectroscopic properties. ORCA uses std. Gaussian basis functions and is fully parallelized. The article provides an overview of its current possibilities and documents its efficiency.
- 39Becke, A. D. Density-functional Exchange-energy Approximation with Correct Asymptotic Behavior. Phys. Rev. A 1988, 38, 3098– 3100, DOI: 10.1103/PhysRevA.38.309839Density-functional exchange-energy approximation with correct asymptotic behaviorBecke, A. D.Physical Review A: Atomic, Molecular, and Optical Physics (1988), 38 (6), 3098-100CODEN: PLRAAN; ISSN:0556-2791.Current gradient-cor. d.-functional approxns. for the exchange energies of at. and mol. systems fail to reproduce the correct 1/r asymptotic behavior of the exchange-energy d. A gradient-cor. exchange-energy functional is given with the proper asymptotic limit. This functional, contg. only one parameter, fits the exact Hartree-Fock exchange energies of a wide variety of at. systems with remarkable accuracy, surpassing the performance of previous functionals contg. two parameters or more.
- 40Lee, C.; Yang, W.; Parr, R. G. Development of the Colle-Salvetti Correlation-energy Formula into a Functional of the Electron Density. Phys. Rev. B 1988, 37, 785– 789, DOI: 10.1103/PhysRevB.37.78540Development of the Colle-Salvetti correlation-energy formula into a functional of the electron densityLee, Chengteh; Yang, Weitao; Parr, Robert G.Physical Review B: Condensed Matter and Materials Physics (1988), 37 (2), 785-9CODEN: PRBMDO; ISSN:0163-1829.A correlation-energy formula due to R. Colle and D. Salvetti (1975), in which the correlation energy d. is expressed in terms of the electron d. and a Laplacian of the 2nd-order Hartree-Fock d. matrix, is restated as a formula involving the d. and local kinetic-energy d. On insertion of gradient expansions for the local kinetic-energy d., d.-functional formulas for the correlation energy and correlation potential are then obtained. Through numerical calcns. on a no. of atoms, pos. ions, and mols., of both open- and closed-shell type, it is demonstrated that these formulas, like the original Colle-Salvetti formulas, give correlation energies within a few percent.
- 41Becke, A. D. Density-functional Thermochemistry. III. The Role of Exact Exchange. J. Chem. phys. 1993, 98, 5648– 5652, DOI: 10.1063/1.46491341Density-functional thermochemistry. III. The role of exact exchangeBecke, Axel D.Journal of Chemical Physics (1993), 98 (7), 5648-52CODEN: JCPSA6; ISSN:0021-9606.Despite the remarkable thermochem. accuracy of Kohn-Sham d.-functional theories with gradient corrections for exchange-correlation, the author believes that further improvements are unlikely unless exact-exchange information is considered. Arguments to support this view are presented, and a semiempirical exchange-correlation functional (contg. local-spin-d., gradient, and exact-exchange terms) is tested for 56 atomization energies, 42 ionization potentials, 8 proton affinities, and 10 total at. energies of first- and second-row systems. This functional performs better than previous functionals with gradient corrections only, and fits expt. atomization energies with an impressively small av. abs. deviation of 2.4 kcal/mol.
- 42Stephens, P. J.; Devlin, F. J.; Chabalowski, C. F.; Frisch, M. J. Ab Initio Calculation of Vibrational Absorption and Circular Dichroism Spectra Using Density Functional Force Fields. J. Phys. Chem. 1994, 98, 11623– 11627, DOI: 10.1021/j100096a00142Ab Initio Calculation of Vibrational Absorption and Circular Dichroism Spectra Using Density Functional Force FieldsStephens, P. J.; Devlin, F. J.; Chabalowski, C. F.; Frisch, M. J.Journal of Physical Chemistry (1994), 98 (45), 11623-7CODEN: JPCHAX; ISSN:0022-3654.The unpolarized absorption and CD spectra of the fundamental vibrational transitions of the chiral mol. 4-methyl-2-oxetanone are calcd. ab initio. Harmonic force fields are obtained using d. functional theory (DFT), MP2 and SCF methodologies, and a [5s4p2d/3s2p] (TZ2P) basis set. DFT calcns. use the LSDA, BLYP, and Becke3LYP (B3LYP) d. functionals. Mid-IR spectra predicted using LSDA, BLYP, and B3LYP force fields are of significantly different quality, the B3LYP force field yielding spectra in clearly superior, and overall excellent, agreement with expt. The MP2 force field yields spectra in slightly worse agreement with expt. than the B3LYP force field. The SCF force field yields spectra in poor agreement with expt. The basis set dependence of B3LYP force fields is also explored: the 6-31G* and TZ2P basis sets give very similar results while the 3-21G basis set yields spectra in substantially worse agreement with expt.
- 43Neese, F.; Wennmohs, F.; Hansen, A.; Becker, U. Efficient, Approximate and Parallel Hartree-Fock and Hybrid DFT Calculations. A ‘Chain-of-Spheres’ Algorithm for the Hartree-Fock Exchange. Chem. Phys. 2009, 356, 98– 109, DOI: 10.1016/j.chemphys.2008.10.03643Efficient, approximate and parallel Hartree-Fock and hybrid DFT calculations. A 'chain-of-spheres' algorithm for the Hartree-Fock exchangeNeese, Frank; Wennmohs, Frank; Hansen, Andreas; Becker, UteChemical Physics (2009), 356 (1-3), 98-109CODEN: CMPHC2; ISSN:0301-0104. (Elsevier B.V.)In this paper, the possibility is explored to speed up Hartree-Fock and hybrid d. functional calcns. by forming the Coulomb and exchange parts of the Fock matrix by different approxns. For the Coulomb part the previously introduced Split-RI-J variant of the well-known d. fitting' approxn. is used. The exchange part is formed by semi-numerical integration techniques that are closely related to Friesner's pioneering pseudo-spectral approach. Our potentially linear scaling realization of this algorithm is called the 'chain-of-spheres exchange' (COSX). A combination of semi-numerical integration and d. fitting is also proposed. Both Split-RI-J and COSX scale very well with the highest angular momentum in the basis sets. It is shown that for extended basis sets speed-ups of up to two orders of magnitude compared to traditional implementations can be obtained in this way. Total energies are reproduced with an av. error of <0.3 kcal/mol as detd. from extended test calcns. with various basis sets on a set of 26 mols. with 20-200 atoms and up to 2000 basis functions. Reaction energies agree to within 0.2 kcal/mol (Hartree-Fock) or 0.05 kcal/mol (hybrid DFT) with the canonical values. The COSX algorithm parallelizes with a speedup of 8.6 obsd. for 10 processes. Min. energy geometries differ by less than 0.3 pm in the bond distances and 0.5° in the bond angles from their canonical values. These developments enable highly efficient and accurate SCF calcns. including nonlocal Hartree-Fock exchange for large mols. In combination with the RI-MP2 method and large basis sets, second-order many body perturbation energies can be obtained for medium sized mols. with unprecedented efficiency. The algorithms are implemented into the ORCA electronic structure system.
Moving Frontiers in Quantum Chemistry.
- 44Weigend, F.; Ahlrichs, R. Balanced Basis Sets of Split Valence, Triple Zeta Valence and Quadruple Zeta Valence Quality for H to Rn: Design and Assessment of Accuracy. Phys. Chem. Chem. Phys. 2005, 7, 3297– 3305, DOI: 10.1039/b508541a44Balanced basis sets of split valence, triple zeta valence and quadruple zeta valence quality for H to Rn: Design and assessment of accuracyWeigend, Florian; Ahlrichs, ReinhartPhysical Chemistry Chemical Physics (2005), 7 (18), 3297-3305CODEN: PPCPFQ; ISSN:1463-9076. (Royal Society of Chemistry)Gaussian basis sets of quadruple zeta valence quality for Rb-Rn are presented, as well as bases of split valence and triple zeta valence quality for H-Rn. The latter were obtained by (partly) modifying bases developed previously. A large set of more than 300 mols. representing (nearly) all elements-except lanthanides-in their common oxidn. states was used to assess the quality of the bases all across the periodic table. Quantities investigated were atomization energies, dipole moments and structure parameters for Hartree-Fock, d. functional theory and correlated methods, for which we had chosen Moller-Plesset perturbation theory as an example. Finally recommendations are given which type of basis set is used best for a certain level of theory and a desired quality of results.
- 45Breneman, C. M.; Wiberg, K. B. Determining Atom-centered Monopoles from Molecular Electrostatic Potentials. The Need for High Sampling Density in Formamide Conformational Analysis. J. Comput. Chem. 1990, 11, 361– 373, DOI: 10.1002/jcc.54011031145Determining atom-centered monopoles from molecular electrostatic potentials. The need for high sampling density in formamide conformational analysisBreneman, Curt M.; Wiberg, Kenneth B.Journal of Computational Chemistry (1990), 11 (3), 361-73CODEN: JCCHDD; ISSN:0192-8651.An improved method for computing potential-derived charges is described which is based upon the CHELP program available from QCPE. This approach (CHELPG) is shown to be considerably less dependent upon mol. orientation than the original CHELP program. In the second part of this work, the CHELPG point selection algorithm was used to analyze the changes in the potential-derived charges in formamide during rotation about the C-N bond. In order to achieve a level of rotational invariance less than 10% of the magnitude of the electronic effects studied, an equally-spaced array of points 0.3 Å apart was required. Points found to be greater than 2.8 Å from any nucleus were eliminated, along with all points contained within the defined VDW distances from each of the atoms. The results are compared to those obtained by using CHELP. Even when large nos. of points (ca. 3000) were sampled using the CHELP selection routine, the results did not indicate a satisfactory level of rotational invariance. On the basis of these results, the original CHELP program was inadequate for analyzing internal rotations.
- 46Vanommeslaeghe, K.; Hatcher, E.; Acharya, C.; Kundu, S.; Zhong, S.; Shim, J.; Darian, E.; Guvench, O.; Lopes, P.; Vorobyov, I.; Mackerell, A. D., Jr. CHARMM General Force Field: A Force Field for Drug-like Molecules Compatible with the CHARMM All-atom Additive Biological Force Fields. J. Comput. Chem. 2010, 31, 671– 690, DOI: 10.1002/jcc.2136746CHARMM general force field: A force field for drug-like molecules compatible with the CHARMM all-atom additive biological force fieldsVanommeslaeghe, K.; Hatcher, E.; Acharya, C.; Kundu, S.; Zhong, S.; Shim, J.; Darian, E.; Guvench, O.; Lopes, P.; Vorobyov, I.; Mackerell, A. D., Jr.Journal of Computational Chemistry (2010), 31 (4), 671-690CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)The widely used CHARMM additive all-atom force field includes parameters for proteins, nucleic acids, lipids, and carbohydrates. In the present article, an extension of the CHARMM force field to drug-like mols. is presented. The resulting CHARMM General Force Field (CGenFF) covers a wide range of chem. groups present in biomols. and drug-like mols., including a large no. of heterocyclic scaffolds. The parametrization philosophy behind the force field focuses on quality at the expense of transferability, with the implementation concg. on an extensible force field. Statistics related to the quality of the parametrization with a focus on exptl. validation are presented. Addnl., the parametrization procedure, described fully in the present article in the context of the model systems, pyrrolidine, and 3-phenoxymethyl-pyrrolidine will allow users to readily extend the force field to chem. groups that are not explicitly covered in the force field as well as add functional groups to and link together mols. already available in the force field. CGenFF thus makes it possible to perform "all-CHARMM" simulations on drug-target interactions thereby extending the utility of CHARMM force fields to medicinally relevant systems. © 2009 Wiley Periodicals, Inc.J Comput Chem, 2010.
- 47Li, P.; Roberts, B. P.; Chakravorty, D. K.; Merz, K. M., Jr. Rational Design of Particle Mesh Ewald Compatible Lennard-Jones Parameters for + 2 Metal Cations in Explicit Solvent. J. Chem. Theory Comput. 2013, 9, 2733– 2748, DOI: 10.1021/ct400146w47Rational Design of Particle Mesh Ewald Compatible Lennard-Jones Parameters for +2 Metal Cations in Explicit SolventLi, Pengfei; Roberts, Benjamin P.; Chakravorty, Dhruva K.; Merz, Kenneth M.Journal of Chemical Theory and Computation (2013), 9 (6), 2733-2748CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)Metal ions play significant roles in biol. systems. Accurate mol. dynamics (MD) simulations on these systems require a validated set of parameters. Although there are more detailed ways to model metal ions, the nonbonded model, which employs a 12-6 Lennard-Jones (LJ) term plus an electrostatic potential, is still widely used in MD simulations today due to its simple form. However, LJ parameters have limited transferability due to different combining rules, various water models, and diverse simulation methods. Recently, simulations employing a Particle Mesh Ewald (PME) treatment for long-range electrostatics have become more and more popular owing to their speed and accuracy. In the present work, we have systematically designed LJ parameters for 24 +2 metal (M(II)) cations to reproduce different exptl. properties appropriate for the Lorentz-Berthelot combining rules and PME simulations. We began by testing the transferability of currently available M(II) ion LJ parameters. The results showed that there are differences between simulations employing Ewald summation with other simulation methods and that it was necessary to design new parameters specific for PME based simulations. Employing the thermodn. integration (TI) method and performing periodic boundary MD simulations employing PME, allowed for a systematic investigation of the LJ parameter space. Hydration free energies (HFEs), the ion-oxygen distance in the first solvation shell (IOD), and coordination nos. (CNs) were obtained for various combinations of the parameters of the LJ potential for four widely used water models (TIP3P, SPC/E, TIP4P, and TIP4PEW). Results showed that the three simulated properties were highly correlated. Meanwhile, M(II) ions with the same parameters in different water models produce remarkably different HFEs but similar structural properties. It is difficult to reproduce various exptl. values simultaneously because the nonbonded model underestimates the interaction between the metal ions and water mols. at short-range. Moreover, the extent of underestimation increases successively for the TIP3P, SPC/E, TIP4PEW, and TIP4P water models. Nonetheless, we fitted a curve to describe the relationship between ε (the well depth) and radius (Rmin/2) from exptl. data on noble gases to facilitate the generation of the best possible compromise models. Hence, by targeting different exptl. values, we developed three sets of parameters for M(II) cations for three different water models (TIP3P, SPC/E, and TIP4PEW). These parameters we feel represent the best possible compromise that can be achieved using the nonbonded model for the ions in combination with simple water models. From a computational uncertainty anal. we est. that the uncertainty in our computed HFEs is on the order of ±1 kcal/mol. Further improvements will require more advanced nonbonded models likely with inclusion of polarization.
- 48Lee, J.; Cheng, X.; Swails, J. M.; Yeom, M. S.; Eastman, P. K.; Lemkul, J. A.; Wei, S.; Buckner, J.; Jeong, J. C.; Qi, Y. CHARMM-GUI Input Generator for NAMD, GROMACS, AMBER, OpenMM, and CHARMM/OpenMM Simulations Using the CHARMM36 Additive Force Field. J. Chem. Theory Comput. 2016, 12, 405– 413, DOI: 10.1021/acs.jctc.5b0093548CHARMM-GUI Input Generator for NAMD, GROMACS, AMBER, OpenMM, and CHARMM/OpenMM Simulations Using the CHARMM36 Additive Force FieldLee, Jumin; Cheng, Xi; Swails, Jason M.; Yeom, Min Sun; Eastman, Peter K.; Lemkul, Justin A.; Wei, Shuai; Buckner, Joshua; Jeong, Jong Cheol; Qi, Yifei; Jo, Sunhwan; Pande, Vijay S.; Case, David A.; Brooks, Charles L.; MacKerell, Alexander D.; Klauda, Jeffery B.; Im, WonpilJournal of Chemical Theory and Computation (2016), 12 (1), 405-413CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)Proper treatment of nonbonded interactions is essential for the accuracy of mol. dynamics (MD) simulations, esp. in studies of lipid bilayers. The use of the CHARMM36 force field (C36 FF) in different MD simulation programs can result in disagreements with published simulations performed with CHARMM due to differences in the protocols used to treat the long-range and 1-4 nonbonded interactions. In this study, we systematically test the use of the C36 lipid FF in NAMD, GROMACS, AMBER, OpenMM, and CHARMM/OpenMM. A wide range of Lennard-Jones (LJ) cutoff schemes and integrator algorithms were tested to find the optimal simulation protocol to best match bilayer properties of six lipids with varying acyl chain satn. and head groups. MD simulations of a 1,2-dipalmitoyl-sn-phosphatidylcholine (DPPC) bilayer were used to obtain the optimal protocol for each program. MD simulations with all programs were found to reasonably match the DPPC bilayer properties (surface area per lipid, chain order parameters, and area compressibility modulus) obtained using the std. protocol used in CHARMM as well as from expts. The optimal simulation protocol was then applied to the other five lipid simulations and resulted in excellent agreement between results from most simulation programs as well as with exptl. data. AMBER compared least favorably with the expected membrane properties, which appears to be due to its use of the hard-truncation in the LJ potential vs. a force-based switching function used to smooth the LJ potential as it approaches the cutoff distance. The optimal simulation protocol for each program has been implemented in CHARMM-GUI. This protocol is expected to be applicable to the remainder of the additive C36 FF including the proteins, nucleic acids, carbohydrates, and small mols.
- 49Humphrey, W.; Dalke, A.; Schulten, K. VMD: Visual Molecular Dynamics. J. Mol. Graph. 1996, 14, 33– 38, DOI: 10.1016/0263-7855(96)00018-549VDM: visual molecular dynamicsHumphrey, William; Dalke, Andrew; Schulten, KlausJournal of Molecular Graphics (1996), 14 (1), 33-8, plates, 27-28CODEN: JMGRDV; ISSN:0263-7855. (Elsevier)VMD is a mol. graphics program designed for the display and anal. of mol. assemblies, in particular, biopolymers such as proteins and nucleic acids. VMD can simultaneously display any no. of structures using a wide variety of rendering styles and coloring methods. Mols. are displayed as one or more "representations," in which each representation embodies a particular rendering method and coloring scheme for a selected subset of atoms. The atoms displayed in each representation are chosen using an extensive atom selection syntax, which includes Boolean operators and regular expressions. VMD provides a complete graphical user interface for program control, as well as a text interface using the Tcl embeddable parser to allow for complex scripts with variable substitution, control loops, and function calls. Full session logging is supported, which produces a VMD command script for later playback. High-resoln. raster images of displayed mols. may be produced by generating input scripts for use by a no. of photorealistic image-rendering applications. VMD has also been expressly designed with the ability to animate mol. dynamics (MD) simulation trajectories, imported either from files or from a direct connection to a running MD simulation. VMD is the visualization component of MDScope, a set of tools for interactive problem solving in structural biol., which also includes the parallel MD program NAMD, and the MDCOMM software used to connect the visualization and simulation programs, VMD is written in C++, using an object-oriented design; the program, including source code and extensive documentation, is freely available via anonymous ftp and through the World Wide Web.
- 50Park, I.-S.; Hausinger, R. P. Site-Directed Mutagenesis of Klebsiella aerogenes Urease: Identification of Histidine Residues that Appear to Function in Nickel Ligation, Substrate Binding, and Catalysis. Protein Sci. 1993, 2, 1034– 1041, DOI: 10.1002/pro.556002061650Site-directed mutagenesis of Klebsiella aerogenes urease: Identification of histidine residues that appear to function in nickel ligation, substrate binding, and catalysisPark, Il Seon; Hausinger, Robert P.Protein Science (1993), 2 (6), 1034-41CODEN: PRCIEI; ISSN:0961-8368.Comparison of six urease sequences revealed the presence of 10 conserved histidine residues (H96 in the γ subunit, H39 and H41 in β, and H134, H136, H219, H246, H312, H320, and H321 in the α subunit of the K. aerogenes enzyme). Each of these residues in K. aerogenes urease was substituted with alanine by site-directed mutagenesis, and the mutant proteins were purified and characterized in order to identify essential histidine residues and assign their roles. The γH96A, βH39A, βH41A, αH312A, and αH321A mutant proteins possess activities and nickel contents similar to wild-type enzyme, suggesting that these residues are not essential for substrate binding, catalysis, or metal binding. In contrast, the αH134A, αH136A, and αH246A proteins exhibit no detectable activity and possess 53%, 6%, and 21% of the nickel content of wild-type enzyme. These results are consistent with αH134, αH136, and αH246 functioning as nickel ligands. The αH219A protein is active and has nickel (∼1.9% and ∼80%, resp., when compared to wild-type protein) but exhibits a very high Km value (1,100 mM compared to 2.3 mM for the wild-type enzyme). These results are compatible with αH219 having some role in facilitating substrate binding. Finally, the αH320A protein (Km = 8.3 mM) only displays ∼0.003% of the wild-type enzyme activity, despite having a normal nickel content. Unlike the wild-type and αH219A ureases, this mutant protein was not inactivated by diethylpyrocarbonate (DEP), consistent with αH320 being the DEP-reactive general base that facilitates catalysis.
- 51Nelson, M. T.; Humphrey, W.; Gursoy, A.; Dalke, A.; Kalé, L. V.; Skeel, R. D.; Schulten, K. NAMD: A Parallel, Object-Oriented Molecular Dynamics Program. Int. J. High Perform. Comput. Appl. 1996, 10, 251– 268, DOI: 10.1177/109434209601000401There is no corresponding record for this reference.
- 52Darden, T.; York, D.; Pedersen, L. Particle Mesh Ewald: An N·log (N) Method for Ewald Sums in Large Systems. J. Chem. Phys. 1993, 98, 10089– 10092, DOI: 10.1063/1.46439752Particle mesh Ewald: an N·log(N) method for Ewald sums in large systemsDarden, Tom; York, Darrin; Pedersen, LeeJournal of Chemical Physics (1993), 98 (12), 10089-92CODEN: JCPSA6; ISSN:0021-9606.An N·log(N) method for evaluating electrostatic energies and forces of large periodic systems is presented. The method is based on interpolation of the reciprocal space Ewald sums and evaluation of the resulting convolution using fast Fourier transforms. Timings and accuracies are presented for three large cryst. ionic systems.
- 53Huang, J.; MacKerell, A. D., Jr CHARMM36 All-atom Additive Protein Force Field: Validation Based on Comparison to NMR Data. J. Comput. Chem. 2013, 34, 2135– 2145, DOI: 10.1002/jcc.2335453CHARMM36 all-atom additive protein force field: Validation based on comparison to NMR dataHuang, Jing; MacKerell, Alexander D. JrJournal of Computational Chemistry (2013), 34 (25), 2135-2145CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)Protein structure and dynamics can be characterized on the atomistic level with both NMR expts. and mol. dynamics (MD) simulations. Here, the authors quantify the ability of the recently presented CHARMM36 (C36) force field (FF) to reproduce various NMR observables using MD simulations. The studied NMR properties include backbone scalar couplings across hydrogen bonds, residual dipolar couplings (RDCs) and relaxation order parameter, as well as scalar couplings, RDCs, and order parameters for side-chain amino- and methyl-contg. groups. The C36 FF leads to better correlation with exptl. data compared to the CHARMM22/CMAP FF and suggest using C36 in protein simulations. Although both CHARMM FFs contains the same nonbond parameters, the authors' results show how the changes in the internal parameters assocd. with the peptide backbone via CMAP and the χ1 and χ2 dihedral parameters leads to improved treatment of the analyzed nonbond interactions. This highlights the importance of proper treatment of the internal covalent components in modeling nonbond interactions with mol. mechanics FFs.
- 54Bannwarth, C.; Ehlert, S.; Grimme, S. GFN2-xTB─An Accurate and Broadly Parametrized Self-Consistent Tight-Binding Quantum Chemical Method with Multipole Electrostatics and Density-Dependent Dispersion Contributions. J. Chem. Theory Comput. 2019, 15, 1652– 1671, DOI: 10.1021/acs.jctc.8b0117654GFN2-xTB-An Accurate and Broadly Parametrized Self-Consistent Tight-Binding Quantum Chemical Method with Multipole Electrostatics and Density-Dependent Dispersion ContributionsBannwarth, Christoph; Ehlert, Sebastian; Grimme, StefanJournal of Chemical Theory and Computation (2019), 15 (3), 1652-1671CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)An extended semiempirical tight-binding model is presented, which is primarily designed for the fast calcn. of structures and noncovalent interactions energies for mol. systems with roughly 1000 atoms. The essential novelty in this so-called GFN2-xTB method is the inclusion of anisotropic second order d. fluctuation effects via short-range damped interactions of cumulative at. multipole moments. Without noticeable increase in the computational demands, this results in a less empirical and overall more phys. sound method, which does not require any classical halogen or hydrogen bonding corrections and which relies solely on global and element-specific parameters (available up to radon, Z = 86). Moreover, the at. partial charge dependent D4 London dispersion model is incorporated self-consistently, which can be naturally obtained in a tight-binding picture from second order d. fluctuations. Fully anal. and numerically precise gradients (nuclear forces) are implemented. The accuracy of the method is benchmarked for a wide variety of systems and compared with other semiempirical methods. Along with excellent performance for the "target" properties, we also find lower errors for "off-target" properties such as barrier heights and mol. dipole moments. High computational efficiency along with the improved physics compared to it precursor GFN-xTB makes this method well-suited to explore the conformational space of mol. systems. Significant improvements are furthermore obsd. for various benchmark sets, which are prototypical for biomol. systems in aq. soln.
- 55Melo, M. C.; Bernardi, R. C.; Rudack, T.; Scheurer, M.; Riplinger, C.; Phillips, J. C.; Maia, J. D.; Rocha, G. B.; Ribeiro, J. V.; Stone, J. E. NAMD Goes Quantum: An Integrative Suite for Hybrid Simulations. Nat. Methods 2018, 15, 351– 354, DOI: 10.1038/nmeth.463855NAMD goes quantum: an integrative suite for hybrid simulationsMelo, Marcelo C. R.; Bernardi, Rafael C.; Rudack, Till; Scheurer, Maximilian; Riplinger, Christoph; Phillips, James C.; Maia, Julio D. C.; Rocha, Gerd B.; Ribeiro, Joao V.; Stone, John E.; Neese, Frank; Schulten, Klaus; Luthey-Schulten, ZaidaNature Methods (2018), 15 (5), 351-354CODEN: NMAEA3; ISSN:1548-7091. (Nature Research)Hybrid methods that combine quantum mechanics (QM) and mol. mechanics (MM) can be applied to studies of reaction mechanisms in locations ranging from active sites of small enzymes to multiple sites in large bioenergetic complexes. By combining the widely used mol. dynamics and visualization programs NAMD and VMD with the quantum chem. packages ORCA and MOPAC, we created an integrated, comprehensive, customizable, and easy-to-use suite (http://www.ks.uiuc.edu/Research/qmmm). Through the QwikMD interface, setup, execution, visualization, and anal. are streamlined for all levels of expertise.
- 56Bursch, M.; Neugebauer, H.; Grimme, S. Structure Optimisation of Large Transition-Metal Complexes with Extended Tight-Binding Methods. Angew. Chem., Int. Ed. 2019, 58, 11078– 11087, DOI: 10.1002/anie.20190402156Structure Optimisation of Large Transition-Metal Complexes with Extended Tight-Binding MethodsBursch, Markus; Neugebauer, Hagen; Grimme, StefanAngewandte Chemie, International Edition (2019), 58 (32), 11078-11087CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)Large transition-metal complexes are used in numerous areas of chem. Computer-aided theor. investigations of such complexes are limited by the sheer size of real systems often consisting of hundreds to thousands of atoms. Accordingly, the development and thorough evaluation of fast semiempirical quantum chem. methods that are universally applicable to a large part of the periodic table is indispensable. Herein, we report on the capability of the recently developed GFNn-xTB method family for full quantum-mech. geometry optimization of medium to very large transition-metal complexes and organometallic supramol. structures. The results for a specially compiled benchmark set of 145 diverse closed-shell transition-metal complex structures for all metals up to Hg are presented. Further the GFNn-xTB methods are tested on three established benchmark sets regarding reaction energies and barrier heights of organometallic reactions.
- 57Maruno, Y.; Shoji, M.; Koizumi, K.; Nishiyama, Y.; Kitagawa, Y.; Kawakami, T.; Okumura, M.; Yamaguchi, K. Theoretical Studies on Magnetic Interactions between Ni (II) Ions in Urease. Polyhedron 2005, 24, 2778– 2783, DOI: 10.1016/j.poly.2005.03.16157Theoretical studies on magnetic interactions between Ni(II) ions in ureaseMaruno, Yusuke; Shoji, Mitsuo; Koizumi, Kenichi; Nishiyama, Yusuke; Kitagawa, Yasutaka; Kawakami, Takashi; Okumura, Mitsutaka; Yamaguchi, KizashiPolyhedron (2005), 24 (16-17), 2778-2783CODEN: PLYHDE; ISSN:0277-5387. (Elsevier B.V.)The chem. species of bridge oxygen (WB) in active site of urease was examd. in terms of a magnetic interaction between Ni(II) ions. The effective exchange integrals (Jab) values between Ni(II) ions were calcd. by using three candidates of WB, i.e., H2O, OH- and O2-. The unrestricted hybrid d. functional theory (UHDFT) calcns. were carried out using the smallest models and realistic models. In the case of the smallest models, the exptl. Jab value (-6.3 cm-1) was a middle value of WB = H2O (-2.6 cm-1) and WB = OH- (-11.4 cm-1), while we could not obtain an appropriate electronic structure using the O2--bridge model. The Jab became -16.7 cm-1 by using more realistic OH--bridge model, while one spin d. on the Ni2 ion was delocalized to Asp360 in case of H2O-bridge model. Natural orbital analyses revealed that the magnetic orbital through Lys-bridge delocalized to Asp360 and it seemed to assoc. with the spin transfer. From the results, the OH--bridge seemed to be the most likely candidate for the structure of active site in urease.
- 58Kitagawa, Y.; Saito, T.; Ito, M.; Nakanishi, Y.; Shoji, M.; Koizumi, K.; Yamanaka, S.; Kawakami, T.; Okumura, M.; Yamaguchi, K. Geometry Optimization Method Based on Approximate Spin Projection and its Application to F2, CH2, CH2OO, and Active Site of Urease. Int. J. Quantum Chem. 2007, 107, 3094– 3102, DOI: 10.1002/qua.2145658Geometry optimization method based on approximate spin projection and its application to F2, CH2, CH2OO, and active site of ureaseKitagawa, Yasutaka; Saito, Toru; Ito, Masahide; Nakanishi, Yasuyuki; Shoji, Mitsuo; Koizumi, Kenichi; Yamanaka, Shusuke; Kawakami, Takashi; Okumura, Mitsutaka; Yamaguchi, KizashiInternational Journal of Quantum Chemistry (2007), 107 (15), 3094-3102CODEN: IJQCB2; ISSN:0020-7608. (John Wiley & Sons, Inc.)A new geometry optimization method based on an approx. spin projection (AP) procedure is proposed to eliminate a spin contamination effect in an optimized structure on a low spin (LS) state of a broken symmetry (BS) method. First, an energy gradient with the AP (AP gradient) is derived and it is applied to the geometry optimization of F2, CH2, and CH2OO in order to obtain their structures without the spin contamination. The optimization method corrects H-C-H angle of the CH2 about 10° in comparison with the BS method, and indicates that the spin contamination in optimized geometry of the BS LS states is considerably large. Next, an active site of Urease that consists of two Ni(II) ions is optimized by the AP method. The result indicates that structure of the active site is sensitive to a water mol. called W3.
- 59Frisch, M. J.; Gaussian 16, Revision B.01; Gaussian Inc.: Wallingford, CT, 2016.There is no corresponding record for this reference.
- 60Bussi, G.; Laio, A.; Parrinello, M. Equilibrium Free Energies from Nonequilibrium Metadynamics. Physi. Rev. Lett. 2006, 96, 090601, DOI: 10.1103/PhysRevLett.96.09060160Equilibrium Free Energies from Nonequilibrium MetadynamicsBussi, Giovanni; Laio, Alessandro; Parrinello, MichelePhysical Review Letters (2006), 96 (9), 090601/1-090601/4CODEN: PRLTAO; ISSN:0031-9007. (American Physical Society)We propose a new formalism to map history-dependent metadynamics in a Markovian process. We apply this formalism to model Langevin dynamics and det. the equil. distribution of a collection of simulations. We demonstrate that the reconstructed free energy is an unbiased est. of the underlying free energy and anal. derive an expression for the error. The present results can be applied to other history-dependent stochastic processes, such as Wang-Landau sampling.
- 61Castro, C. B.; Ferreira, M. P.; Caterina, G. M. N. Metalloenzyme Mechanisms Correlated to Their Turnover Number and Metal Lability. Curr. Res. Chem. Biol. 2021, 1, 100004, DOI: 10.1016/j.crchbi.2021.100004There is no corresponding record for this reference.
- 62Krajewska, B.; van Eldik, R.; Brindell, M. Temperature- and Pressure-Dependent Stopped-Flow Kinetic Studies of Jack Bean Urease. Implications for the Catalytic Mechanism. J. Biol. Inorg. Chem. 2012, 17, 1123– 1134, DOI: 10.1007/s00775-012-0926-862Temperature- and pressure-dependent stopped-flow kinetic studies of jack bean urease. Implications for the catalytic mechanismKrajewska, Barbara; van Eldik, Rudi; Brindell, MalgorzataJBIC, Journal of Biological Inorganic Chemistry (2012), 17 (7), 1123-1134CODEN: JJBCFA; ISSN:0949-8257. (Springer)Urease, a Ni-contg. metalloenzyme, features an activity that has profound medical and agricultural implications. The mechanism of this activity, however, has not been as yet thoroughly established. Accordingly, to improve its understanding, in this study we analyzed the steady-state kinetic parameters of the enzyme (jack bean), Km and kcat, measured at different temps. and pressures. Such an anal. is useful as it provides information on the mol. nature of the intermediate and transition states of the catalytic reaction. We measured the parameters in a noninteracting buffer using a stopped-flow technique at 15-35° and in the pressure range 5-132 MPa, the pressure-dependent measurements being the first of their kind performed for urease. While temp. enhanced the activity of urease, pressure inhibited the enzyme; the inhibition was biphasic. Analyzing Km provided the characteristics of the formation of the ES complex, and analyzing kcat, the characteristics of the activation of ES. From the temp.-dependent measurements, the energetic parameters were derived, i.e. thermodn. ΔHo and ΔSo for ES formation, and kinetic ΔH≠ and ΔS≠ for ES activation, while from the pressure-dependent measurements, the binding ΔVb and activation vols. were detd. The thermodn. and activation parameters obtained are discussed in terms of the current proposals for the mechanism of the urease reaction, and they are found to support the mechanism proposed by Benini et al., in which the Ni-Ni bridging hydroxide-not the terminal hydroxide-is the nucleophile in the catalytic reaction.
- 63El-Hefnawy, M. E.; Sakran, M.; Ismail, A. I.; Aboelfetoh, E. F. Extraction, Purification, Kinetic and Thermodynamic Properties of Urease from Germinating Pisum Sativum L. seeds. BMC Biochem. 2014, 15, 15, DOI: 10.1186/1471-2091-15-1563Extraction, purification, kinetic and thermodynamic properties of urease from germinating Pisum sativum L. seedsEl-Hefnawy, Mohamed E.; Sakran, Mohamed; Ismail, Ali I.; Aboelfetoh, Eman FahmyBMC Biochemistry (2014), 15 (), 15/1-15/8, 8 pp.CODEN: BBMIB3 ISSN:. (BioMed Central Ltd.)Background: Urease, one of the highly efficient known enzymes, catalyzes the hydrolysis of urea into ammonia and carbon dioxide. The present study aimed to ext. urease from pea seeds (Pisum sativum L). The enzyme was then purified in three consequence steps: acetone pptn., DEAE-cellulose ion-exchange chromatog. and gel filtration chromatog. (Sephacryl S-200 column). Results: The purifn. fold was 12.85 with a yield of 40%. The mol. wt. of the isolated urease was estd. by chromatog. to be 269,000 Daltons. Maximum urease activity (190 U/g) was achieved at the optimum conditions of 40°C and pH of 7.5 after 5 min of incubation. The kinetic parameters, Km and Vmax, were estd. by Lineweaver-Burk fits and found to be 500 mM and 333.3 U/g, resp. The thermodn. consts. of activation, ΔH, Ea, and ΔS, were detd. using Arrhenius plot and found to be 21.20 kJ/mol, 23.7 kJ/mol and 1.18 kJ/mol/K, resp. Conclusions: Urease was purified from germinating Pisum sativum L. seeds. The purifn. fold, yield and mol. wt. were detd. The effects of pH, concn. of enzyme, temp., concn. of substrate and storage period on urease activity were examd. This may provide an insight on the various aspects of the property of the enzyme. The significance of extg. urease from different sources could play a good role in understanding the metab. of urea in plants.
- 64Musiani, F.; Arnofi, E.; Casadio, R.; Ciurli, S. Structure-Based Computational Study of the Catalytic and Inhibition Mechanisms of Urease. J. Biol. Inorg. Chem. 2001, 6, 300– 314, DOI: 10.1007/s00775000020464Structure-based computational study of the catalytic and inhibition mechanisms of ureaseMusiani, Francesco; Arnofi, Elisa; Casadio, Rita; Ciurli, StefanoJBIC, Journal of Biological Inorganic Chemistry (2001), 6 (3), 300-314CODEN: JJBCFA; ISSN:0949-8257. (Springer-Verlag)The viability of different mechanisms of catalysis and inhibition of the nickel-contg. enzyme urease was explored using the available high-resoln. structures of the enzyme isolated from Bacillus pasteurii in the native form and inhibited with several substrates. The structures and charge distribution of urea, its catalytic transition state, and three enzyme inhibitors were calcd. using ab initio and d. functional theory methods. The DOCK program suite was employed to det. families of structures of urease complexes characterized by docking energy scores indicative of their relative stability according to steric and electrostatic criteria. Adjustment of the parameters used by DOCK, in order to account for the presence of the metal ion in the active site, resulted in the calcn. of best energy structures for the nickel-bound inhibitors β-mercaptoethanol, acetohydroxamic acid, and diamidophosphoric acid. These calcd. structures are in good agreement with the exptl. detd. structures, and provide hints on the reactivity and mobility of the inhibitors in the active site. The same docking protocol was applied to the substrate urea and its catalytic transition state, in order to shed light onto the possible catalytic steps occurring at the binuclear nickel active site. These calcns. suggest that the most viable pathway for urea hydrolysis involve a nucleophilic attack by the bridging, and not the terminal, nickel-bound hydroxide onto a urea mol., with active site residues playing important roles in orienting and activating the substrate, and stabilizing the catalytic transition state.
Supporting Information
Supporting Information
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.jpcb.1c10200.
Figure S1, active site models for 1 and 2 excluding H222, H323, and A366 with key distances and spin density; Cartesian coordinates of these models obtained with the UB3LYP/def2-SVP level; the created topology and parameter files; Figure S2, RMSDs with respect to the initial conformations during a 10 ns classical MD simulation for 1 and 2; Figure S3, overlay of smaller models (1, 2, and 3) optimized in the singlet and quintet states at GFN2-xTB with respect to the structure obtained with the broken-symmetry DFT method; Figure S4, structure and Cartesian coordinates of the QM region for 1 optimized in the open-shell singlet state at the UB3LYP/def2-SV(P) level; Figure S5, potential energy profile of the dissociation of W2; Cartesian coordinates for all stationary points of the QM-only cluster calculations optimized at the GFN2-xTB and UB3LYP levels; Tables S1–S5, key bond distance for all stationary points obtained from QM-only cluster and QM/MM metadynamics calculations; full citation for ref (59) (PDF)
Terms & Conditions
Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.