Network


Latest external collaboration on country level. Dive into details by clicking on the dots.

Hotspot


Dive into the research topics where Fahmi Himo is active.

Publication


Featured researches published by Fahmi Himo.


Chemical Reviews | 2014

Quantum Chemical Studies of Mechanisms for Metalloenzymes

Margareta R. A. Blomberg; Tomasz Borowski; Fahmi Himo; Rong-Zhen Liao; Per E. M. Siegbahn

ing a hydrogen atom from substrates with relatively weak X−H bonds (Figure 56). Such a situation was found in the catalytic cycle of ACCO, where cleavage of the N−H bond is facilitated by the ferrous ion capable of (partly) reducing the resulting N-based radical to an anion. In the catalytic cycle of HEPD and in the reaction of HppE with an enantiomer of the native substrate, a C−H bond at the carbon hosting a deprotonated alcohol group is severed by the superoxide. In these cases transfer of a hydrogen atom to the superoxide is coupled to one-electron reduction of Fe(III) to Fe(II), which yields aldehyde or ketone products. In a similar way a thioaldehyde is produced during the initial steps of the IPNS catalytic cycle. Figure 53. Reaction catalyzed by NDO. Figure 54. X-ray structure of the non-heme iron cofactor in the NDO active site (PDB 1O7G). Figure 55. Reaction mechanism for the dihydroxylation reaction catalyzed by NDO suggested on the basis of the results of a DFT study. Relative energy values in kilocalories per mole are given beneath the structures shown and above the arrows for the transition states connecting them. Chemical Reviews Review dx.doi.org/10.1021/cr400388t | Chem. Rev. 2014, 114, 3601−3658 3636 The most often encountered reaction of the Fe(III)−O2 species is an electrophilic attack on an electron-rich (co)substrate that yields an Fe(II) intermediate with a peroxide bridge between the ferrous ion and an organic molecule (Figure 57). In all examples shown in the figure the reactive spin state is a quintet with high-spin Fe(III) antiferromagnetically coupled to the superoxide radical. This particular electronic structure allows for a smooth two-electron reaction, leading to a ferrous intermediate featuring a high-spin Fe(II) ion, i.e., also lying on a quintet PES. When the substrates are easily one-electron-oxidized, as, for example, carotenoids or catecholates, already the ternary enzyme−dioxygen−organic substrate complex may contain an organic radical along with the superoxide anion stabilized by coordination to the metal (Figure 58). With proper (antiferromagnetic) alignment of the spins of the unpaired electrons on the two radicals, direct coupling between them proceeds with a straightforward formation of a new C−O bond. With the O2 ligand already protonated, several different reaction scenarios are usually plausible, and one of them, arguably the simplest, involves a transfer of the HOO group from the metal ion to the organic substrate (Figure 59). In the reaction of HEPD, it is the distal, i.e., originally protonated, oxygen atom that attacks the aldehyde group and the proton is transferred to the second (proximal) oxygen atom with the help of a phosphonic group of the substrate. In this reaction HOO acts as a nucleophile. In the case of HGD, the HOO ligand has a partial radical character, and hence, in its attack on the aromatic ring, it behaves as an electrophilic reagent. Moreover, in the HGD and ACO cases, it is the proximal oxygen atom that is directly transferred from the metal ion to the organic radical, i.e., the proton remains on the same oxygen atom throughout the reaction. When a catalytic reaction involves a homolytic O−O bond cleavage, one end of the peroxo group is in contact with the metal ion, and it is reduced to HO− or RO− when the O−O bond breaks (Figure 60). The electron required for the reduction is provided usually by the ferrous ion; in intradiol dioxygenases, which host Fe(III) in the active site, a tyrosinate ligand can serve as a reductant. Heterolytic cleavage of the O−O bond typically yields highvalent iron(IV)−oxo species, and such a reaction requires Fe(II), a deprotonated proximal oxygen atom, and usually also that the distal oxygen atom has a chance to develop a second covalent bond when the O−O gets broken (Figure 61). The bonding partner for the distal oxygen can be a hydrogen (PDHs, IPNS, ACCO) or a (co)substrate’s carbon (αKAOs, ACO, Dke1) atom. Heterolysis to Fe(IV)O usually proceeds on the quintet PES. Finally, heterolytic O−O bond cleavage may proceed without changing the oxidation state of the metal but instead with coupled two-electron oxidation of an organic substrate (Figure Figure 56. Examples of Fe(II)/Fe(III)−O2 reactions involving hydrogen atom abstraction. Only the most relevant fragment of the substrate is shown. Figure 57. Examples of Fe(III)−O2 reactions involving electrophilic attack and two-electron oxidation of an organic (co)substrate. Figure 58. Examples of Fe(II)/Fe(III)−O2 reactions involving radical coupling between the superoxide and an organic radical. Only the most relevant fragment of the substrate is shown. Chemical Reviews Review dx.doi.org/10.1021/cr400388t | Chem. Rev. 2014, 114, 3601−3658 3637 62). In both cases depicted in the figure, a hydroperoxo group is bound to high-spin Fe(III), and when theO−Obond cleaves, the OH group remains on iron and the other oxygen atom forms two covalent bonds with the organic substrate. This kind of heterolytic O−O cleavage requires that the donor orbital of the organic substrate, i.e., the one that provides two electrons for reduction of the peroxo group, has a good overlap with the O−O σ* orbital. 5.2. Dinuclear Non-Heme Iron Enzymes 5.2.1. Methane Monooxygenase. MMO is an enzyme which inserts one oxygen from O2 into methane to form methanol. The active site of the soluble form of MMO is shown in Figure 63. It contains an iron dimer complex linked by oxygen-derived ligands and has four glutamates and two histidines. Due to the presence of a well-resolved X-ray structure, and the technical importance of the reaction catalyzed, this was actually the first redox-active enzymemechanism that was treated with the cluster model using modern DFT functionals in 1997. Several groups were active at an early stage, and the most essential parts of the reaction mechanism were determined more than a decade ago. A comprehensive review of this development was written by Friesner et al. In short, the active species (compound Q) has a diamond core structure with two bridging oxo groups and is in an Fe2(IV,IV) state, one of themost oxidized species in nature. One of the oxo groups of compound Q activates nethane by an abstraction of one hydrogen atom. The TS is linear in C···H···O to form an Fe2(III,IV) state and amethyl radical. The loss of entropy at the TS is a large part of the barrier. In the second step, the methyl radical recombines with the bridging hydroxide formed in the first step. The TS for this step was first located by Basch et al. This rebound mechanism was criticized by interpretations of radical clock experiments, which appeared to show that there could not be a sufficiently long-lived alkyl radical to be consistent with a two-step mechanism. A concerted mechanism with simultaneous cleavage of the C−H bond and formation of the C−O bond was therefore suggested. Several explanations were suggested to resolve this apparent discrepancy between experiment and theory. In one of them, it was concluded that the radical clock probe molecules Figure 59. Examples of HOO transfer reactions. Only the most relevant fragment of the substrate is shown. Figure 60. Examples of O−O bond homolysis facilitated by oneelectron reduction of the proximal oxygen atom. Only the most relevant fragment of the substrate is shown. Figure 61. Examples of heterolytic O−O bond cleavage yielding iron(IV)−oxo species. Figure 62. Examples of heterolytic O−Obond cleavage proceeding with direct two-electron oxidation of the organic molecule. Chemical Reviews Review dx.doi.org/10.1021/cr400388t | Chem. Rev. 2014, 114, 3601−3658 3638 were chemically so different from methane that different mechanisms were likely. The probes are much easier to ionize and could form cations instead of radicals. Another suggestion was a so-called two-channel mechanism of the dynamics involving a bound radical intermediate. A third possibility could be something analogous to the two-state reactivity mechanism suggested for P450 to explain similar discrepancies in that case, but this has not been tested. In summary, a concerted mechanism, as suggested by the experiments, has never been found in DFT modeling calculations for methane hydroxylation by MMO, at least when reasonable models have been tried, and the suggestion has therefore not survived. In the initial phase of the MMO studies, there were problems converging to proper electronic states. These problems were solved by Friesner et al., who were able to obtain the correct antiferromagnetic coupling of compound Q with two high-spin irons. A state of key importance is also the first intermediate after Q, with an electronic structure characterized as Fe2(III,IV)− O•, discussed further below in connection with mixedMn−Fe dimers. It was found that already at the TS for hydrogen abstraction the iron dimer is in an Fe2(III,IV) state as it is in the product of this step. The bridging oxygen radical would then act as a hydrogen atom abstractor. At the TS, the spins are divided between a bridging oxo ligand and the methyl, while the iron spins stay essentially constant from the Fe2(III,IV)−O state to the product. The O−O bond cleavage to reach compound Q is also a significant step in the catalytic cycle. Again, several groups were involved in studying this step at an early stage. There was essential agreement among these studies on the mechanism. First, a peroxide (compound P) is formed between the two irons in an Fe2(III,III) state. Several different structures of P are nearly degenerate. In the TS for the O−O cleavage, the oxygens are symmetric, but only one of the irons is redox-active. In the final Figure 63. Active site of methane monooxygenase. Figure 64. R1 and R2 proteins in RNR. Chemical Reviews Review dx.doi.org/10.1021/cr400388t | Chem. Rev. 2014, 114, 3601−3658 3639 stage, the other iron also changes its oxidation state to IV, and compound Q is formed. In the study by Friesner et al.,


Journal of Biological Inorganic Chemistry | 2009

Recent developments of the quantum chemical cluster approach for modeling enzyme reactions

Per E. M. Siegbahn; Fahmi Himo

The quantum chemical cluster approach for modeling enzyme reactions is reviewed. Recent applications have used cluster models much larger than before which have given new modeling insights. One important and rather surprising feature is the fast convergence with cluster size of the energetics of the reactions. Even for reactions with significant charge separation it has in some cases been possible to obtain full convergence in the sense that dielectric cavity effects from outside the cluster do not contribute to any significant extent. Direct comparisons between quantum mechanics (QM)-only and QM/molecular mechanics (MM) calculations for quite large clusters in a case where the results differ significantly have shown that care has to be taken when using the QM/MM approach where there is strong charge polarization. Insights from the methods used, generally hybrid density functional methods, have also led to possibilities to give reasonable error limits for the results. Examples are finally given from the most extensive study using the cluster model, the one of oxygen formation at the oxygen-evolving complex in photosystem II.


Chemistry: A European Journal | 2013

Arylation with Unsymmetrical Diaryliodonium Salts : A Chemoselectivity Study

Joel Malmgren; Stefano Santoro; Nazli Jalalian; Fahmi Himo; Berit Olofsson

Phenols, anilines, and malonates have been arylated under metal-free conditions with twelve aryl(phenyl)iodonium salts in a systematic chemoselectivity study. A new “anti-ortho effect” has been identified in the arylation of malonates. Several “dummy groups” have been found that give complete chemoselectivity in the transfer of the phenyl moiety, irrespective of the nucleophile. An aryl exchange in the diaryliodonium salts has been observed under certain arylation conditions. DFT calculations have been performed to investigate the reaction mechanism and to elucidate the origins of the observed selectivities. These results are expected to facilitate the design of chiral diaryliodonium salts and the development of catalytic arylation reactions that are based on these sustainable and metal-free reagents.


Wiley Interdisciplinary Reviews: Computational Molecular Science | 2011

The quantum chemical cluster approach for modeling enzyme reactions

Per E. M. Siegbahn; Fahmi Himo

This Overview describes the general concepts behind the quantum chemical cluster approach for modeling enzyme active sites and reaction mechanisms. First, the underlying density functional electronic structure method is briefly recapitulated. The cluster methodology is then discussed, including the important observation on the convergence of the solvation effects. The concepts are illustrated using examples from recent applications, such as the discrimination between different reaction mechanisms in phosphotriesterase, the elucidation of origins of regioselectivity in the epoxide‐opening reaction of haloalcohol dehalogenase, and finally the use of the cluster methodology to establish the detailed structure of the oxygen‐evolving complex in photosystem II.


Biophysical Journal | 1997

Density functional calculations on model tyrosyl radicals

Fahmi Himo; Astrid Gräslund; Lars Eriksson

A gradient-corrected density functional theory approach (PWP86) has been applied, together with large basis sets (IGLO-III), to investigate the structure and hyperfine properties of model tyrosyl free radicals. In nature, these radicals are observed in, e.g., the charge transfer pathways in photosystem II (PSII) and in ribonucleotide reductases (RNRs). By comparing spin density distributions and proton hyperfine couplings with experimental data, it is confirmed that the tyrosyl radicals present in the proteins are neutral. It is shown that hydrogen bonding to the phenoxyl oxygen atom, when present, causes a reduction in spin density on O and a corresponding increase on C4. Calculated proton hyperfine coupling constants for the beta-protons show that the alpha-carbon is rotated 75-80 degrees out of the plane of the ring in PSII and Salmonella typhimurium RNR, but only 20-30 degrees in, e.g., Escherichia coli, mouse, herpes simplex, and bacteriophage T4-induced RNRs. Furthermore, based on the present calculations, we have revised the empirical parameters used in the experimental determination of the oxygen spin density in the tyrosyl radical in E. coli RNR and of the ring carbon spin densities, from measured hyperfine coupling constants.


Coordination Chemistry Reviews | 2003

Density functional methods applied to metalloenzymes

Timothy Lovell; Fahmi Himo; Wen-Ge Han; Louis Noodleman

Abstract Density functional calculations for structures, spin states, redox energetics and reaction pathways are presented for some selected metalloenzymes. The specific enzymes examined are: (1) Fe and Mn superoxide dismutase for redox energetics and the role of second shell residues; (2) galactose oxidase (Cu enzyme) and (3) glyoxalase I (Zn enzyme) for reaction pathways, mechanisms, intermediates, and transition states (reaction barriers); (4) iron-oxo dimer enzymes methane monooxygenase and ribonucleotide reductase for characterizing the oxidized and reduced forms in terms of structures and protonation states, and for a proposed structure for the high-valent intermediate Q in MMO. The interaction of the active site with the surrounding protein environment is also explored in a number of cases either by using expanded quantum mechanically treated clusters, or by using electrostatic/dielectric representations of the protein–solvent environment.


Journal of Inorganic Biochemistry | 2011

Ribonucleotide reductase inhibition by metal complexes of Triapine (3-aminopyridine-2-carboxaldehyde thiosemicarbazone): A combined experimental and theoretical study

Ana Popović-Bijelić; Christian R. Kowol; Maria E. S. Lind; Jinghui Luo; Fahmi Himo; Éva A. Enyedy; Vladimir B. Arion; Astrid Gräslund

Triapine (3-aminopyridine-2-carboxaldehyde thiosemicarbazone, 3-AP) is currently the most promising chemotherapeutic compound among the class of α-N-heterocyclic thiosemicarbazones. Here we report further insights into the mechanism(s) of anticancer drug activity and inhibition of mouse ribonucleotide reductase (RNR) by Triapine. In addition to the metal-free ligand, its iron(III), gallium(III), zinc(II) and copper(II) complexes were studied, aiming to correlate their cytotoxic activities with their effects on the diferric/tyrosyl radical center of the RNR enzyme in vitro. In this study we propose for the first time a potential specific binding pocket for Triapine on the surface of the mouse R2 RNR protein. In our mechanistic model, interaction with Triapine results in the labilization of the diferric center in the R2 protein. Subsequently the Triapine molecules act as iron chelators. In the absence of external reductants, and in presence of the mouse R2 RNR protein, catalytic amounts of the iron(III)-Triapine are reduced to the iron(II)-Triapine complex. In the presence of an external reductant (dithiothreitol), stoichiometric amounts of the potently reactive iron(II)-Triapine complex are formed. Formation of the iron(II)-Triapine complex, as the essential part of the reaction outcome, promotes further reactions with molecular oxygen, which give rise to reactive oxygen species (ROS) and thereby damage the RNR enzyme. Triapine affects the diferric center of the mouse R2 protein and, unlike hydroxyurea, is not a potent reductant, not likely to act directly on the tyrosyl radical.


Proceedings of the National Academy of Sciences of the United States of America | 2010

Mechanism of tungsten-dependent acetylene hydratase from quantum chemical calculations

Rong-Zhen Liao; Jian-Guo Yu; Fahmi Himo

Acetylene hydratase is a tungsten-dependent enzyme that catalyzes the nonredox hydration of acetylene to acetaldehyde. Density functional theory calculations are used to elucidate the reaction mechanism of this enzyme with a large model of the active site devised on the basis of the native X-ray crystal structure. Based on the calculations, we propose a new mechanism in which the acetylene substrate first displaces the W-coordinated water molecule, and then undergoes a nucleophilic attack by the water molecule assisted by an ionized Asp13 residue at the active site. This is followed by proton transfer from Asp13 to the newly formed vinyl anion intermediate. In the subsequent isomerization, Asp13 shuttles a proton from the hydroxyl group of the vinyl alcohol to the α-carbon. Asp13 is thus a key player in the mechanism, but also W is directly involved in the reaction by binding and activating acetylene and providing electrostatic stabilization to the transition states and intermediates. Several other mechanisms are also considered but the energetic barriers are found to be very high, ruling out these possibilities.


Journal of Computational Chemistry | 2010

Quantum chemical modeling of enzymatic reactions: the case of histone lysine methyltransferase.

Polina Georgieva; Fahmi Himo

Quantum chemical cluster models of enzyme active sites are today an important and powerful tool in the study of various aspects of enzymatic reactivity. This methodology has been applied to a wide spectrum of reactions and many important mechanistic problems have been solved. Herein, we report a systematic study of the reaction mechanism of the histone lysine methyltransferase (HKMT) SET7/9 enzyme, which catalyzes the methylation of the N‐terminal histone tail of the chromatin structure. In this study, HKMT SET7/9 serves as a representative case to examine the modeling approach for the important class of methyl transfer enzymes. Active site models of different sizes are used to evaluate the methodology. In particular, the dependence of the calculated energies on the model size, the influence of the dielectric medium, and the particular choice of the dielectric constant are discussed. In addition, we examine the validity of some technical aspects, such as geometry optimization in solvent or with a large basis set, and the use of different density functional methods.


Journal of Physical Chemistry B | 2008

Peptide hydrolysis by the binuclear zinc enzyme aminopeptidase from Aeromonas proteolytica: a density functional theory study.

Shi-Lu Chen; Tiziana Marino; Wei-Hai Fang; Nino Russo; Fahmi Himo

Aminopeptidase from Aeromonas proteolytica (AAP) is a binuclear zinc enzyme that catalyzes the cleavage of the N-terminal amino acid residue of peptides and proteins. In this study, we used density functional methods to investigate the reaction mechanism of this enzyme. A model of the active site was constructed on the basis of the X-ray crystal structure of the native enzyme, and a model dipeptide was used as a substrate. It was concluded that the hydroxide is capable of performing a nucleophilic attack at the peptide carbonyl from its bridging position without the need to first become terminal. The two zinc ions are shown to have quite different roles. Zn2 binds the amino group of the substrate, thereby orienting it toward the nucleophile, while Zn1 stabilizes the alkoxide ion of the tetrahedral intermediate, thereby lowering the barrier for the nucleophilic attack. The rate-limiting step is suggested to be the protonation of the nitrogen of the former peptide bond, which eventually leads to the cleavage of the C-N bond.

Collaboration


Dive into the Fahmi Himo's collaboration.

Top Co-Authors

Avatar

Rong-Zhen Liao

Beijing Normal University

View shared research outputs
Top Co-Authors

Avatar
Top Co-Authors

Avatar
Top Co-Authors

Avatar
Top Co-Authors

Avatar

Louis Noodleman

Scripps Research Institute

View shared research outputs
Top Co-Authors

Avatar
Top Co-Authors

Avatar

Jian-Guo Yu

Beijing Normal University

View shared research outputs
Top Co-Authors

Avatar
Top Co-Authors

Avatar

Peter Hammar

Royal Institute of Technology

View shared research outputs
Top Co-Authors

Avatar

Valery V. Fokin

Scripps Research Institute

View shared research outputs
Researchain Logo
Decentralizing Knowledge