electronic and catalytic properties of iron porphyrin ...electronic and catalytic properties of iron...

287
Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the University of Manchester for the degree of Doctor of Philosophy (PhD) in the Faculty of Engineering and Physical Sciences 2015 Mala Alhaji Sainna* Supervisor: Sam P. de Visser School of Chemical Engineering and Analytical Science

Upload: others

Post on 25-Jan-2020

6 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

Electronic and catalytic properties of iron porphyrin

complexes: Trends and reaction mechanisms.

A thesis submitted to the University of Manchester for the degree of

Doctor of Philosophy (PhD)

in the Faculty of Engineering and Physical Sciences

2015

Mala Alhaji Sainna*

Supervisor: Sam P. de Visser

School of Chemical Engineering and Analytical Science

Page 2: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

2

Table of Contents LIST OF FIGURES ................................................................................................................................ 5

LIST OF SCHEMES ............................................................................................................................. 11

LIST OF TABLES ................................................................................................................................ 13

ACKNOWLEDGEMENT .................................................................................................................... 14

ABSTRACT .......................................................................................................................................... 15

DECLARATION .................................................................................................................................. 16

COPY RIGHT STATEMENT .............................................................................................................. 17

CHAPTER 1 ......................................................................................................................................... 18

Introduction ........................................................................................................................................... 19

1.1 Preface................................................................................................................................... 19

1.2 Homogeneous Catalysis ........................................................................................................ 21

1.2.1 Ligand effects. ............................................................................................................... 23

1.2.2 Bite angle effect ............................................................................................................. 24

1.2.3 Cone angle effect ........................................................................................................... 26

1.3 Ligand according to donor atoms. ......................................................................................... 27

1.3.1 Imido and Alkoxy groups as an anionic ligand. ........................................................... 27

1.3.2 Neutral and Anionic hydrocarbyl groups ..................................................................... 28

1.3.3 Oxazolines, imines , Amines and related ligands .......................................................... 28

1.3.4 Carbon monoxide and Carbene .................................................................................... 30

1.4 Common anions .................................................................................................................... 32

1.5 Vacant site Creation and coordination of the substrate ......................................................... 33

1.6 Insertion versus migration ..................................................................................................... 34

1.6.1 β-Elimination and de-insertion: .................................................................................... 39

1.6.2 Oxidative addition and Reductive elimination: ............................................................. 41

1.7 Activation of a substrate toward nucleophilic attack ...................................................... 41

1.7.1 Role of Alkenes and alkynes .......................................................................................... 41

1.8 Metalloenzymes ................................................................................................................... 43

1.9 Epoxidation .......................................................................................................................... 48

Page 3: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

3

1.10 Hydroxylation ...................................................................................................................... 50

1.11 Dissertation summary ......................................................................................................... 52

CHAPTER 2 ......................................................................................................................................... 57

METHODOLOGY ............................................................................................................................... 57

2.1 Solving the Schrodinger equation ......................................................................................... 59

2.2 Born-oppenheimer approximation ........................................................................................ 61

2.3 Slater determinant ................................................................................................................. 63

2.3.1 Anti-symmetry Principle ............................................................................................... 63

2.4 Hartree-Fock theory (HF) ..................................................................................................... 64

2.5 Electron Correlation .............................................................................................................. 66

2.6 Density Functional Theory (DFT) ........................................................................................ 67

2.7 Functionals ............................................................................................................................ 70

2.7.1 Local Density Approximation (LDA) ............................................................................ 70

2.7.2 Generalized Gradient Approximation (GGA) ............................................................... 71

2.7.3 Hybrid Functionals ....................................................................................................... 71

2.8 Basis Sets .............................................................................................................................. 72

2.8.1 Slater-type Orbital (STO) .............................................................................................. 72

2.8.2 Gaussian-type orbital (GTO) ........................................................................................ 73

2.8.3 Polarisation and Diffuse basis functions ...................................................................... 75

2.9 Zero-Point Energy ................................................................................................................. 76

2.10 Transition states .................................................................................................................... 77

2.11 Computational Software ....................................................................................................... 78

CHAPTER 3 ......................................................................................................................................... 79

ABSTRACT ......................................................................................................................................... 80

3.1 Introduction ........................................................................................................................... 81

3.2 Methods................................................................................................................................. 85

3.3 Results and Discussion ......................................................................................................... 87

3.4 Conclusion ............................................................................................................................ 99

Page 4: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

4

ABSTRACT ....................................................................................................................................... 101

4.1 Introduction ......................................................................................................................... 102

4.2 Methods............................................................................................................................... 106

4.2.4 COMPUTATION ............................................................................................................. 106

4.3 Results ................................................................................................................................. 108

4.4 Theoretically derived reaction paths, energetics and structures. ......................................... 111

4.5 Discussion ........................................................................................................................... 119

4.6 Conclusion .......................................................................................................................... 128

CHAPTER 5 ....................................................................................................................................... 129

ABSTRACT ....................................................................................................................................... 130

5.1 Introduction. ........................................................................................................................ 131

5.2 Methods............................................................................................................................... 134

5.3 Results ................................................................................................................................. 136

5.4 Discussion ........................................................................................................................... 149

5.5 Conclusion .......................................................................................................................... 161

CHAPTER 6 ....................................................................................................................................... 162

ABSTRACT ....................................................................................................................................... 163

6.1 Introduction ......................................................................................................................... 164

6.2 Methods............................................................................................................................... 169

6.2.1. DFT model calculations .............................................................................................. 169

6.2.2. QM/MM calculations .................................................................................................. 170

6.3 Results and Discussion ....................................................................................................... 173

6.3.1. DFT model reactions .................................................................................................. 173

6.3.3. QM/MM studies of desaturation ................................................................................. 180

6.3.4. Valence Bond rationalization of rebound and desaturation mechanisms ................... 184

6.3.5. Molecular orbital rationalization of rebound and desaturation mechanisms ............ 188

6.4 Conclusion .......................................................................................................................... 191

CHAPTER 7 ....................................................................................................................................... 192

Page 5: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

5

ABSTRACT ....................................................................................................................................... 193

7.1 Introduction ......................................................................................................................... 194

7.2 Methods............................................................................................................................... 198

7.3 Results and Discussion ....................................................................................................... 201

7.4 Conclusions ......................................................................................................................... 218

ABSTRACT ....................................................................................................................................... 220

8.1 Introduction ......................................................................................................................... 221

8.2 Methods............................................................................................................................... 226

8.3 Results ................................................................................................................................. 227

8.4 Discussion ........................................................................................................................... 238

8.5 Conclusion .......................................................................................................................... 248

9.1 Concluding remarks ............................................................................................................. 249

References .......................................................................................................................................... 251

LIST OF FIGURES

Figure 1.1: Effect of ligands and valence states on the selectivity in a metal centred catalysed

reaction of butadiene. ............................................................................................................... 22

Figure 1.2: PyBox Ligand from an oxazoline group. .............................................................. 30

Figure 1.3: Example of a typical carbene Ligand .................................................................... 31

Figure 1.4: Dissociative and associative ligand exchange ....................................................... 34

Figure 1.5: insertion mechanism .............................................................................................. 35

Figure 1.6: migration mechanism ............................................................................................ 36

Figure 1.7: disproval of “outersphere” insertion...................................................................... 37

Figure 1.8: charge distribution in a migration reaction............................................................ 38

Figure 1.9: β – hydride elimination.......................................................................................... 39

Page 6: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

6

Figure 1.10: Migratory de - insertion ....................................................................................... 40

Figure 1.11: Nucleophilic attack to a coordinated alkene ........................................................ 42

Figure 1.12: Substitution of metal in cofactor to incorporate alternative metal into a binding

site. The Apo-protein must be synthesised first before introducing the new metal. Picture is

obtained from protein data bank 1N2C (Schindelin, H.; Kisker, C. et al 1997). ..................... 44

Figure 1.13: Schematic representation of the catalytic cycle of a typical substrate

hydroxylation reaction of substrate R–H by P450 enzymes and a visual depiction of the active

species (Highlighted in a rectangular box) (Nam, W., et al., 2000). ...................................... 46

Figure 1.14: Reactivity patterns of P40 Compound I with selected substrates. ...................... 48

Figure 1.15: Potential energy profile for the epoxidation reaction of 2-butene by 4,2

[Fe(IV)-

oxo Porphyrin with a thiolate axial ligand as calculated with DFT methods. Data taken from

ref. (de Visser 2012).Values in Kcal/mol. ............................................................................... 49

Figure 1.16: High-lying occupied and low lying virtual orbitals of Compound I at an

optimised geometry of P450 Cpd I of quartet and doublet spin electron arrangements. ......... 51

Figure 2.1: A Potential Energy Surface (PES) indicating a reaction from reactant to product

that passes through a transition state (TS). .............................................................................. 78

Figure 3.1: Relative spin state energies of (5/2,

5/2), (

5/2,

3/2) and (

3/2,

3/2) states of structure 1b

as calculated with various DFT/BS2 methods in Gaussian. Calculations done with basis set

BS3 labeled with superscript a. (a) E values relative to the S = (5/2,

5/2) state. (b) E+ZPE

values relative to the S = (5/2,

5/2) state. ................................................................................... 89

Figure 3.2: Quartet/sextet spin state energies for (A) individual core I and (B) core II. All

energies are relative to the quartet spin state in kcal mol–1

. A negative value implies an S = 5/2

ground state. ............................................................................................................................. 92

Figure 3.3: Optimized geometries of the S = (5/2,

5/2), [S = (

5/2,

3/2)] and {S = (

3/2,

3/2)} states

of 1a•X and 1b•X with X = I3–, BF4

– or ClO4

– at the B3LYP/BS1 level of theory in Jaguar.

Bond lengths are given in angstroms and is the average deviation of the iron from the plane

through the four nitrogen atoms. .............................................................................................. 95

Page 7: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

7

Figure 4.1 Active site of P450 as taken from the 2WM4 pdb file. ........................................ 104

Figure 4.2: Time dependence of relative ion abundancies for the reaction of

[FeIV

(O)(TPFPP+•

)]+ (m/z 1044) with indene. Product ions are [Fe

III(TPFPP)]

+ (m/z 1028),

[Fe(TPFPP)(C9H8)O]+ (m/z 1160) and C9H8

+• (m/z 116). Experiments were performed in the

presence of indene at 5.2 10–8

mbar in the FT-ICR cell. .................................................... 110

Figure 4.3: Molecular valence orbitals of 4A. ........................................................................ 112

Figure 4.4: Optimized geometries of the 4,2

A2u and 4,2

A1u states of 4,2

A as calculated

with UB3LYP/BS1 [UB3LYP/BS2] {UB3LYP-D3/BS2} with Fe–O bond lengths in

angstroms. ............................................................................................................................. 116

Figure 4.5: UB3LYP/BS1 optimized geometries of epoxidation transition states with

bond lengths in angstroms. ................................................................................................. 118

Figure 4.6: Correlation between experimental and computational barrier heights. ............... 121

Figure 4.7: (a) Correlation between experimentally determined RT ln kexp (for raw data,

see Table 1) versus known ionization energies (IE). (b) Correlation between calculated

epoxidation activation enthalpy (in kcal mol–1

) and experimental ionization energy for

the substrates in Fig 5. ......................................................................................................... 122

Figure 4.8: VB curve crossing diagram for the C–O bond formation step in olefin

epoxidation (R2C=CH2) by [FeIV

(O)(TPFPP+•

)]+. Valence electrons are identified with

a dot and lines (curved and straight) in the VB structures represent bonds. ................ 124

Figure 4.9: Correlation between calculated epoxidation activation enthalpy (in kcal

mol–1

) and BDECH for the substrates. ................................................................................ 127

Figure 5.1: Oxidants and substrates used in this work. .......................................................... 134

Figure 5.2: UB3LYP/BS1 optimized geometries of 4,2

1X and 4,2

2X in the gas phase with bond

lengths in angstroms. Group spin densities are obtained at UB3LYP/BS2//UB3LYP/BS1 and

are reported in atomic units. ................................................................................................... 138

Figure 5.3: Potential energy profile of styrene epoxidation by 4,2

1Cl as calculated with

UB3LYP/BS2//UB3LYP/BS1. All energies are in kcal mol–1

relative to isolated reactants in

the doublet spin state and include ZPE corrections. Also shown are optimized geometries of

critical points with bond lengths in angstroms and the imaginary frequency in the transition

Page 8: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

8

state in wave numbers. Free energies are given in parenthesis and include UB3LYP/BS2

energies corrected with thermal and entropic corrections at 298 K. Data in square brackets

obtained after a UB3LYP-D/BS1 geometry optimization. .................................................... 140

Figure 5.4: Optimized geometries of rate determining transition states 2TSX,Z for the reaction

of para-Z-styrene with 21X (X = Cl

–/NCCH3). Geometries optimized at UB3LYP/BS1 with

bond lengths given in angstroms and the imaginary frequency in wave numbers. Also given

are barrier heights (E+ZPE) for 2TSX,Z with energies calculated at

UB3LYP/BS2//UB3LYP/BS1+ZPE relative to isolated reactants in kcal mol–1

and free

energies of activation in solvent (G+Esolv+Edisp) relative to 2RCX,Z. ................................... 142

Figure 5.5: Analysis of structural features of the transition states TSX,Z calculated at

UB3LYP/BS1 as a function of the height of the epoxidation barrier with respect to the: (a) C–

O distance, (b) Fe–X distance, and (c) imaginary frequency in the transition state. Data given

for X = Cl– axial ligand (diamonds) and X = acetonitrile (squares). ..................................... 145

Figure 5.6: Optimized geometries of epoxidation transition states 2TS′X,Z for the reaction of

2X (X = Cl–/NCCH3) with para-Z-styrene. Bond lengths are in angstroms and the value of the

imaginary frequency in wave numbers. ................................................................................. 147

Figure 5.7: UB3LYP/BS2//UB3LYP/BS1 calculated group spin density ranges for para-Z-

styrene epoxidation by 1Cl (top) and 1AN (bottom). ............................................................... 150

Figure 5.8: Styrene epoxidation barrier heights (E‡+ZPE) of

2TSX,Z plotted against the

ionization energy of the corresponding substrate. (a) Energies relative to isolated reactants.

(b) Energies relative to a reactant complex (RC)................................................................... 153

Figure 5.9: Valence bond curve crossing diagram for para-Z-styrene epoxidation by iron(IV)-

oxo porphyrin cation radical oxidants. Lewis structures give relevant valence-orbitals with

a dot. ....................................................................................................................................... 154

Figure 5.10: Correlations of (a) Epoxidation barrier height of para-NO2-styrene with BDEOH.

(b) Epoxidation barrier height of para-N(CH3)2-styrene with EACpdI(X). .............................. 158

Figure 5.11: Correlation between epoxidation barrier height of all data for 1X (X = Cl–,

NCCH3) with parameter .................................................................................................... 160

Figure 6.1: Extract of the active site of P4502C19 as taken from the 4GQS pdb file. .......... 166

Page 9: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

9

Figure 6.2: Potential energy profile of ethylcarbamate activation by 4,2

Cpd I of P450 as

calculated with DFT. Energies are given in kcal mol–1

and are calculated at

UB3LYP/BSII//UB3LYP/BSI level of theory with ZPE and solvent corrections included.

Values in parenthesis are free energies in solvent. Optimized geometries give bond lengths in

angstroms and the imaginary frequency in the transition states in wave numbers. ............... 174

Figure 6.3: Optimized geometries of hydrogen atom abstraction transition states of VA, ET

and DHA by 4,2

Cpd I of P450 as calculated with DFT. Bond lengths are in angstroms, the

imaginary frequency is in wave numbers and (free) energies are given in kcal mol–1

and are

calculated at UB3LYP/BSII//UB3LYP/BSI level of theory with ZPE and solvent corrections

included. ................................................................................................................................. 177

Figure 6.4: Potential energy profile of eicosanoic acid activation by 2Cpd I of P450 as

calculated with QM/MM. Energies are given in kcal mol–1

and are calculated at

UB3LYP/BSIV//UB3LYP/BSIII level of theory with ZPE corrections included. Optimized

geometries give bond lengths in angstroms and angles in degrees. ....................................... 180

Figure 6.5: Geometry scans for the rotation along the Fe–O bond from 2I as calculated

with QM/MM. Energies are given in kcal mol–1

and each data point represents a full

geometry optimization with fixed H–O–Fe–Nheme dihedral angle. Also shown are the maxima

of the scans with key hydrogen bonding interactions identified. The atom labelled with a

yellow star is C. .................................................................................................................... 182

Figure 6.6: Valence bond curve crossing diagrams for product formation from radical

intermediates. (a) Radical rebound leading to hydroxylation products. (b) Hydrogen atom

transfer to give desaturation products. ................................................................................... 184

Figure 7.1: Optimized geometry of 3 as calculated with B3LYP. Bond lengths are given in

angstroms and group NBO charges Q in atomic units. The right-hand-side displays the

natural bond orbitals and their ordering for those involving the central carbon atom with it

ligands. ................................................................................................................................... 202

Figure 7.2: Free energy profile of alkyl chain growth on 3 via either (i) CH3-transfer followed

by H-transfer (mechanism from the center to the left) or (ii) H-transfer followed by CH3-

transfer (mechanism from the center to the right). Free energies (in kcal mol–1

) are obtained

Page 10: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

10

with B3LYP-D3/BS2 and contain zero-point, thermal and entropic corrections at 298K.

Values in parenthesis include solvent corrections to the free energy. Optimized geometries

report bond lengths in angstroms, angles in degrees and the imaginary frequency in the

transition states in wave numbers. ......................................................................................... 204

Figure 7.3: Energies of initial H-atom or CH3-transfer reactions from 3 as calculated with

different DFT methods. All structures optimized at B3LYP/BS1 and single point calculations

with basis set BS2 applied. Relative energies are given in kcal mol–1

and include ZPE

corrections calculated at B3LYP/BS1. .................................................................................. 207

Figure 7.4: Optimized geometries of 3’, IMe’ and IH’ with bond lengths in angstroms. ........ 209

Figure 7.5: Valence bond curve crossing diagrams for methyl transfer (part a) and hydrogen

atom transfer (part b) from 3. Valence electrons are identified with a dot. ........................... 211

Figure 7.6: Bond dissociation free energies (BDFEs in kcal mol–1

) of key bonds in structures

3, IMe and IH. Reactions calculated according to Eqs 1 – 4. Part (a) gives adiabatic BDFE

values and part (b) diabatic BDFE values. Values in parenthesis are solvent corrected free

energies, whereas those out of parenthesis are gas-phase data. ............................................. 213

Figure 7.7: Energy decomposition of the methyl and hydrogen atom transfer reaction from 3.

Free energies given are in kcal mol–1

. .................................................................................... 215

Figure 8.1: Active site structure of P450 with key amino acids and substrate (camphor) and

solvent water (W) highlighted. Amino acids labelled as in the pdb file. ............................... 223

Figure 8.2: High-lying occupied and low-lying virtual orbitals of 4RMn. Orbital energies are

reported in au. ........................................................................................................................ 228

Figure 8.3: Optimized geometries with bond lengths in angstroms of 4RMn and

5RFe; group

spin densities () taken from UB3LYP/BS2 calculations. Also given are relative energies

(including ZPE and solvent corrections in kcal mol–1

) of all low lying spin states for RMn and

RFe. ......................................................................................................................................... 231

Figure 8.4: (a) Potential energy profile of hydrogen atom abstraction from DHA by 4,6,2

RMn as

calculated using DFT methods with energies in kcal mol–1

relative to the quartet spin reactant

complex. Energies are taken from the UB3LYP/BS2 calculations in the gas-phase, while

solvent corrected values are in parenthesis. Free energies with solvent, entropic and

Page 11: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

11

dispersion corrections are given in square brackets. (b) Optimized geometries of the transition

states for hydrogen atom abstraction from DHA and CHD with bond lengths in angstroms

and the imaginary frequency in the transition state in cm–1

. .................................................. 234

Figure 8.5: Optimized geometries of 4RX structures with different substituents X with bond

lengths in angstroms and the vibrational frequency in wave numbers. Note that 4RMn has X =

t-Bu. ....................................................................................................................................... 236

Figure 8.6: (a) Barrier heights (E‡+ZPE in solvent) for H-atom abstraction from DHA by

various complexes manganese based complexes RX (X = t-Bu, i-Pr, Et, Me or H). (b) Barrier

heights as a function of the Mn–O frequency (MnO) in the reactant complex. ..................... 237

Figure 8.7: Extracts of the active site environments of nonheme iron dioxygenases

representing from left-to-right: TauD (1OS7 pdb), AlkB (3I2O pdb) and CDO (2IC1 pdb).

Amino acids are labelled as in the pdb file. ........................................................................... 243

Figure 8.8: (a) Orientation of substrate attack on the metal(IV)-oxo group with angles in

degrees and group spin densities in au. (b) Electron transfer processes and LUMO orbital that

is filled with one electron in the H-abstraction process. ........................................................ 243

LIST OF SCHEMES

Scheme 3.1: Structures investigated in this work. ................................................................... 84

Scheme 4.1: Models investigated in this work. ..................................................................... 106

Scheme 4.2: Substrates investigated in this work. ................................................................. 108

Scheme 4.3: Pathways observed for the reaction of [FeIV

(O)(TPFPP+•

)]+ ions (R = C6F5) with

selected substrates (Sub) as studied with FT-ICR MS. ......................................................... 110

Scheme 6.1: (a) Competitive hydroxylation and desaturation metabolism pathways of drug

molecules by P450 Cpd I. (b) Reaction products observed for valproic acid and

ethylcarbamate. ...................................................................................................................... 168

Scheme 6.2: Atoms in the QM region of the QM/MM calculation. Wiggly lines represent the

cuts between the QM and MM regions. ................................................................................. 173

Scheme 6.3: Reaction Mechanism of Ethyl Carbamate Activation by Cpd I of P450. ......... 173

Page 12: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

12

Scheme 6.4: Orbital mixing patterns for the pathways from radical intermediates to products.

................................................................................................................................................ 189

Scheme 7.1: Catalyst investigated in this work for alkyl formation on a carbide center. ..... 196

Scheme 7.2: Hybridization scheme of 3. ............................................................................... 203

Scheme 7.3: VB description of second reaction steps. (a) hydrogen-atom abstraction. (b)

methyl-transfer. ...................................................................................................................... 216

Scheme 8.1: Models studied in this work. ............................................................................. 225

Page 13: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

13

LIST OF TABLES

Table 3.1. Relative energies of optimized geometries in different spin states of 1a•X and

1b•X complexes.a ..................................................................................................................... 96

Table 4.1 Kinetic data and product distributions obtained fro the gas phase reaction of

[FeIV

(O)(TPFPP+•

)]+ with selected olefins as determined by FT-ICR

MS……………………......109

Table 4.2. Relative energies of several low-lying electronic states of [Fe(O)(Por+•)]+

(A).a.......................................................................................................................................113

Table 5.1. Substrate chemical properties and charge-transfer (QCT) values in the transition

states. ...................................................................................................................................... 160

Table 6.1. Free energies of activation of hydrogen atom abstraction, rebound and desaturation

barriers.a ................................................................................................................................. 178

Table 8.1. Calculated barrier heights for hydrogen atom abstraction from dehydroanthracene

and 1,4-cyclohexadiene by metal-oxo complexes (Energies are in kcal mol–1

). ................... 244

WORD COUNT: 69,110

Page 14: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

14

ACKNOWLEDGEMENT

My deepest appreciation goes to my distinguish supervisor Samuël P. de Visser who possess

the manner, support and the essence of an intellect: he persistently conveyed strength of

adventure in consideration to research and brotherly support; without his guidance and

persistent help this research would not have been possible.

I am heartily thankful to Mr. Balarabe Z. Ahmed for giving me the moral support I require.

I also thank the National Service of computational chemistry software (NSCCS) for a

generous CPU time.

I cannot exempt my appreciation from my research colleagues within the MIB building and

the entire University faculty and the support staff for their help and patience.

And finally I would also like to take this opportunity to thank Petroleum Technology

Development Fund (PTDF) for its generosity in funding my study; I am very honoured to be

a recipient of this award.

Page 15: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

15

ABSTRACT

The University of Manchester,

School of Chemical Engineering and Analytical Science

ABSTRACT OF THESIS submitted by Mala Alhaji Sainna for the degree of Doctor of

Philosophy (PhD) and entitled “Electronic and catalytic properties of iron porphyrin

complexes: Trends and reaction mechanisms”

The cytochrome P450s belong to the superfamily of proteins containing a heme cofactor and,

thus, are termed hemoproteins. They perform important oxidation reactions in the body, and

are, for instance, involved in the metabolism of drug molecules in the liver as well as the

detoxification of xenobiotics and biosynthesis of hormones. The active species of these

enzymes is identified as iron(IV)-oxo heme cation radical species (also known as Compound

I), but it is short-lived and difficult to trap and characterize experimentally. Therefore,

theoretical modelling as implemented in this thesis was instituted in order to give important

answers to questions such as the mechanisms of substrate monoxygenation, the oxidant

activity in enzymes and the effect of protein architecture on chemical catalysis. The present

thesis focuses on addressing these issues using a combination of density functional theory

and quantum mechanics/molecular mechanics studies. The work gives insight into the nature

of heme, axial ligand bound to heme and the impact of substrate on oxidation reactions. We

find trends in reaction mechanisms and rate constants and rationalize rate constants and

reaction processes. The work has also given insight into the nature of high valent iron(IV)-

oxo heme cation radical oxidants and their reactivity patterns with respect to a broad range of

substrates. Hence, the offered studies have shown how small structural differences in the

active site will result in dramatic differences in reactivity patterns and how nature approach

and catalyzes vital reaction mechanisms.

Page 16: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

16

DECLARATION

No portion of this work referred to within the thesis has been submitted in support of an

application for another degree or qualification of this or any other university or institute of

learning.

Page 17: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

17

COPY RIGHT STATEMENT

i. The author of this dissertation including any appendices and/or schedules to this dissertation

owns certain copyright or related rights in it and s/he has given the University of Manchester

certain rights to use such copyright, including for administrative purposes.

ii. Copies of this dissertation, either in full or in extracts and whether in hard or electronic copy,

may be made only in accordance with the copyright, Designs and Patents Act 1998 (as

amended) and regulations issued under it or, where appropriate, in accordance with licensing

agreements which the university has entered into. This page must form part of any such

copies made.

iii. The ownership of certain copyright, patents, designs, trade marks and other intellectual

property (the “Intellectual Property”) and any reproductions of copyright works in the

dissertation, for example graphs and tables (“Reproductions”), which may be described in

this dissertation, may not be owned by the author and may be owned by third parties. Such

intellectual properties and reproductions cannot and must not be made available for use

without the prior written permission of the owner(s) of the relevant intellectual property

and/or reproductions.

iv. Further information on the conditions under which disclosure, publication and

commercialisation of this dissertation, the copyright and any intellectual property and/or

reproductions described in it may take place is available in the University IP Policy (see

http://documents.manchester.ac.uk/display.aspx?DocID=487), in nay relevant Dissertation

restriction declarations deposited in the University Library, The University Library

regulations (see http://www.manchester.ac.uk/library/aboutus/regulations) and in the

University’s Guidance for the Presentation of Dissertations.

Page 18: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

18

CHAPTER 1

INTRODUCTION

Page 19: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

19

Introduction

1.1 Preface

Catalysis can be classified into homogeneous catalysis in which the reactant and the catalyst

exist in the same phase, and a heterogeneous catalysis in which the reaction occurs on the

interface between different phases; for instance the substrate is in solution and reacts with a

solid catalyst surface. Nowadays transition metal catalysts are the most commonly applied

examples of both homogeneous and heterogeneous catalysts (Busca, G.,2014). Undoubtedly,

although some of the inner transition metals, such as Pd, Ru, Ir and Rh may be efficient

catalysts for chemical reactions, it may not always be practical in an industrial setting to

utilise those catalysts due to difficulties to obtain and purification. Thus, their toxicity and

reactivity in nature leads to environmental and economic problems associated with their

usage. Consequently, industry is always seeking to find more environmentally friendly

alternatives for process technologies (Reetz, M.T.,2014). Therefore, current drives in

chemical research is replacement of toxic and expensive second and third row catalysts with

environmentally benign and cheaper options, and thus work has focus on developing catalysts

that use iron instead (Enthaler, S.,2013).

Transition metal catalysis has revolutionised organic chemistry enabling chemical

transformation of molecules that is very difficult to achieve otherwise (Thomas, J.M.,2014).

Some of the uniqueness of transition metals is their ability to offer varieties of oxidation

states and coordination spheres, which makes them highly susceptible to electron abstraction

and/or donation processes. The nature by which these processes act can have a strong

influence on, the regioselectivity as well as the stereospecificity of chemical reaction,etc

Page 20: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

20

(Kratsch, J., Roesky, P.W.;2014). Moreover, the ability of transition metals to bind multiple

atoms of a single ligand (hapticity) leads to interesting binding patterns in organometallic

complexes (de Visser, S.P.,2001, Shelby, M.L., Mara, M.W. et al 2014). In nature, these

transition metal catalysts form the class of enzymes called metalloenzymes, which are highly

multipurpose and take part in biological processes that include, for instance, metabolism,

biosynthesis and biodegradation processes. Often, these metalloenzymes utilise iron in their

active centre bound to the protein via a multidentate ligand system containing a heme or a

non-heme structure. One class of the heme metalloenzymes that are extensively studied are

the cytochrome P450s, which are monoxygenases, For example, The P450s are involved in

drug metabolism reactions in the liver (Gonzalez-Diaz, H., 2013, de Visser, S.P. 2013). These

catalysts have an extensive array of reactivity patterns; however, all proceed via short-lived

intermediates, which are difficult to study experimentally. With the advance in computational

methods, we are now able to investigate the behaviour and the electronic properties of these

species in detail, and are also able to predict the spectroscopic parameters of short-lived

intermediates (Shi, S.; Klotz, U. 2012, Murphy, C., 2014, Nyman, G., 2014, Tsipis, A.C.,

2014, Fioroni, M., Dworeck, T. et al 2014).

In this thesis, we discuss the reactivity patterns and chemical properties of catalytic cycle

intermediates of enzymes and synthetic model complexes. Our studies give insight into the

nature and the effect of ligands in the reaction mechanism, the rate constant and the product

distributions. Moreover, our work investigated the influence of the environment of the

reactants as well as their activity towards substrates.

Page 21: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

21

1.2 Homogeneous Catalysis

Homogeneous catalysis utilising transition metal complexes is one of the most studied

research areas and has received an enormous boost in recent years. Many interesting

discoveries were made over the years both by researchers and industries as well as in

academia. The homogeneous catalysis can be defined as a catalytic system in which the

reactant or the substrates for a reaction and the parent catalyst component are all merged

within a single phase, usually the liquid phase. Lately, a tapered definition has become viral

according to which homogeneous catalyst involves organometallic complexes as the

catalysts; although, it is a clear fact that an organometallic compound has a bond between the

central metal and a carbon atom which is not the case with all known homogeneous catalysts.

Therefore, an organometallic complexes can be categorised as a class of homogeneous

catalyst but there are few more interesting and important reactions in the literature that

employ homogeneous catalyst which are distinct from organometallic complexes based on

their ligand systems (Anderson, G. K.;1984, Mawby, R.J.; Basolo F.; et al 1964, Noack,

K.;Calderazzo, F. 1967). Examples of some of these catalysts include:

Diels-Alder reactions (Lewis acid as catalysts),

Ester hydrolysis which is a General acid and base catalysis,

Organic catalysts (thiazolium ion in Cannizzarro reactions),

Co-ordination complexes,

Enzymatic processes,

Porphyrin complexes,

Biomimetic complexes

Page 22: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

22

Although the above listed catalysts are categorised as homogeneous catalysts, they primarily

differ by their ligands effect and coordination. A single metal can give a variety of products

from the same substrate by simply altering its ligand environment. One good example found

in literature relates to the varieties of substrates formed by a single metal by simply changing

the ligands around the metal centre as indicated in Figure 1.1. The figure shows the various

products obtained from 1, 3-butadiene using a nickel catalyst with different ligand

coordination.

Figure 1.1: Effect of ligands and valence states on the selectivity in a metal centred catalysed

reaction of butadiene.

We will not discuss every class of homogeneous catalyst in detail in this chapter; however we

are going to give a brief description of the few we can relate to within our work.

Page 23: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

23

1.2.1 Ligand effects.

Homogeneous catalysis usage represents an important and resourceful method for carrying

out chemical transformations. There are many processes in industry that utilise organo-

transition metal catalysts (Flood, T.C.; Jensen, J.E.; et al 1981, Kataoka, Y.; Shibahara, A.; et

al 2001, van Leeuwen, P. W.N.M; Roobeek, C.F. et al 1994). The development of selective

chemical processes generally requires modifying ligands on the transition metal centre. These

ligands have a unique ability to stabilise metals in several oxidation states and geometries,

and they can also be tuned to radically change the reactivity of a catalyst. Hence, ligand

systems are one of the several tools utilised by chemists in studying and subsequently

controlling the reactivity of homogeneous catalysts. Apart from the ligand-based approach in

the control of transition metal reactivity, there are alternative approaches, including those

focused on altering the d-electron energy of the metal through either metal based oxidation or

reduction. This procedure can help in controlling the electronic and steric properties of the

metal coordination sphere, although the ligand-based approach is more widely used. These

procedures can also be used in controlling the thermodynamic binding affinities of transition

metals, for instance, it has been shown that reactivity pattern can be changed of a centrally

bounded transition metal as a function of ligand oxidation state as a result of redox-active

ligands. In particular, these redox-active ligands enable the stability of different oxidation

states and, thereby, resulted in transition metals that produced compounds with switchable

electronic state of reactivity (van Leeuwen, P. W. N. M.; Morokuma, K.; et al 1995, Maseras,

F.; Lledós, A. 2002, Dekker, G. P. C. M.; Buijs, A.; et al 1992). Moreover, it was found that

the catalyst produced from these kinds of ligands could be a sluggish catalyst in one

particular oxidation state and brilliantly active after single or multiple oxidation or

Page 24: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

24

reduction.Furthermore, they were also found to display selectivity for a definite

transformation depending on the oxidation state associated with the complex (Lundquist, E.

G.; Folting, K.; et al 1990). The redox-active ligands can also be used as an electron shuttle

in a catalytic process as shown by Hembre and McQueen (García Alonso, F. J.; Llamazares,

A.; et al 1991).

1.2.2 Bite angle effect

The term bite angle is usually associated to geometric parameters used to classify chelating

ligands in coordination chemistry; in particular, structures involved in organometallic

chemistry and some biomimetic and enzymatic complexes. Although it was commonly

applied to phosphine ligands due to its ability to adopt a wide range of chelating ring sizes,

the parameters are also applicable to any sort of chelating ligands. We are however going to

use the widely understood diphosphines as an example to give a brief explanation of these

features. Many instances indicate that the ligand bite angle is somehow associated to catalytic

performance in a number of reactions. Early examples are the platinum diphosphines

catalysed hydroformulations (Versluis, L.; Ziegler, T.; et al 1990). In recent years, a

connection between the ligand bite angles and the selectivity of a catalyst has been observed;

a specific example is the rhodium catalysed hydroformylation, the Diels-Alder reactions as

well as the nickel catalysed hydrocyanation (Brookhart, M.; Green, M. L. H. 1983). The L-

M-L angle found in transition metal complexes is believed to be a concession between the

attached ligands preferred bite angle and the angle preferred by the metal centre. The former

is mainly determined by constraints forced by the ligand atoms and the backbone. Electronic

effects seem to have a more indirect influence by changing the preferred metal-ligand bond

Page 25: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

25

length. The metal preferred bite angle however, is primarily determined by electronic

requirements; for instance the nature and number of d-orbitals associated in the formation of

molecular orbitals. Other ligands attached to the metal centre can also have influence on the

metal orbitals, for example, -bonding ligands.

As per the aim in explaining the effect of bite angles in the catalytic reactions, there are

distinctions worth mentioning between two different effects (Mole, L.; Spencer, J. L.; et al

1991, Conroy-Lewis, F. M.; Mole, L.; et al 1991): (1) steric bite effect: This effect is related

to the steric interactions between ligand-ligand or ligand-substrate, which is observed when

the bite angle is modified by altering the backbone and keeping the attached substituents at

the donor atoms the same. The steric interaction observed from this has an influence upon

altering the energies of the transition states and the catalyst resting states, thus modifying the

activity or selectivity of the main catalytic system. (2) electronic bite angle effect: As the

name implies this is associated to the electronic changes within the catalytic centre as the bite

angle is altered (Thorn D. L.; Hoffmann, R. 1978). It can be defined as an orbital effect, due

to the fact that the bite angle determines metal hybridisation and as a consequence metal

orbital energies and reactivity. This effect can also show itself as a stabilisation or as a

destabilisation at the initial reaction, final or at the transition state of a reaction. It was also

observed that when the substituents at the donor atom are kept the same while the bite angle

is changed, the steric properties also change (Hickey, C. E.; Maitlis, P. M. 1984, Amatore, C.;

Jutand, A. 2000).

Page 26: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

26

1.2.3 Cone angle effect

The cone angle notion was found to be of practical importance when dealing with

homogeneous catalysis due to the fact that the size of the ligand could affect the reactivity of

the attached metal centre. A known example shows where the selectivity of hydroformylation

catalysts was strongly influenced by the size of the co-ligands (Calhorda, M. J.; Brown J. M.;

et al 1991, Widenhoefer, R. A.; Buchwald, S. L.; 1998); Thus, some phosphine ligands are

wide enough to conquer more than they should within the coordination sphere of a metal

centre. The concept of cone angle was first introduced by Tolmans parameter Ѳ (theta) in an

attempt to define a reliable steric parameter complementary to the electronic parameter.

Tolman proposed to measure the steric bulk of phosphine ligand from CPK models (CPK

models are also called Space-filling models after the chemists Robert Corey, Linus

Pauling and Walter Koltun, who founded their use) (Robert B.; Pauling, L. 1953). They

constructed a cone which embraces all the atoms of the substituents on the phosphorus atom

at a distance of 2.28Å from the metal centre and the cone angle is then measured which are

considered as the desired steric parameters. Tolman proposed quite a few other methods

which were later modified according to the chemical series investigated. Sterically, it is

understood that more bulky ligands give less stability in complexes. In a series of comparable

ligands, such as aryl phosphites, this leads to a reliable index. Crystal structure determination

has indicated that practically the angles recognized within the structure complexes are smaller

than the Ѳ-vales would propose. For instance, in the case of cis-triphenylphoshine molecules,

a double of the cis-triphenylphoshine may give an angle of 95o between P-M-P, whereas the

Ѳ-values would actually predict 145o in reality interlinking of the R-substituents leads to

smaller effective cone angles. In some instances, the steric interactions nearby the metal

Page 27: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

27

centre could be important while for other properties interacting in a more distant from the

metal centre may dominate, but the ligands never form a perfect cone angle (Fitton, P.; Rick,

E. A. 1971, Goertz, W.; Kamer, P. C. J.; et al 2001).

Recently, numerous efforts have started in order to describe the steric properties of ligands

using molecular mechanics and analysis of data taken from X-ray studies as well as data

mining (Marcone, J. E.; Moloy, K. G. 1998). The accessible molecular surface (AMS)

method is used. In the AMS method, the effective contours of the ligands are calculated,

resembling that of the calculation of accessible surface for enzymes. Crystallographic data

have also been utilised to calculate and compare steric ligand properties, more in particular

for bidentate ligands (Brown , J. M.; Cooley, N. A. 1988, Schrock, R. R. 1979).

1.3 Ligand according to donor atoms.

1.3.1 Imido and Alkoxy groups as an anionic ligand.

Alkoxides and imido groups are known anionic ligands used in titanium and zirconium

catalysts for alkenes polymerisations; they are either used in the catalysts as the only anions

or in combination with cyclopentadienyl ligands (Bäckvall, J. E.; Åkermark B.; et al 1979).

Alkoxide ligands are also used in titanium catalysts for epoxidation of alkenes. The

alkoxides when connected to a neutral ligand like an imine or a phosphine are known to be

excellent ligands for divalent metal catalysts of nickel and palladium for alkene reactions, in

which the second valency of the complex is made up by a hydrocarbyl group involved in a

chain growth reaction or a hydride during the chain transfer process. The alkoxides and

amides are involved in hydrogen transfer catalysts where they play an active part in terms of

abstracting a proton from the alcohol substrate (Luo, X.-L.; Crabtree, R. H. 1989); whereas

Page 28: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

28

the elements of column 1-3 on the periodic table such as oxygen and nitrogen are much more

stable when bound as an anionic ligand to a metal centre. In the case of transition metal

complexes the differences between oxygen or nitrogen-based anions is much smaller, even

more so as you proceed towards the right hand side of the periodic table (Luo, X.-L.;

Crabtree, R. H. 1989). As a consequence a large number of Metal-to-heteroatom bonds

contribute in catalytic reactions leading to carbon-to-heteroatom bonds and it allows us to

carry out a large part of our organometallic chemistry in the presence of alcohol, amines or

water etc.

1.3.2 Neutral and Anionic hydrocarbyl groups

Aromatics such as the cyclopentadienyl appear as ligands and became extremely important in

catalysis for transition metals such as Ti, Zr, Hf and Ru for many decades. They also occur as

ligands in ruthenium complexes that are used in hydrogen transfer reaction where two

hydrogen atoms are transferred from donor molecules such as alcohol, to a ketone yielding

another alcohol as the product. Emphasis was made on the enantiospecific variant (Vaska, L.

Diluzio, J. W. 1962, Vaska, L. 1968). For example, when a weakly coordinating anionic

ligand is applied to an early-transition metal that is cationic, these metal ions can be solvated

by aromatics if no stronger ligands are present. Aromatics ligands have as well been used in

nickel complexes and shown to make the catalysts highly active in the addition or vinyl-type

polymerisation of norbornenes (James, B. R.; Mahajan, D. et al 1979).

1.3.3 Oxazolines, imines , Amines and related ligands

All the above listed compounds contain a donating nitrogen atom. A nitrogen ligand is known

to be having the strongest donating function and has been used for many years in catalysis.

Page 29: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

29

They occur naturally in imidazoles and porphyrins and bind to transition metals easily, and

they are known to be involved in oxidation reactions (Ungváry, F. 1972). Several of these

complexes were mimicked and are used in the field of homogeneous catalysis to understand

reactions such as, the oxidation of C-H bonds or phenol oxidative coupling reactions (Major,

A.; Horváth, I. T.; et al 1988).

The amines and pyridines are among the oldest most studied ligands in the field of

coordination chemistry and catalysis. For instance pyridines were used in hydrogenation

catalysts based on copper (I) and (II) long before phosphines in platinum complexes

(Mirbach, M. F. 1984), and the sp3 hybridised nitrogen in amines are considered an analogue

of phosphines; they are known to be strong -donor capability as compared to phosphines,

and they are excellent in stabilising high-valent metal complexes such as tetravalent

palladium and platinum (Chinn, M. S.; Heinekey, D. M. 1990). Nitrogen ligands are

generally much more stable than phosphines especially in pyridine and imidazole ligand

complexes. Pyridine complexes, for instance, are not susceptible to oxidation as phosphines,

they are also not known to follow the decomposition pathways of phosphines such as C-P

bond cleavage, phosphide formation or the phosphide hydrolysis (Van Leeuwen, P. W. N.

M.; Roobeek, C. F.; 1986), hence nitrogen ligands are by far more preferable than phosphine

in oxidation catalysis. The bonding characteristics of pyridines or any other sp2 hybridised

nitrogen to that of phosphorus donor ligands are quite distinct. Although they are both good

-donor and poor -acceptor ligands, they both stabilise higher oxidation states of metal

complexes instead of a low oxidation states. For instance in the catalysis of divalent metal or

nickel they have a very similar behaviour, however, in the case of reductive elimination

Page 30: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

30

reactions leading to zero-valent complexes, they have different characteristics and

performance, whereby phosphorus is more preferable in this instance.

The oxazolines are thought to have an additional advantage in comparison to other imine that

are nitrogen based ligands due to the fact that the asymmetric derivatives are readily

accessible. A very good example of an oxazoline is the group of PyBox ligands (Figure 1.1)

which is a typically chiral oxazolines (Inuki, T.; Kojima, T. 1967). The mono-oxazoline has

also found a wide recognition and application, the mono-oxazoline contain a phosphine group

as a second donor atom. The substitution of one of the oxygen atoms in oxazoline by a

nitrogen atom yield an imidazoline which contain the asymmetric carbon atom but offer an

electronic variation through substitution at the amine-nitrogen atom (Fleming, I. 1976). The

imidazoline substituent might as well be applicable for immobilisation via covalent linkages.

Figure 1.2: PyBox Ligand from an oxazoline group.

1.3.4 Carbon monoxide and Carbene

Carbenes act both as an intermediate and as a ligand in the field of catalysis. They are known

to occur as an intermediate in a reaction of alkene metathesis as well as alkene

cyclopropanation. They carry hydrogen and carbon substituent as intermediates, hence they

Page 31: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

31

are classified as a “Schrock carbenes”. As ligand, however, they often contain nitrogen

substituent and are therefore classified as a “Fischer carbenes”. They have received a great

attention within the last few years as ligands in metal complexes catalysis (Denmark, S. E.;

Fu, 2003, Landmana, M.; Pretorius, R.; et al 2014), however the structural motive was

already in existence and in exploration since the early seventies (Ueki, M.; Matsumoto, Y.; et

al 2001, Sainna, M. A., Singh, D. et al 2015).

Figure 1.3: Example of a typical carbene Ligand

The carbine as ligands, are strong -donor and rarely -acceptors, they are sometimes called

“singlet carbenes”.

The carbon monoxides are often used as either a reactant or a ligand. It is found to be a very

strong -acceptor as a ligand and a modest -donor. In comparison of its steric properties to

that of a phosphorus ligand, it is believed to be one of the smallest ligands available with a χ-

value of around 55 when characterised using IR. One of the exciting features of the CO

ligands is its ease of characterising using IR as mentioned earlier and also in situ IR, in a

transparent frequency area (1800-2200 cm-1

). Although its disadvantages include its demand

of high pressure of about 10 bar in order to obtain a concentration of order 0.1M in an

organic liquid.

Page 32: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

32

1.4 Common anions

In most cases the anions form part of the coordination complexes as well as the

organometallic complexes and their role should not be undervalued. In many cases they act as

reactants, for instance in cross-coupling reactions whereby a salt is made as a second product

and even in the simple series of halides its anion function toward the metal centre makes an

enormous difference in catalytic behaviour (Ishihara, K.; Kurihara, H.; et al 1998). The

function of halides and carboxylates as a ligand goes beyond doubt as they are known to bind

strongly to metal ions. For instance, a carboxylate bound to a phosphine is a moiety found as

an impressive ligand in the shell process for ethane oligomerisation (Reetz, M. T.; Kyung, S.

H.; et al 1986), whereas an iodide which is a halide family is found to be important ingredient

of methanol carbonylation chemistry as they are known to form part of the ligand

environment in iridium and rhodium metal complexes (Ishihara, K.; Ohara, S.; et al 2000).

The introduction and implementation of weakly and non-coordinating anions (WCA, NCA)

has also received a great attention over the years as they are indispensable in homogeneous

catalysis. The cationic counter-ions obtained as a result of using a WCA is willingly

accessible for the reactants especially reactions involving alkenes such as polymerisations,

carbonylations and hydrogenations reactions. The introduction of weakly coordinating anions

has led to numerous advances. In the early history of homogeneous catalysis the role of

“Ziegler catalysts” where alkene coordination is facilitated by the role of creating vacant sites

at a metal ion was played by Lewis acids abstracting chloride ions from the catalyst precursor

(Garrou, P. E. 1985), but this has changed in the recent years as the WCA has also headed to

Page 33: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

33

an additional control over the species formed in the reactor compared to the use of Lewis

acids.

1.5 Vacant site Creation and coordination of the substrate

The sole purpose of a catalyst is to bring the reactant together and alter the activation barrier

of the reaction; however bringing the reactants together requires a metal centre with a vacant

site. Metal catalysis starts with the creation of a vacant site we could say. For a homogeneous

catalyst in a condensed phase creating a vacant site can be challenging as a condensed phase

solvent molecules will always be co-ordinated to the active metal ion and therefore “vacant

site” is an inaccurate description and this is because substrates are present in excess and so

are the ligands. Hence, a competition in complex formation exists between the desired

susbstrate and other potential ligands present within the same phase or the solution. Usually a

negative order in one of the ligands concentrations can be found in the rate of product

formations expression; a zero order in the concentration of the substrate is obtained when the

substrate co-ordinates strongly to the metal centre. i.e. saturation kinetics (Billig, E.;

Jamerson, J. D.; et al 1980), also known as c.f. Michaelis-Menten kinetics (Goel, A. B. 1984).

Another consideration needs to be taken into account when creating a vacant site and co-

ordination of substrate is the classical way by which substitution reactions are described

(Sakakura, T. 1984). Two thrilling mechanisms are distinguished, an associative mechanisms

and a dissociative mechanism. In the case of dissociative mechanism the breaking of the

bond between the metal and the leaving ligand is the rate controlling step, a solvent molecule

thereby occupies the open site which is a phenomenon that does not appear in the rate

equation. Subsequently the replacement of the solvent by the substrate occurs in the first step.

Page 34: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

34

In the associative process (SN2) however, the displacement is the bimolecular process with

simultaneous bond breaking of the ligand and the formation of the metal with the substrate

(Abatjoglou, A. G.; Billig, E.; et al 1984); In square planar complexes as found for the group

9 and 10 metals, the associative process is most common.

Figure 1.4: Dissociative and associative ligand exchange

1.6 Insertion versus migration

The terms insertion and migration denotes the process in which an unsaturated molecule

inserts to a metal-anion bond. In Figure 1.4 and 1.5, a typical instance of the aforementioned

reaction showing acetyl fragment as formed from a co-ordinated CO and a methyl group on a

platinum complex, the two reacting groups mentioned must occupy positions cis to one

another otherwise the reaction does not occur (Abatjoglou, A. G.; Bryant, D. R. 1984).

Page 35: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

35

Figure 1.5: insertion mechanism

The most important difference between the insertion mechanism and the migration

mechanism is that, the insertion mechanism consist of carbon monoxide inserting into the

metal-methyl bond while the acetyl bond formed takes the place of the methyl group; that

means the -bonded fragment retains its position trans to *, whereas in the migration

mechanism the methyl group moves to the co-ordinated carbon monoxide resulting to an

acetyl group occupying the position cis to *. The mechanisms depicted in Figure 1.5 above

is evidenced by NMR spectroscopic studies on a platinum complex of diphosphine with the

platinum containing slightly distinct phosphine groups such that they can be distinguished in

the NMR spectrum of both the acetyl and alkyl species (Kong, K-C.; Cheng, C-H. 1991). The

example of migration mechanism with rhodium complex above has proven the fact that

migration mechanism is the path way for complexes containing “piano-stool” structures;

theoretical calculation also ascertained the migration mechanism (Kikukawa, K.; Takagi, M.;

et al 1979), showing the anionic methyl group shifting to the positively charged carbon atom.

Therefore the migration is considered more accurate for this process. Although there are

several experiments that suggest the insertion is the actual mechanism. Most of the systems

studied consist of Mn and Fe as the metal involved. A typical example on a platinum

complex is depicted in Figure 1.6 below. The most stable starting geometry indicated that the

methyl group and the phosphine are in cis positions, because of the influence of the trans. The

Page 36: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

36

same mechanism happens for the acetyl product. This explanation might be considered as a

proof for an insertion reaction, however, as the figure shows the migration might be followed

by an isomerisation or an isomerisation may take place after the migration. This is reinforced

by the fact that asymmetric ligands usually undergo insertion reaction pathway more slowly

than the symmetric ligands. Thus, diphosphine and bipyridine complexes undergo faster

reactions as compared to the mixed phosphine-nitrogen ligands (Sisak, A.; Ungváry, F.; et al

1983). Generally, the experiments discussed do not disprove the cis-migration mechanisms

though, although it presumably resulted from topomerisations of the intermediates or

products (Bouaoud, S-E.; Braunstein, P.; et al 1986).

Figure 1.6: migration mechanism

According to the depicted insertion reaction mechanisms shown in the Figure 1.4 above, we

assumed the reacting carbon monoxide is co-ordinated to the Pt metal. There is a clear

Page 37: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

37

evidence proven experimentally that methyl migration is indeed co-ordinated to carbon

monoxide (Szulc, A.; Meyerstein, D.; et al 1998). The standard evidence comes from a

relatively inert complex where both the migration as well as the exchange of the co-ordinated

CO with free CO is slow. A clear example is The reaction of CH3Mn(CO)5 as shown in the

Figure 1.7 below, result in the formation of CH3(CO)Mn(CO)4(13

CO) when reacted with 13

C

labelling free CO in which the CO is present as co-ordinating carbon monoxide instead of the

acetyl group (Alcock, N.W.; Bergamini, P.; et al 1987). Therefore, there does not seem to be

a direct evidence of reaction between the methyl manganese unit and the freshly received

carbon monoxide. There is no established proof an insertion of an uncomplexed unsaturated

susbstrate into a metal-to-carbon -bond. It is worth mentioning as in when dealing with

heterogeneous catalysis, however, coordination of the susbtrate to an active metal surface is

acknowledged as the “Langmuir-Hinshelwood” mechanism, whereas the reaction of a

molecule in a gas-phase with some of its fragments on the surface is named the “Eley-Rideal”

mechanism.

Figure 1.7: disproval of “outersphere” insertion

Another essential type of the migration reaction mechanism is the one consisting of alkenes

in place of carbon monoxide as discussed earlier. In this type of migration reaction a hydride

migrates to a co-ordinated ethene molecule cis to the hydride which may result in an empty

space in the coordination sphere of the metal. There are two ways of lifting these sort of co-

Page 38: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

38

ordinative unsaturations: (1) either an agostic interaction of the unsaturated metal with the -

hydrogen may occur which is a mechanism supported by both experimentalists and the

theoretician (van Leeuwen, P. W. N. M.; Roobeek, C. F.; et al 1990); or, (2) the occupation

of the vacant site by an incoming ligand.

The activation of a co-ordinated alkenes taking place before migration is currently unclear

although the coordinated alkenes are subject to -back donation and-donation but the

overall result of the electron density is unpredictable. Although, molecular orbital

calculations at an extended level using Hückel theory (Goodson, F. E.; Wallow, T. I.; et al

1997) indicated that in many instances the co-ordinated alkenes are not activated towards

nucleophilic attack (migration) and cannot a priori be that the hydride will undergo a rapid

migration. A strong back-donation leads to more electron-rich alkenes which reduce their

vulnerability for the attack of the migrating group. A strong polarisation of the alkene will

occur when an asymmetric bonding of the alkene is invoked as shown in Figure 1.8 below.

The depicted structure representative below, the alkene is shown to be activated towards the

hydride migration (Bianchini, C.; Meli, A. 1998, Chen, J.B.; Angelici, R. 2000).

Figure 1.8: charge distribution in a migration reaction

Page 39: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

39

1.6.1 β-Elimination and de-insertion:

The β-elimination literally refers to the reverse of the migration of η-bonded anionic groups

to co-ordinated alkenes. The migration reaction (or nucleophilic reaction) reduces the total

number of electrons of the complex by two, thereby creating a vacant site at the metal,

whereas the -elimination behaves the exact opposite as it requires a vacant site at the

complex (abandoning coordinated solvents in the complex), and during the reaction process

the electron count is increased by two instead. Reactions resembling the -elimination occur

in many organic reactions although the variation is in the intramolecular nature of the present

process as the eliminated alkene may be retained within the complex. In organic chemistry, a

two-step process is observed in the reaction mechanism, for instance proton elimination with

a base followed by the leaving of the anion. In transition metal chemistry, however, the

availability of d-orbitals facilitates a concerted cis -elimination (Dong, L.; Duckett, S. B.; et

al 1992).

Figure 1.9: β – hydride elimination

The -elimination is a reaction that is often required to be suppressed in order to achieve a

desirable feature. There are a few ways to achieve the suppression of the -elimination:

Page 40: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

40

1. By maintaining a coordinative saturation of the complex although this might be a

counterproductive idea as the next step in the catalytic cycle of the complex will

surely require a vacant site to react.

2. By the selection of metals that are stable with metal alkyl complexes with esteem to

hydride and liberated alkene. For the metals on the left side of the periodic table such

as the early transition metals and the lanthanides, there is a relative stability with the

alkyl, hence the best alkene polymerisation catalysts are obtained among these

aforementioned metals.

3. Steric hindrance could be instigated in order to avoid the correct stereochemistry

required for -elimination leading to stability of the metal alkyl complex. In the

modern polymerisation catalysts for polypropene this feature is observed, thereby

leading to polymers with higher molecular weight.

The de-insertion is also used instead of the -insertion especially when dealing with CO. The

process is absolutely similar because it also requires a vacant site for the reaction to occur,

and the electron count of the metal rises by two in the process of de-insertion. In the de-

insertion, an insertion takes place between a -bonded and -bonded fragment in mutual cis-

positions, as was discussed earlier. The mechanisms of de-insertion only proceed if there is a

vacant site cis to the acyl group. A typical example of the de-insertion is shown in the Figure

1.10 below as provided by the experimentalist (Kaneda, K.; Sano, K.; et al 1979).

Figure 1.10: Migratory de - insertion

Page 41: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

41

1.6.2 Oxidative addition and Reductive elimination:

In an oxidative addition reaction, a compound AB is bound to a metal complex M during

which the bond of the bounded compound AB is broken to form a new bond with the metal

MA, and MB.A while B is reduced, and both will obtain a minus one charge thereby raising

the oxidation state of the metal by two, also the co-ordination number of the metal also

increases by two, while the electron around the metal complex raises by two and the d-

electron count of the metal decreases by two electrons. In oxidative addition, electronic

ligand effects are exceedingly likely. Donors strongly stimulate the formation of high-

valence states and therefore give rise to oxidative additions.

Reductive elimination is considered the reverse reaction of the oxidative addition discussed.

In the reductive elimination the formal valence state of the metal is reduced by either two

electron or one for a bimetallic reaction thereby reducing the total electron count of the

complex by two. Stabilisation of the low-valent state of the product promotes reductive

elimination; this means a good -acceptor ligands, bulky ligands as well as ligands preferring

bite angles more suited for tetrahedral than for square-planar complexes, when dealing with

elements of group 10 metals.

1.7 Activation of a substrate toward nucleophilic attack

1.7.1 Role of Alkenes and alkynes

Bounding of an alkene to an electronegative metal or often a positively charged metal tend to

activate the alkene toward a nuclophilic attack. Once the nucleophilic attack occurs, the

alkene complex will convert into a -bonded alkyl complex with the nucleophile at the β-

Page 42: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

42

position (Chan, A. S. C.; Caroll, W. E.; et al 1983). The overall product obtained from the

nucleophilic attack is virtually the same as that of insertion reaction mechanism; the two are

distinct by the fact that the insertion reaction gives rise to a syn-addition while the

nucleophilic attack to an anti-addition. Sometimes the two reaction mechanisms are referred

to as inner and outer sphere attack for the insertion and nucleophilic attack reactions

respectively. Properly substituted alkenes could demonstrate either the syn or anti manner of

the addition reaction. These types of addition reaction is known as a key-step in palladium

catalysed reactions in “Wacker-type” process (van Leeuwen, P. W. N. M. 2001, Chen, J.;

Daniels, L. M.; et al 1990, Angelici, R. 1988).

Figure 1.11: Nucleophilic attack to a coordinated alkene

Although, the activation of the alkene by the metal toward nucleophilic attack is unclear as

the counteracting influences of donation and back-donation is under consideration.

The same reaction is observed with the alkynes as discussed for the alkenes above with a

product having an anti-isomer. An alkene is obtained as a product of the anti-addition after an

appropriate elimination of the metal. In the previous paragraph it was mentioned that

insertion into a metal hydride bond followed by hydrogenation will lead to the syn-product.

Page 43: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

43

At this stage, it is assumed that the main characteristic and the properties that determine the

features of a catalyst are fairly recapped within the above paragraphs. We are going to

quickly discuss a metalloenzyme and biomimetic complexes which the core work mostly

depend upon as well as organometallic complexes, although we will put more emphases on

the metalloenzymes.

The organometallic catalysts usually consist of a metal centre surrounded by organic and

inorganic ligands. The properties and characteristics of an organometallic catalyst are usually

determined by both the central metal and the ligands coordination as the above discussed

characteristics clearly indicated.

1.8 Metalloenzymes

Although nature chose first row transition metals to develop metalloenzymes for biological

catalysis, second and third row transition metals are also utilised in the development of

biological systems. Metalloenzymes are protein which contains a metal cofactor that is

directly linked to the protein or to an enzyme-bound nonprotein system (Holm, R.,

Kennepohl, P., et al 1996). The utilisation of nature towards the first row transition elements

is due to the fact that the first row transition metals ions are more abundant in soluble form in

the evolution of living systems. In other terms, the metalloenzyme development is not only

predicted by optimum functionality but as well by other evolutionary pressures; for instance

by the availability of metals within the cell (Valdez C.E, Smith Q.A, et al 2014). On the other

hand, a broader collection of metals can be utilised in metalloenzyme design where the main

aim is obtaining a maximum catalytic efficiency. It is therefore essential to have a good

understanding on the binding of non-biological metals into existing metal-binding proteins

Page 44: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

44

and its effect on enzymatic systems. Most proteins consist of metal-binding sites that contain

either amino acid side-chains or a subordinate ligand; these metal-binding sites can

accommodate non-native metals with similar coordination geometries and electronic

properties. To enhance the binding activity of the protein, the side-chain can be improved by

protein engineering while the subordinate ligands can be chemically modified. In addition to

the use of existing metal-binding scaffolds, non-native binding sites can be designed and

inserted back into the protein to create artificial metalloenzymes , as discussed in some

articles in literature (Lu Y, Yeung N, et al 2009, Petrik I.D.; Liu J, et al 2014, Yu F,

Cangelosi V.M, et al 2014, Lewis J.C, 2013).

Figure 1.12: Substitution of metal in cofactor to incorporate alternative metal into a binding

site. The Apo-protein must be synthesised first before introducing the new metal. Picture is

obtained from protein data bank 1N2C (Schindelin, H.; Kisker, C. et al 1997).

Although within these briefly discussed metalloenzymes there are a highly prominent

oxidants known as the cytochromes P450s; these are group of well-studied heme enzymes

that are known to react via oxygen atom transfer to substrates. The name P450 originated

from the fact that the strong electron donating group of the cysteinate axial ligand of the

enzyme in the carbon monoxide bound ferrous heme form gives a red-shifted robust Soret

Page 45: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

45

band through a high-energy -* conversion of the porphyrin ring at 450 nm, hence the name

cytochrome P450. The P450s utilise a high valent iron(IV)-oxo heme cation radical species as

its active oxidant to catalyse a range of chemical reactions, such as: 1) aromatic

hydroxylation, 2) double bond epoxidation leading to an epoxide, 3) sulphoxidation reaction

mechanism and 4) aliphatic hydroxylation of a saturated compound to give either an alcohol

or an unsaturated species. The heme enzymes are subdivided into a monoxygenases and

dioxygenases (de Visser, S.P., 2009). A monoxygenase binds an oxygen molecule (O2) into

the active site of the enzyme, where it is utilised on a transition metal centre to oxidize a

substrate. In the process, a water molecule is released as well. A dioxygenase, by contrast

uses both oxygen atoms of O2, which are donated to substrate(s). These oxygen binding

enzymes are naturally classified according to their ligand system as heme or a non-heme

metal-loenzymes (Conroy-Lewis, F. M.; Mole, L.; et al 1991, Shaik, S., de Visser, S.P., et al

2004). Heme enzymes are among the most thoroughly studied enzymes in the past 50 years.

The work has gained thorough understanding of its structure, mechanisms, biological

function and its diversified isozymes. It has long been known that the P450s catalyse the

reductive activation as well as molecular oxygen scission that binds to their heme metal iron

thereby leading to the insertion of an atom of oxygen into a substrate concomitant to the

production of a water molecule from the other oxygen atom (Pan, Z., Zhang, R., et al 2006).

The reaction discussed above requires two electrons and two proton transfers to the heme

metal. Equation 1 below shows the typical P450 reaction overview of aliphatic

hydroxylation.

RH + O2 + 2e- + 2H+ R – OH + H2O (1)

Page 46: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

46

As mentioned earlier, the cytochrome P450s catalyse a diverse range of chemical reactions

other than hydroxylation as represented in Equation 1 above; because of this there are many

P450 isozymes with different structure and folding patterns which means that all the

reactivity patterns catalysed have different proton and electron relay channels. Nevertheless,

all P450 isozymes undergo the same catalytic cycle starting from the resting state (A) as

depicted in Figure 1.13

Figure 1.13: Schematic representation of the catalytic cycle of a typical substrate

hydroxylation reaction of substrate R–H by P450 enzymes and a visual depiction of the active

species (Highlighted in a rectangular box) (Nam, W., et al., 2000).

N N

N NFe3+

O

O

S

O

ON N

N NFe3+

O

O

S

O

O

N N

N NFe2+

O

O

S

O

O

N N

N NFe2+

O

O

S

O

O

N N

N NFe3+

O

O

S

O

O

N N

N NFe3+

O

O

S

O

O

N N

N NFe4+

O

O

S

O

O

N N

N NFe3+

O

O

S

O

O

Page 47: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

47

In the first step of the cycle, the displacement of the distal water ligand due to substrate

binding occurs to form (B). This leads to a shift in ferric heme iron spin from a low-spin (S =

½) to a high-spin (S = 5/2). The spin-change triggers an electron transfer from the redox

partner to reduce the heme iron to a ferrous state to form the iron(II) complex (C). The

ferrous heme binds oxygen to form the ferrous-oxy intermediate (D) and its reduction gives

the peroxy state (E) by the delivery of a second electron from the redox partner. Protonation

produces the ferric hydroperoxo complex known as Compound 0 (F). An extra protonation

leading to a scission of the dioxygen bound as well as the production of a water molecule is

made to form Compound I (G). The intermediate formed on the heme is a ferryl-oxo heme

cation radical species known as compound I. The compound I (structure G outlined in a

rectangular box in figure 1.12) are known to be highly reactive; the compound I and

compound 0 are also known to have a very short-lifetime and it took huge experimental

efforts to characterise and trap these species experimentally within the past few decades.

Compound I attacks a nearby substrate via hydrogen atom abstraction/radical rebound

process according to the cycle depicted above to form the alcohol product (R –OH). The

latter dissociates from the substrate binding pocket, which allows it to refill with water

molecules and a new substrate. The heme relaxes back to the resting state ferric iron thereby

completing the catalytic cycle. In some instances, the formation of non-productive pathways

leads to dead-end products; for instance the ferrous-oxy species can decay to reform ferric

P450 with production of superoxide.

In this report we will discuss work performed on the mechanism of P450 and biomimetic

model complexes. The work has given insight into the factors that determine the rate constant

of oxygen atom transfer, in particular, substrate epoxidation and has led to a predictive model

Page 48: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

48

for substrate activation. Although we have not re-visited all the reactivity patterns of the

iron(IV)-oxo species in P450 enzymes with substrates, we have investigated a few which will

be briefly discussed and summarised within the chapter.

Figure 1.14: Reactivity patterns of P40 Compound I with selected substrates.

1.9 Epoxidation

P450 compound I is known to react via oxygen atom transfer to olefins, thereby generating

epoxides. For example, the formation of an epoxide is an initial step in the metabolism of

monosaturated fatty acids by P450 enzymes. Double bond epoxidation by P450 isozymes is

widely studied computationally (de Visser S.P, 2012, de Visser S.P., Kumar, D., et al 2004,

Shaik, S., de Visser, S.P., et al 2002). The mechanism of olefin epoxidation by P450 enzymes

Page 49: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

49

involves a step-wise reaction via a radical intermediate (Pan, Z., Zhang, R., et al 2006, Nam,

W., 2000). The reaction starts with an electrophilic attack of the oxo group on the double

bond through a rate-determining barrier (first transition state) that leads to a C – O bond

formation between the oxo group of the iron (IV)-oxo oxidant with the carbon of the

substrate. The above mechanism gives a radical intermediate. This step is followed by a ring

closure barrier (second transition state) to form epoxide products. It is a known fact that the

ring closure barrier is negligible on the double spin state, which means the reaction is pseudo-

concerted (de Visser, S.P., 2012). The mechanism leading to the epoxidation reaction by a

substrate 2-butene is displayed in figure 1.14 using a potential energy profile to display the

energy barriers at each stage of the mechanism.

Figure 1.15: Potential energy profile for the epoxidation reaction of 2-butene by 4,2

[Fe(IV)-

oxo Porphyrin with a thiolate axial ligand as calculated with DFT methods. Data taken from

ref. (de Visser 2012).Values in Kcal/mol.

Page 50: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

50

Epoxidation reactions of terminal olefins sometimes lead to suicidal complexes (de Visser

S.P, 2012, de Visser S.P., Kumar, D., et al 2004, Shaik, S., de Visser, S.P., et al 2002), where

a covalent bond between the heme and the substrate is formed. This ultimately leads to loss

of the metal from the heme, and hence, deactivates the enzyme. Computational modelling

showed that dead-end complexes are formed when a state crossing occurs from the radical

intermediate to a cationic state (de Visser, S.P., Ogliaro, F., et al 2001)

1.10 Hydroxylation

The substrate hydroxylation reaction is a common reaction catalysed by P450 enzymes, both

with aliphatic and aromatic substrates; although the reaction mechanisms are quite distinct. In

aliphatic hydroxylation, the reaction starts with a hydrogen atom abstraction prior to a radical

rebound to form alcohol as a product; it also proceeds via a two state reactivity pattern on

competing doublet and quartet spin states. However, in the aromatic hydroxylation the

hydrogen atom abstraction from C-H is thermodynamically difficult and requires a

substantial amount of energy to break, hence the reaction path undergoes a mechanism

known as the electrophilic substitution mechanism (de Visser, S.P., 2012). The aromatic

hydroxylation reaction studies in the past (de Visser, S.P., 2012, Bojic, M., Sedgeman, C.A.,

et al 2015) have all confirmed to have given the same pathways.

The reaction typically starts via an electrophilic attack of the oxo group on one of the carbon

atoms of the attacking aromatic substrate to form a conjoined complex known as the

“Meisenheimer complex” with a configuration of either a cationic or a radical intermediate

(de Visser, S.P. 2012). Typically compound I has a distinctive electronic configuration

consisting of three unpaired electrons: of the three unpaired electrons two are located on

Page 51: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

51

orthogonal *FeO orbitals (*xz and *yz), whereas the third is located on the heme within an

orbital in D4h symmetrical orientation labelled a2u (Green, M. T., 1999, Ogliaro, F.; de Visser,

S. P.; 2001). The optimised geometry of compound I as depicted in figure 1.4. (de Visser, S.

P.; Shaik, S.; et al 2003, Ogliaro, F.; de Visser, S. P.; 2001), and the high-lying occupied as

well as the low-lying virtual molecular orbitals that are essential for the mechanism of

hydroxylation reaction. The a2u orbital mentioned earlier is found on the heme and it is

responsible in mixing with a lone-pair orbital on the axial ligand (Ogliaro, F.; Cohen, S.; et

al 2000).

Figure 1.16: High-lying occupied and low lying virtual orbitals of Compound I at an

optimised geometry of P450 Cpd I of quartet and doublet spin electron arrangements.

In the cationic intermediate a double electron transfer converts the active iron centre to an

oxidation state of Fe(III) with a doublet orbital occupation of x2-y22 *xz

2 *yz1 a2u

2 ФR

0 or a

quartet orbital occupation of x2-y22 *xz

1 *yz1 *z2

1 a2u

2 ФR

0,

where ФR represent the orbital

N N

N NFeIV

O

S

Cys

HOO

OHO

22 yx

2*z

xy*

yz* ua2

ua1

xz*

Quartet spin

Doublet spin

Page 52: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

52

on the substrate . In the radical intermediate however, an electron transfer from the aromatic

substrate to the oxidant occurs hence giving a Fe (IV) type complex with an electronic

configuration of x2-y22 *xz

1 *yz1 a2u

2 ФR

1. As can be observed from Figure 1.15, the doublet

spin orbital occupation in the cationic intermediate is found to be very much lower in energy

as compared to the radical intermediate pathway. In the subsequent step the ipso-proton is

abstracted by a nitrogen atom of the porphyrin forming an intermediate of the proton transfer

which leads to the restoration of the aromaticity of the substrate ring; the proton transfer

intermediate is known to be an enormously exothermic process. In the final step of the

reaction the hydrogen atom bounded to the nitrogen atom of the porphyrin ring rebound back

to either the oxo group to form either an alcohol product or to the ortho-carbon to form

ketone.

1.11 Dissertation summary

Clearly, homogeneous catalysts have important chemical, industrial and biological functions.

However, there is only limited understanding of the molecular details of what happens during

a reaction mechanism. Evidently there are many un-answered questions on the catalytic

behaviours of verse majority of synthetic catalysts both bio-dependent and industrial catalysts

which requires detailed computational studies in order to explicitly understand there

behaviours down to an electronic level. To contribute on the understanding of these

mechanistic behaviours on a quantum level, we studied a few synthetic complexes during the

period of my doctorate degree. The quantum mechanical calculations done within this thesis

is believed to have contributed in the understanding of the mechanisms of these reactions as

Page 53: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

53

well as the origin of the reaction pathways and environmental perturbation affecting the

mechanisms. Below is a brief description of each of the chapters elaborated within the thesis.

Chapter 3:

This chapter reported the computational results I have obtained on a synthesised crystal

structure of 1,2-bis[-hydroxo iron(III) 5-(2,3,7,8,12,13,17,18-octaethylporphyrinyl)]ethane

with I3–, ClO4

- and BF4

– counter anions which shows some unique behaviour of the system

from a spectroscopic characterisation shown by Sankar and co-workers (Ghosh, S.; Rath, S.

2010). The synthetic dimer system shows a character of two unequivalent spin states with one

iron centre having an admixed intermediate-spin state of S=3/2 with a slight contribution of

S=5/2, and a S=

5/2 on the second iron or a spin state of either S=

5/2, S=

5/2 or S=

3/2, S=

3/2 on

both of the iron centre depending on the counter anion approaching. The spectroscopic

characterisation reported that, as the approaching counter anion altered, the spin state

ordering on the -hydroxo complexes changes. In order to fully understand and rationalise

how the spin state ordering is affected by external perturbations, we also have done a

comprehensive computational benchmarking which was also thoroughly discussed in the

chapter. Our calculations demonstrated that subtle environmental perturbations, such as

entropic corrections ruffles the spin state ordering and relative energies and are likely to be

the root cause of the variation in spin state ordering observed experimentally.

Chapter 4:

In this chapter, I have synthesized a model complex and studied it with low-pressure Fourier

transform-ion cyclotron resonance (FT-ICR) mass spectrometry (MS). I found out that all

Page 54: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

54

substrates react with the selected substrates by a more or less efficient oxygen atom transfer

process; Whereas substrates with low ionization energies react by a charge-transfer channel,

which enabled me to determine the electron affinity of [FeIV(O)(TPFPP+•)]+ for the first

time. The computational results we presented here confirm the observed trends excellently

and rationalize the reactivities within the framework of thermochemical considerations and

valence bond schemes.

In summary my findings gave an excellent agreement with ideal-gas measured rate constants

and have provided a general trend in epoxidation reactions.

Chapter 5:

This chapter basically report a systematic study into the core chemical properties of the

Cytochrome P450 oxidant and substrates in order to understand the factors that affect

reactivity patterns. The work purposely focused on investigating the effect bestowed by an

epoxidation of styrene and a para-substituted styrene substrate by a biomimetic compound I

with either an anionic or a neutral axial ligand. We chose chloride as our anionic ligand and

an acetonitrile for our cationic axial ligand. Our findings reported that the activation enthalpy

of the reaction is determined by the ionization potential of the susbstrate, the electron affinity

of the oxidant and the C-O bond strength which is determined by the bond dissociation

energy (BDEOH). We also generated a model that enable us to predict the rate constants and

reactivities of substrates epoxidation reactions by iron(IV)-oxo porphyrin cation radical

oxidants.

In summary the discussion presented within the chapter states that electron withdrawing

substituents lead to early transition states while electron donating groups on the substrates

Page 55: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

55

offers late transition states. This trends affects the barrier heights in such a way that electron

withdrawing substituents correlate the barrier height with BDEOH, while the electron affinity

of the oxidant is proportional to the barrier height for substrates with electron donating

substituents.

Chapter 6:

The calculations within this chapter gave a multistate reactivity patterns, whereby the

bifurcation of hydroxylation versus desaturation was detailed. The product distributions differ

on each of the spin state surfaces; hence we find spin-selective product formation. I did a

thorough analysis of the electronic and thermochemical factors that determine the bifurcation

pathways and I come up with a model that predicts the regioselectivity of the aliphatic

hydroxylation over desaturation pathways from valence bond and molecular orbital theories.

Chapter 7:

This chapter summarises the work I have done in order to establish the key features of a

synthetic homogeneous catalyst that is capable of catalysing an alkyl chain growth. I have

shown some detailed results we achieved on a combined DFT, an NBO and a VB analysis;

and also a detailed thermochemical studies has been performed and establish the intrinsic

properties of the complex. Initially I analyse the reactants as having a Ru2Pt-carbene core

where the carbene’s lone-pair transfers to the Pt through bonding configuration. The

characterisation also shows that a low energy mechanism of alkyl formation via consecutive

CH3• followed by H• transfer to the bridging carbene to form alkyl chain is feasible.

Conclusively, my work identifies a novel catalyst for the synthesis of alkanes that starts from

Page 56: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

56

a trimetal carbene. This unique structure is shown to be capable of intramolecular methyl and

hydrogen atom transfer to the carbene to form 2-CHCH3 products as precursor to alkanes.

Chapter 8:

The computational analysis presented in this chapter gave an extended knowledge on the

importance of distal hydrogen bonding on both heme and non-heme complexes. I show for

the first time, the effect of hydrogen bonding interaction with the distal Oxo ligands. It is

known that Iron (IV)-oxo intermediates are actively involved in several hydrogen bonding

interactions with its neighbouring molecules in an active site of an enzyme which triggers our

curiosity to gain a better understanding on it. The calculation therefore gave a new insight

into the distal hydrogen bonding and it shows that the hydrogen bond may play a vital role in

proton relay mechanisms in the formation of metal-oxo intermediates thereby decreasing the

hydrogen atom abstraction ability of the intermediate. Indeed, in nonheme iron enzymes,

where no proton relay takes place, there generally is no donating hydrogen bond to the

iron(IV)-oxo moiety.

Page 57: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

57

CHAPTER 2

METHODOLOGY

Page 58: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

58

Computational techniques have been developed with the aim to predict and improve

understanding of molecular behavior. The widely known and accepted procedures used in the

calculation of atomistic systems include ab-initio methods, in which a system is calculated

from scratch without any prior knowledge of the chemical system, whereas in

(semi)empirical Quantum mechanics methods experimentally derived parameters are taken

into account and build upon. By contrast, Molecular mechanics method (MM) are commonly

used for calculations of very large systems (typically with tens of thousands of atoms, such as

proteins). However, the MM methods do not model bond-formation and bond breaking

processes. In all methods aforementioned, the accuracy of the result solidly depends on the

computational theory applied. In order to combine the accuracy of QM methods with the

speed of MM methods, the combined quantum mechanics/molecular mechanics (QM/MM)

method has been developed. QM methods generally are computed based on the Schrödinger

equation (equation 2.0), which was introduced by the Austrian physicist Erwin Schrödinger

(Schrödinger, E., 1926), in 1926 and is the basic foundation of the electronic structure

calculations.

H = E

In Eq 2.0, H denotes the Hamilton operator, represents the eigenfunction for a given

Hamiltonian (i.e. the wavefunction), and E is the energy of the system. The wavefunction

takes the position of the electrons and nuclei in the system as variables, leading to the

following equation:

𝐻 Ψ 𝑖 (𝑥1, … , 𝑥𝑁 , 𝑅1, … , 𝑅𝑀) = 𝐸 Ψ𝑖(𝑥1, … , 𝑥𝑁 , 𝑅1, … , 𝑅𝑀) (2.1)

Page 59: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

59

The position of the electrons in Eq 2.1 is described with 𝑥𝑁, while N and 𝑅𝑀 describe the

position of the nuclei, M. The properties of the system will be assumed using the knowledge

of 𝛹 in the equation.

The wavefunction is assumed to be orthonormal in space as represented in the equation 2.2

below (Schrödinger, E., 1926):

⟨Ψ𝑖|Ψ𝑗⟩ = 𝛿𝑖𝑗 (2.2)

The term 𝛿𝑖𝑗 is known as the Kronecker symbol, with a range as indicated 𝛿𝑖𝑗 = 1 𝑖𝑓 𝑖 =

𝑗 and 𝛿𝑖𝑗 = 0 𝑓𝑜𝑟 𝑖 ≠ 𝑗.

2.1 Solving the Schrodinger equation

To solve the Schrodinger equation, the Hamilton operator needs to be established in the

following form (Schrödinger, E., 1926):

𝐻 = 𝑇𝑒 + 𝑇𝑛 + 𝑉𝑛𝑒 + 𝑉𝑒𝑒 + 𝑉𝑛𝑛 (2.3)

Where the terms 𝑇𝑒 and 𝑇𝑛 representing the terms for kinetic energy of the electrons and

nuclei, The value 𝑉𝑛𝑒 is the attractive potential between the nuclei and the electrons, and the

potentials 𝑉𝑒𝑒 and 𝑉𝑛𝑛 represent the electron-electron and nuclei-nuclei repulsion energies

respectively. The individual terms can be equated into the following expressions:

𝑇𝑒 = −1

2 ∑ ∇𝑖

2

𝑁

𝑖=1

(2.4)

𝑇𝑛 = −1

2 ∑

1

𝑀𝐴∇𝐴

2

𝑀

=1

(2.5)

Page 60: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

60

𝑉𝑛𝑒 = − ∑

𝑁

𝑖=1

∑𝑍𝐴

𝑟𝑖𝐴

𝑀

𝐴=1

(2.6)

𝑉𝑒𝑒 = − ∑

𝑁

𝑖=1

∑1

𝑟𝑖𝑗

𝑁

𝑗>1

(2.7)

𝑉𝑛𝑛 = − ∑

𝑀

𝐴=1

∑𝑍𝐴𝑍𝐵

𝑅𝐴𝐵

𝑀

𝐵>𝐴

(2.8)

Substituting the above terms into equation 2.3 above gives the following equation 2.9:

𝐻 = −1

2 ∑ ∇𝑖

2

𝑁

𝑖=1

−1

2 ∑

1

𝑀𝐴∇𝐴

2

𝑀

𝐴=1

− ∑

𝑁

𝑖=1

∑𝑍𝐴

𝑟𝑖𝐴

𝑀

𝐴=1

+ ∑

𝑁

𝑖=1

∑1

𝑟𝑖𝑗

𝑁

𝑗>1

𝑀

𝐴=1

∑𝑍𝐴𝑍𝐵

𝑅𝐴𝐵

𝑀

𝐵>𝐴

(2.9)

The A and B denote the two individual atoms in molecule M (or AB), whereas i and j

represent the N number of electrons within the system.

The equation above is the full Hamiltonian for a bimolecular system where the first

expression represents the kinetic energy contribution of the electrons; the second expression

represents the kinetic energy of nuclei M, the third expression is the electron-nucleus

attraction, the fourth expression is the nucleus-nucleus association and the last is representing

the electron-electron repulsion (Schrödinger, E., 1926). However, as mentioned earlier, the

expression in equation 2.9 above gets more complex according to the size of the system. This

equation is the basis of the electronic Schrödinger equation and is solved in Quantum-

chemical calculations.

Page 61: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

61

2.2 Born-Oppenheimer approximation

The Born-Oppenheimer approximation is the foremost of quite a few approximations used in

quantum chemistry to simplify solving the Schrödinger equation. The molecular challenge

encountered by the Schrödinger equation is simplified by separating nuclear and electronic

motion by this approximation which is reasonable due to the fact that the mass of the electron

is 1822 times smaller than the mass of the proton (Millikan R, A., 1911). Therefore the nuclei

move very slowly with respect to the movement of the electrons, and the electrons react

immediately to changes in nuclear position. Consequently, the electron distribution within a

molecular system will depend on the positions of the nuclei, and not on their velocities.

Literally, the nuclei are fixed with respect to the position of the electrons, and electronic

motion can be described as occurring in a field of fixed nuclei. This gives rise to a

wavefunction of an electron that describes the electrons in the field of the nuclei; therefore

the Hamiltonian is defined as the following equation (Atkins, P., W. and Friedman, R., S.

1997):

𝐻𝐵𝑂 = −1

2 ∑ ∇𝑖

2 − ∑ ∑𝑍𝐴

𝑟𝑖𝐴+ ∑ ∑

1

𝑟𝑖𝑗 + 𝑉𝑛𝑚

𝑁

𝑗>1

(2.10)

𝑁

𝑖=1

𝑀

𝐴=1

𝑁

𝑖=1

𝑁

𝑖=1

The 𝑉𝑛𝑚 is a constant representing the nucleus-nucleus repulsion; as can be observed the

equation 2.10 can factorized to give equation 2.11 below;

𝐻𝑒𝑙𝑒𝑐 = ∑ (−1

2 ∇𝑖

2 − ∑𝑍𝐴

𝑟𝑖𝐴+ ∑

1

𝑟𝑖𝑗

𝑗>𝑖𝐴

) (2.11)

𝑖

Page 62: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

62

Now, the wavefunction depends solely on the number of electrons in the system as indicated

below:

𝐻𝑒𝑙𝑒𝑐 Ψ 𝑖(𝑒𝑙𝑒𝑐) (𝑥1, 𝑥2, … , 𝑥𝑖, 𝑥𝑗 , … , 𝑥𝑁) = 𝐻𝑒𝑙𝑒𝑐 Ψ 𝑖(𝑒𝑙𝑒𝑐)(𝑥1, 𝑥2, … , 𝑥𝑖, 𝑥𝑗 , … , 𝑥𝑁) (2.12)

The electrons can be described by their spin quantum number (ms) which can take up a value

of either – ½ or ½; the definition is based upon the alignment of the spin with respect to an

arbitrary axis; where x represent the probability of finding an electron at a given point in

space. The aforementioned spins are known as and they are spinfunctions for ms = + ½

and -½ correspondingly; and are considered orthonormalised:

⟨𝛼|𝛼⟩ = ⟨𝛽|𝛽⟩ = 1 (2.12)

⟨𝛼|𝛽⟩ = ⟨𝛽|𝛼⟩ = 0 (2.13)

Generally, the wavefunction is split in to one-electron wavefunctions, that contain a spatial

component and a spin component as represented in equation 2.14:

Ψ (𝑥) = Ψ (𝑟). 𝜎 𝜎 = 𝛼 𝑜𝑟 𝛽 (2.14)

Page 63: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

63

2.3 Slater determinant

2.3.1 Anti-symmetry Principle

The wavefunction Ψ is not visible; however it is expressed according to the representation in

equation 2.15 (Atkins, P., W. 1977);

|Ψ(𝑥1, 𝑥2, … , 𝑥𝑁)|2𝑑𝑥1𝑑𝑥2 … 𝑑𝑥𝑁 (2.15)

Where 𝑑𝑥1𝑑𝑥2 … represent the probability of finding an electron at a specific point in space.

As electrons are equivocal and indistinguishable, the probability does not change by the

exchange of two electrons;

|Ψ(𝑥1, 𝑥2, … , 𝑥𝑖𝑥𝑗 , … , 𝑥𝑁)|2

= |Ψ(𝑥1, 𝑥2, … , 𝑥𝑖𝑥𝑗 , … , 𝑥𝑁)|2

(2.16)

However, the sign of the wavefunction is changed because of the exchange of two electrons;

the Ψ becomes anti-symmetrical with respect to the electron changes which are a clear

representation of the quantum-mechanical generalization of the Pauli’s exclusion principle

which states that no two electrons can occupy the same state within a chemical system (Krane

K. S., 1987). Since the precise wavefunction is unknown, a trial wavefunction that obeys the

anti-symmetrical principle is carefully generated. A trial wavefunction is generated and

conforms to the rule with anti-symmetry product of N one-electron wavefunctions 𝑥𝑖(𝑥𝑖).

This product is represented by the symbol Φ𝑆𝐷 and is referred to as the Slater determinant

(Atkins P. W., 1977):

Φ𝑆𝐷 = 1

√𝑁! |

𝑥1(𝑥1) 𝑥2(𝑥1) … 𝑥𝑁(𝑥1)

𝑥1(𝑥2) 𝑥2(𝑥2) … 𝑥𝑁(𝑥2)

𝑥1(𝑥𝑁) 𝑥2(𝑥𝑁) … 𝑥𝑁(𝑥𝑁)| (2.17)

Page 64: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

64

The columns in the representation above are single electron wavefunctions (orbital), x(x),

whereas the row represents the electron indices.

In order to establish the best wavefunction to describe a chemical system, the variational

method (also known as Rayleigh-Ritz variational method) is used. It is based on the two most

pronounced methods known as the Hartree-Fock method and Density functional theory

(DFT) (Abdulsattar, Mudar A. 2012). The variational method states that the value calculated

for the total energy of a trial wavefunction can only be greater than or equal to the ground

state energy 𝐸0.

𝐸 = < Ψ|𝐻|Ψ > ≥ 𝐸0 (2.18)

2.4 Hartree-Fock theory (HF)

In Hartree-Fock theory, the Hamiltonian is typically by two individual components: 1) The

core Hamiltonian, Hc(i), which describes the kinetic energy of the electrons as well as the

electron-nucleus attraction, and 2) The electron-electron repulsion:

𝐻 = ∑ [𝐻𝑐 (𝑖) + ∑1

𝑟𝑖𝑗𝑗>𝑖

]

𝑖

𝑎𝑛𝑑 𝐻𝑐(𝑖) = −1

2 ∇𝑖

2 − ∑𝑍𝐴

𝑟𝑖𝐴 (2.19)

𝐴

The core Hamiltonian can be solved exactly as it is, but the electron-electron repulsion must

be treated in such a way that each electron is considered to be moving independent of other

(neighboring) electrons in an average field created by the other electrons. By implementing

the variational method to a single Slater determinant Φ𝑆𝐷, the lowest energy calculation is

possible through the optimsation of the values for 𝑥𝑖. The equation obtained is called

Page 65: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

65

“Hartree-Fock equation (HF)” and solving the HF equation determines the best spin orbitals

at minimized energy E.

𝑓𝑖𝑥𝑖 = 𝜖𝑖𝑥𝑖 (2.20)

In Eq 2.20, 𝑥𝑖 is an eigenfunction of a Fock operator 𝑓, and 𝜖𝑖 is the corresponding orbital

energy for this eigenstate. The eigenvalue of the Fock operator that is negative, i.e. associated

with a spin orbital,−𝜖𝑖, actually corresponds to the ionization potential according to

Koopmans Theorem (Koopman, Tjalling, 1934). The Fock operator is an effective single-

electron operator with the form as follows:

𝑓𝑖 = −1

2 ∇𝑖

2 − ∑𝑍𝐴

𝑟𝑖𝐴+ 𝑉𝐻𝐹(𝑖)

𝐴

(2.21)

In Eq 2.21, the term 𝑉𝐻𝐹(𝑖) is known as the Hartree-Fock potential and it is a representation

of the average repulsive potential by each electron due to the field of N-1 electron. It replaces

the 1

𝑟𝑖𝑗 repulsion operator which was considered too complex to solve. Although by adjusting

the electronic repulsion of different electrons within an average potential, the equation is

solvable. Therefore 𝑉𝐻𝐹comprises of two terms:

𝑉𝐻𝐹(𝑥1) = ∑ (𝐽𝑗(𝑥1) − 𝐾𝑗(𝑥1))

𝑗

(2.22)

𝐽𝑗(𝑥1) = ∫|𝑋𝑗(𝑥2)|2

1

𝑟12 𝑑𝑥2 (2.23)

Page 66: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

66

𝐾𝑗(𝑥1)𝑋𝑖(𝑥1) = ∫ 𝑋𝑖 (𝑥1)1

𝑟12 𝑋𝑖 (𝑥2)𝑑𝑥2 𝑋𝑗 (𝑥1) (2.24)

The operator 𝐽𝑗 represents the potential experienced by an electron at a point 𝑥1 because of

the average charge distribution of another electron within the spin orbital 𝑋𝑗, while the term

𝐾𝑗 has no known classical interpretation. It leads to an exchange of the variables within the

two spin orbitals. To solve the Hartree-Fock equation, the two aforementioned methods can

be applied depending on the instance faced. If a system is closed shell, i.e. containing an even

number of electrons, all the electrons are paired up and the Restricted Hartree-Fock (RHF)

method is used to solve it. However, if there is an odd number of electrons or an even number

of electrons with an unpaired electrons arrangement within the orbitals, the Unrestricted

Hartree-Fock (UHF) can be used. The Roothaan-Hall equation is generally used to solve the

RHF equations, whereas in the case of the UHF, the Berthier-Pople-Nesbert equation is used

to solve the integrals. To solve the 𝑉𝐻𝐹 description of equation 2.22 above, it is essential to

obtain 𝑋, therefore an iterative methodology known as the “Self-Consistent Field method

(SCF)” is required. An initial guess for 𝑉𝐻𝐹 is required to calculated 𝑋 which is used to

generate a new 𝑉𝐻𝐹. The process continues until the cycle is converged based on the criteria;

i.e. until successive potentials are identical.

2.5 Electron Correlation

In the Hartree-Fock (HF) method, the electrons are considered to be moving freely within an

average electronic field of the nuclei so the correlated motion of individual electrons

perturbed by the other electrons is omitted. As a consequence, the total HF energy is always

higher than the real energy (Abdulsattar, Mudar A. 2012). The difference between the

Page 67: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

67

Hartree-Fock energy and the real energy is the correlation energy and the energy gap is the

electronic correlation energy.

𝐸𝑐𝑜𝑟𝑟𝑒𝑙𝑎𝑡𝑖𝑜𝑛 = 𝐸𝑡𝑜𝑡𝑎𝑙 − 𝐸𝐻𝐹 (2.25)

The electron correlation term constitutes only a tiny proportion of the overall energy of the

system, however, it has a large impact on the calculated chemical and physical properties of

the system (Riley, K. E.; Op’t Holt, B. T.; et al 2007). The correlation energy is usually

introduced by taking into account the excitation of one or more electrons from occupied

orbitals to virtual orbitals. The electrons then adapt a specific orbital arrangement that

depends on the number of electrons involved within the system and the Slater determinant is

used to describe each of these states. The sum combination of all states gives the new trial

wavefunction which is considered closer to the real system than the initial determinant.

Ψ = 𝑐0 Φ𝐻𝐹 + ∑ 𝑐𝑖Φ𝑖 (2.26)

𝑖=1

In Eq 2.26, the summation is over all probable excited states. 𝑐𝑖 are the coefficients that

define the contribution of each excited state to the wavefunction.

2.6 Density Functional Theory (DFT)

Density functional theory (DFT) is based on electron densities instead of wavefunction. An

essential step on the way to developed was taken by Slater in the 1950s, where he applied his

XR method (Extrapolated Run method) for molecules and solid state systems. Thus, one-

parameter approximate exchange correlation functional (Slater, Morton. 1950) was used.

Page 68: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

68

However, DFT did not become a full-fledged theory until the formulation of the Hohenberg

and Kohn theorems in 1964 (Hohenberg, P.; Kohn, W. 1964). They introduced orbitals into

the picture the Kohn-Sham formalism, which gave a computational breakthrough which was

later implemented in e.g., Pople’s GAUSSIAN software packages (Frisch, J. A. et al 1986).

Although, there are a number of popular and highly accuarate wave function packages in use,

DFT was further developed as a computationally attractive alternative.

Density functional theory provides a relative efficient and unbiased tool with which to

compute the ground state energy of realistic models of mass materials and their surfaces. The

calculation solidly relies on the development of some approximations for the exchange-

correlation energy functional (Harrison, J.G., 1987). However, remarkable advances have

been achieved in the recent decade in the development of an exchange-correlation functional

that is dependent on local density gradients, as well as semi-local measures of the density and

nonlocal exchange functional (Frisch, J. A. 2009). Density functional theory describes the

energy of a system as a function of the electron density which is in contrast to wave function

quantum mechanics (Frisch, J. A. 2009). Density functional theory method is widely accepted

and proven as a very successful and computationally less expensive as compared to other

methods, it also gives a result that reasonably agree with experiment for relatively large

chemical systems. However, it is less accurate for binding energies and details of the energy

of the surface away from equilibrium geometries. The total energy of a DFT calculation of

the electronic energy 𝐸𝑒𝑙𝑒 and is expressed by equation 2.27 below (Parr, R. G.; Yang, W.

1989);

𝐸𝑒𝑙𝑒 = −1

2∑ ∫ ∅𝑖(𝑟1)

𝑖

∇2∅𝑖(𝑟1)𝑑𝑟1 + ∑ ∫𝑍𝐴

|𝑅𝐴 − 𝑟1|𝐴

𝜌(𝑟1)𝑑𝑟1 +1

2∫

𝜌(𝑟1)𝜌(𝑟2)

|𝑟1 − 𝑟2|𝑑𝑟1𝑑𝑟2 + 𝐸𝑥𝑐 (2.27)

Page 69: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

69

The first term expresses the kinetic energy of the system, the second term represents the

attraction between electron and nucleus in the system, the next is the Coulomb interaction

between two individual electrons, whereas the last term represents the exchange-correlation

functional (Exc) (Ziegler, 1991). In Eq 2.27, 𝜌(𝑟) is the electron density, r is a term

representing the interaction between electron-electron, electron-nucleus. Thus, in DFT the

solution to the electronic energy (𝐸𝑒𝑙𝑒) follows from the sum of the integrals given in equation

2.27. In addition to the kinetic energies of the electrons, Coulumbic electron-nucleus

attraction, and electron-electron repulsions, there is an exchange-correlation term 𝐸𝑥𝑐 for the

electron pair interaction with an individual exchange and a correlation component Ex and Ec

respectively.

The exchange-correlation functional is unknown, therefore, it is approximated to determine

its contribution to DFT calculations. The exchange-correlation component is divided into an

exchange energy term (equation 2.28) also known as the Slater exchange, and the correlation

energy usually taken from Vosco, Wilk and Nusair (equation 2.29).

𝐸𝑥𝑆𝑙𝑎𝑡𝑒𝑟 = −

9

4𝛼𝑒𝑥(

3

4𝜋)

13

∑ ∫[𝜌1𝛾(𝑟1)]

43

𝛾

𝑑𝑟1 (2.28)

𝐸𝑐𝑉𝑊𝑁 = ∫ 𝜌

1 (𝑟1)휀𝑐[𝜌

1𝛼 (𝑟1), 𝜌

1𝛽 (𝑟1)]𝑑𝑟1 (2.29)

The exchange scale factor αex in the equation has the value 2/3 for an electron gas (Parr, R. G.;

Yang, W. 1989) and the correlation is calculated per electron in a gas from the correlation

Page 70: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

70

functional 휀𝑐 |𝜌1𝛼 , 𝜌1

𝛽| using spin densities 𝜌1

𝛼 and 𝜌1𝛽

. The main reason behind the use of the

Slater exchange 𝐸𝑥𝑆𝑙𝑎𝑡𝑒𝑟and the correlation energy 𝐸𝑐

𝑉𝑊𝑁 is to obtain a wavefunction-type

method, although it was found not to be good enough to compete with other methods such as

MP2; therefore, further corrections due to non-local interactions were required. These were

implementation by, for instance, Lee, Yang and Parr (LYP) and Perdew and Wang (PW95) for

correlation methods and Becke for exchange methods (Becke, A. D. 1996, Lee, C.; Yang, W.

et al 1988).

2.7 Functionals

2.7.1 Local Density Approximation (LDA)

The exchange-correlation functionals are based upon the local density approximation (LDA),

which use the electron density of a uniform electron gas. Even though the LDA method is a

rough approximation, it is the one system that is known to define the density as 𝜌 = 𝑁

𝑉

(where the number of electrons is represented as N, and volume of the gas is denoted as V ),

and the exchange as well as the correlation energy functional are known to a very high

accuracy. The electronic density is 𝜌 is replaced by the spin electronic densities 𝜌𝛼 and 𝜌𝛽

such that 𝜌 = 𝜌𝛼 + 𝜌𝛽 when dealing with an open shell systems. This approximation is

termed the Local Spin-Density Approximation (LSDA). One known instance of an LDA type

functional is the one generated by Vosko and co-workers based on high quantum Monte Carlo

calculation for a uniform electron gas. More accurate results than those obtained from HF

theory are calculated when molecular properties of a system such as vibrational frequencies,

charge moments, elastic moduli etc, are determined. However, generally, the LDA functional

Page 71: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

71

gives occasionally serious flaws and poorly characterizes energetic details, including bond

energies and reaction mechanisms.

2.7.2 Generalized Gradient Approximation (GGA)

Although the LDA method discussed above is known to calculate the exchange and

correlation energies to a very high accuracy, it actually describes energies rather badly, and

hence a new type of functional was introduced known as the “Generalised Gradient

Approximation (GGA)”. This functional uses the gradient of the electron density ∇𝜌. The

inclusion of the electron density gradient assist in describing the non-homogeneity of electron

density rather more realistically. The GGA is usually sectioned into an exchange term and a

correlation term, which can then be solved independently:

𝐸𝑋𝐶𝐺𝐺𝐴 = 𝐸𝑋

𝐺𝐺𝐴 + 𝐸𝐶𝐺𝐺𝐴 (2.30)

2.7.3 Hybrid Functionals

Computational chemistry was greatly enhanced by introduction of the hybrid density

functional procedure developed by Becke. He compared information from experimental test

sets, namely ionization energies, electron and proton affinities, and compared it against DFT

calculated values. The result gave a value with high accuracy which demonstrated the

accuracy of the Hybrid functional method. Although, many different combinations between

exchange and correlation functionals are possible, currently the most commonly used one is

the Becke 3-Parameter (Exchange) Lee, Yang and Parr (correlation; density functional

theory) B3LYP (Lee, C.; Yang, W. 1988). The hybrid density functional method B3LYP is

expressed as the following equation:

Page 72: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

72

EXCB3LYP = AEX

Slater + (1 + A)EXHF + BΔEX

Becke + ECVWN + CΔEC

LYP (2.31)

Hence, the equation comprises of the Slater and Vosko-Wilk-Nusair local density

approximation functions, the Hartree-Fock exchange, and a correction term for the exchange

due to Becke and Lee, Yang and Parr’s correlation functional (Becke, 1993). As mentioned

earlier the B3LYP hybrid functional is a mixture of the LDA and GGA functional, as

presented in equation 2.32 below:

EXCB3LYP = EXC

LDA + 𝑎0 (EXHF − EX

LDA) + 𝑎𝑋 (EXGGA − EX

LDA) + 𝑎𝐶 (𝐸𝐶𝐺𝐺𝐴 − 𝐸𝐶

𝐿𝐷𝐴) (2.32)

Where the terms a0 = 0.20, 𝑎𝑋 = 0.72 and 𝑎𝐶 = 0.81 are three empirical parameters

determined by fitting the predicted values to a set of atomization energies, proton affinities,

ionization potentials and total atomic energies of a system (Becke, 1993).

2.8 Basis Sets

Atomic orbitals and molecular orbitals are described and created using basis functions. They

are denoted as a linear combination of such functions with the coefficients to be determined.

The basis-functions are assumed to be obtained at the atomic nuclei at the centre of the atom

and so bear some resemblance to an atomic orbital. The basis functions are categorized into a

Slater-type orbital (STO), and the Gaussian-type orbital (GTO).

2.8.1 Slater-type Orbital (STO)

The Slater-type orbitals (STO) have the exponential dependence 𝑒−𝜁𝑟 and are very close in

their mathematical expression to the real atomic orbital. It is also worth mentioning that

solving the Schrodinger equation for a H-atom gives the Slater-type orbitals.

Page 73: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

73

휂𝑆𝑇𝑂 = 𝑁𝑟𝑛−1𝑒−𝜁𝑟𝑌𝑙𝑚(휃, 𝜙) (2.33)

Where the term N represents a factor of normalization, 휁 is the exponent; 𝑌𝑙𝑚is the angular

momentum which is the function describing the shape of the orbital; r, 휃 and 𝜙 are spherical

coordinates, while n, l and m are classical quantum numbers representing principal, angular

and magnetic momentum respectively.

The STO very closely described the behavior of atomic orbitals of hydrogen because it

features a cusp at r = 0 and a good exponential decay for larger values of r.

2.8.2 Gaussian-type orbital (GTO)

The Gaussian-type orbitals (GTO) have the exponential dependence𝑒−𝛼𝑟2:

휂𝐺𝑇𝑂 = 𝑁 𝑥𝑙𝑦𝑚 𝑧𝑛 𝑒−𝑎𝑟2 (2.34)

Where N is a normalization factor, x, y and z are the Cartesian coordinates. Although

Gaussian-type orbitals are not really orbitals as they are simpler functions and are frequently

referred to as the “Gaussian Primitive” (Lowe, John P. 1978). Gaussian primitives are

usually obtained from quantum calculations on Hartree-Fock or Hartree-Fock plus some

correlated calculations. Typically the exponents of x, y and z are varied until the lowest total

energy of the atom is attained. For molecular calculations, these Gaussian primitives have to

be contracted thereby the Cartesian linear combinations of them will be used as basis

functions. The basis functions will have its coefficients and exponents fixed.

The GTO, in contrast to the above STO explained, does not show a cusp at r = 0 and steeply

decrease for large values of r. Despite the challenges mentioned, the GTO are better because

the product of two GTOs centered on two different atoms is a third GTO sandwiched between

Page 74: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

74

them, which is not the case with an STO basis function and that made them very difficult to

handle computationally as the four-centre-two-electron integrals are time consuming.

The precision and degree of complexity of a basis set is defined by the number of contracted

functions involved in representing each atomic orbital, the minimum being one contracted

function to describe a basis function. For instance, the basis set STO-3G is formed by a linear

combination of three contracted Gaussian function (represented by G) so as to resemble one

STO. At least, usually two or more basis functions can be used to describe each type of

orbital for more precision and better descriptions. The valence electrons are known to be

actively involved in chemical reactions and therefore, it is essential to have a flexible

description of its electrons. Such basis sets with valence electrons is defined with more

flexibly and differently from the core orbitals and are generally referred to as “Split Valence

Basis Sets”. An example of a split valence basis sets most commonly used in electronic

structure calculations is the 6-31G basis set. These type of basis sets are represented in the

format X-YZG, with nomenclature as follows:

X is the number of primitive GTOs used in describing one single contracted Gaussian

function of the core.

Y and Z are the number of primitive GTOs describing the valence orbitals; more can

be added to increase precision.

Therefore, considering the above nomenclature, the split valence basis set mentioned earlier

as 6-31G comprises two functions, one containing three primitive and the other contain only

one.

Page 75: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

75

2.8.3 Polarisation and Diffuse basis functions

Polarization and diffuse functions can also be included to a basis sets. The polarization

represents the deformation of the electronic cloud, which is induced by the bond between two

atoms. To define this characteristic, functions with higher angular momentum are added to

the basis set. For instance, the addition of a p function to a hydrogen atom induces

polarization; the same applies when a d function is added to a basis set containing p valence

orbitals, and an f function for a d-valence orbitals. Even more precision is increased by the

addition of for instance, p and d polarization functions to a hydrogen atom with 6-31G basis

set thereby becoming a 6-31G(pd) basis set.

The diffuse function, on the other hand explains the section of atomic orbitals distant from

the nuclei that could play a very important role when considering anions or diffuse electronic

clouds; for instance, in second or third row transition metals. The diffuse function was also

found useful when included in calculating anions containing lone pairs.

Another circumstance worth mentioning is the Effective core potential used to describe the

inner core of transition metal. Transition metals have a very large inner core and so the

number of basis functions used to describe it would be very large and that could hinder with

the accuracy of the calculations hence the application of the effective core potential comes in

handy. The ECP will model the effect of the nucleus and the electrons from the inner shell on

the valence electrons as an average effect; these reduce the computational time and includes

the relativistic effects on the system studied as these basis functions are generated from

relativistic atomic calculations.

Page 76: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

76

2.9 Zero-Point Energy

The Born-Oppenheimer approximation is almost always used in the field of quantum

chemistry (Lowe, John P. 1978). Therefore, the most common definition of the molecular

zero-point energy (ZPE), the energy difference between the vibrational ground state and the

lowest point on the Born-Oppenheimer potential energy surface (Lowe, John P. 1978).

However, the definition is not applicable to experimental ZPE as the Born-Oppenheimer

approximation (BOA) is never taken on by real molecules. Experimentally, spectra are

analyzed, though the explicit Born-Oppenheimer approximation correction is only made

when simultaneously fitting of different isotopologs to effective potential. Therefore, the

experimental zero-point energy is defined as the difference between the molecular ground

state and the lowest point on its isotope-specific effective potential. The variation between the

aforementioned definition and the definition used by the quantum chemist must be absorbed

by the empirical scaling factor that is applied in order to determine zero-point energy

theoretically (K.Irikura, 2006).

The computational procedure of determining the ZPE is by solving the secular determinant

which gives an orbital energies that sum up the electronic energy (Ee), that represent the

ground state energy of the gaseous system at 0K temperature. However, since the nuclei

occupy vibrational energy levels, the energy obtained is not the correct absolute energy,

therefore, the electronic energy needs to be corrected for ZPE. Computationally, the ZPE

correction is found from a geometry optimization followed by a vibrational frequency

calculation which estimates the vibrational energy levels and susbsequntly the ZPE value.

The energy of the lowest vibrational energy level is expressed with equation 2.35 below:

Page 77: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

77

𝐸0 = 𝐸𝑒 + 𝑍𝑃𝐸 (2.35)

In Eq 2.35, 𝐸0 is the actual (ZPE corrected) energy of the system at 0 K and 𝐸𝑒 is the

electronic energy which represents the ground state energy of the system at 0 K temperatures.

A frequency calculation can also be used to determine the nature of a stationary point found

by a geometry optimization. There are two portion of information from the frequency output

which are essential to typifying a stationary point (Foresman, J., B. 1993):

The number of imaginary frequencies

The number of negative eigenvalues in the Hessian

Imaginary frequencies are listed in the output of a frequency calculation as negative numbers.

A structure with N number of imaginary frequencies is an nth

order saddle point (Foresman,

J., B. 1998). When a structure contains an imaginary frequency, the calculation indicates that

there is some geometric distortion for which a lower energy structure can be found.

Sometimes, a calculation stuck in a high symmetry constraint and lowering the symmetry of

the structure leads to a more stable structure (Foresman, J., B. 1993). To differentiate between

these two cases, the eigenvalue of the Hessian is taken under consideration.

2.10 Transition states

A transition state refers to the lowest energy maximum state, which connects to two local

minima; it should be a first order saddle point with one maximum along one degrees of

freedom (i.e. the reaction coordinate), and a minimum energy along all other degree of

freedom. Experimentally, transition states are not visible and cannot be characterized as they

have a very short lifetime. However, the activation energy (the energy between the transition

state and the reactant) connects to the rate constant for the chemical reaction. Transition state

Page 78: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

78

calculations are done basically by first running a geometry scan between the reactant and

product geometries, subsequently, the geometry scan with the highest energy is taken and a

full transition state search is done; following with a frequency calculation will characterize

the transition state as a first order saddle point with one imaginary frequency for the correct

mode. It is done by fixing specific intervals while at the same time the rest of the structure

will be fully optimized

Figure 2.1: A Potential Energy Surface (PES) indicating a reaction from reactant to product

that passes through a transition state (TS).

2.11 Computational Software

All calculations implemented within this thesis is performed using electronic structure

method software packages such as Gaussian03 (Frisch, M. et al 2003), Gaussian09 (Frisch,

M. et al 2009), Jaguar 7.6 (Schrödinger LLC, 2007), Orca (Neese, F, 2009), Turbomole

(Turbomole V6.2 2010) and the Chemishell Interface (Sherwood, P. et al 2003) series of

Programs and used without modifications.

Page 79: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

79

CHAPTER 3

PROJECT ONE

Page 80: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

80

Spin state ordering in -hydroxo bridged

diiron(III)bisporphyrin complexes1

ABSTRACT

We provided a detailed computational studies on a hydroxo-bridged diiron(III)-bisporphyrin

reported by Sankar and co-workers (Bhowmick. S, Ghosh, S.K, et al, 2012) who reported the

synthesis, structure and spectroscopic characterization of 1,2-bis[-hydroxo iron(III) 5-

(2,3,7,8,12,13,17,18-octaethylporphyrinyl)]ethane with I3–, BF4

– and ClO4

– counter anions.

Our computational modeling shows that as I3–, BF4

– and ClO4

– counter ion approach the

metal complex, the spin state ordering is affected dramatically. In order to understand how

the spin state ordering is affected by external perturbations, we have done a comprehensive

computational study. The calculations show that subtle environmental perturbations, such as

entropic corrections, perturb the spin state ordering and relative energies and are likely to be

the root cause of the variation in spin state ordering observed experimentally.

1

Mala A. Sainna, Debangsu Sil, Dipankar Sahoo, Bodo Martin, Sankar Prasad Rath, Peter Comba and Sam P. de Visser, “Spin state

ordering in μ-hydroxo bridged diiron(III)bisporphyrin complexes”. Inorganic Chemistry, 2015,54(4)1919-1930

Page 81: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

81

3.1 Introduction

Spin state ordering is an important feature in transition metal catalyzed reactions in biology

as well as in synthetic models, where the spin state of the reactant often determines the

catalytic efficiency of the system. Nature often uses transition metals as catalytic reaction

centers in enzymes, and metal centers are also used as electron transfer complexes. In many

of these systems the active center has two (or more) interacting metal ions which often are

iron centers. Examples of diiron containing enzymes and proteins include hemerythrin

(Wilkins, P.; Wilkins, R. 1987, Stenkamp, R. 1994) , ribonucleotide reductase (RNR)

(Nordlund, P.; Reichard, P. 2006, Tomter, A. B.; Zoppellaro, G.; 2013, Krebs, C.; Dassama,

L. et al 2013), ferritin,( Moënne-Loccoz, P.; Krebs, C. et al 1999) methane monoxygenase

(MMO), (Murray, L., Lippard, S. 2007, Balasubramanian, R.; Rosenzweig, A. 2007) and 9-

stearoyl-acyl carrier protein desaturase (Shanklin, J.; Somerville, C. 1991,Fox, B., Lyle, K.

2004), to name a couple of well-studied systems. These diiron enzymes have diverse

functions in biology and all have the two iron ions separated by about 3 – 5 Å and linked to

the protein via nonheme interactions with histidine and carboxylate (Glu/Asp) residues.

Hemerythrin and ferritin are proteins involved in oxygen transport and iron storage, whereas

other diiron enzymes, such as RNR, catalyze the biosynthesis of deoxyribonucleotides from

ribonucleotides. MMO is one of the most efficient biochemical oxidants in nature and one of

the few enzymes that can hydroxylate methane.

By contrast to these nonheme diiron proteins, cytochrome c oxidase (CcO) has a central

dimetal complex, where one of the metals binds to a heme group and the other is in a

nonheme environment. CcO is a trans-membrane protein that reduces molecular oxygen to

water and thereby releases four electrons that are shuttled through the membrane (Ferguson-

Page 82: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

82

Miller, S.; Babcock, G. 1996, Collman, J., Boulatov, R. 2004, Belevich, I.; Verkhovsky, M.

2006, Siegbahn, P.; Blomberg, M. 2008). To enable oxygen reduction it has a specific active

site structure with two metal centers in close proximity. The intricate details of dioxygen

binding and reduction as well as the electronic and spin state coupling between the two metal

centers in dimetal proteins, such as CcO, are still shrouded in mystery. In order to gain

insight into spin state interactions between metal centers, synthetic analogues have been

created that model the enzymatic systems (Collman, J, Boulatov, R. et al 2004, Chufán, E,

Puiu, S. et al 2007). These studies have given insight into the electronic and spectroscopic

properties of short-lived intermediates in the catalytic cycle of CcO and other dimetal enzyme

centers.

In several diiron enzymes, such as MMO and RNR, catalytic intermediates have been

proposed with either an oxo, peroxo or hydroxo group bridging the two iron centers,

frequently in conjunction with a bridging carboxylate group (Rosenzweig, A.; Lippard, S.

1994, Dassama, L.; Silakov, A. et al 2013, Korendovych, I.; Kryatov, S. et al 2007). The spin

multiplicity on each of the metal centers has an effect on the overall reaction mechanism and

may vary during the catalytic cycle. Thus, in many metal-catalyzed reactions a multistate

reactivity pattern has been established on low lying electronic and spin states (Shaik, S.;

Filatov, M. et al 1998, Shaik, S.; de Visser, S. et al 2002). For instance, computational

modeling established a degenerate pair of doublet and quartet spin states for the active

oxidant of cytochrome P450, i.e. compound I (Green, M. 1999, de Visser, S.; Kumar, D.

2011). It was found that in each of these spin states the oxidant can react with a substrate to

form products with different rate constants and sometimes even different mechanisms.

Page 83: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

83

Understanding the factors that contribute to spin state ordering and relative energies in

transition metal complexes, therefore, is important and may affect reactivity patterns.

Oxo- and hydroxo-bridged diiron active centers are common structural motifs found among

proteins involved in O2 metabolism. The transformation of an oxo to a hydroxo bridge is a

proposed step in the reaction pathways of a great variety of iron and copper redox enzymes.

As enzymatic systems are difficult to study experimentally, synthetic model complexes have

been developed that contain the characteristic features of enzyme active sites. Recently,

Sankar and co-workers reported the spectroscopic characterization of a series of 1,2-bis[-

hydroxo iron(III) 5-(2,3,7,8,12,13,17,18-octaethylporphyrinyl)]ethane complexes, 1b•X

(Scheme 3.1; X = I3–, BF4

– and ClO4

–), in which the two iron(III) centers are either equivalent

or nonequivalent based on the counter anions present, although both cores have identical

molecular structures (Bhowmik, S.; Ghosh, S. et al 2012, Ghosh, S.; Rath, S. 2010). All

complexes were crystallographically characterized and showed considerable doming of both

porphyrin cores and out-of-plane displacement of the metal ions. Also, the spin states were

found to be dependent on the counter ions used. In particular, spectroscopic investigation of

1b•I3 identified core I (shown in red in Scheme 3.1) as having nearly high-spin configuration

(S = 5/2 with a minor contribution of S =

3/2), whereas core II (shown in blue in Scheme 3.1)

had an admixed intermediate-spin state (S = 3/2 with a minor contribution of S =

5/2) both in

the solid and solution phases (Bhowmik, S.; Ghosh, S. et al 2012, Ghosh, S.; Rath, S. 2010).

Complex 1b•BF4 was, however, found to have two iron(III) centers with admixed

intermediate states, whereas 1b•ClO4 was characterized to have high-spin (S = 5/2) iron(III)

centers in the solid but nearly intermediate (S = 3/2) spin in solution. It appears, therefore, that

changing the counter ion affects the spin state ordering and spin coupling between the two

Page 84: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

84

iron centers. Dependence of the spin states of the metal complexes on counter anions has

been found rarely, particularly with anions which are far away and not apparently involved in

a direct interaction with the metal site (Bhowmik, S.; Ghosh, S. et al 2012, Ghosh, S.; Rath,

S. 2010, Ghosh, S.; Bhowmik, S. 2012, Kitchen, J.; White, N. et al 2011, Kitchen, J.;

Brooker, S. 2008, Nihei, M.; Shiga, T. et al 2007, Yamada, M.; Hagiwara, H. et al 2006,

Gütlich, P.; Hauser, A. et al 1994). How counter anions affect the spin state ordering and

coupling of metal centers is currently unknown and encouraged us to do a further

computational study.

In the present work, the synthesis, structure and properties of the corresponding μ-hydroxo

diiron(III)bisporphyrin complexes are reported here and compared with similar complexes

reported before. The ethyl linker, that bridges two porphyrin rings, introduces significant

horizontal and vertical flexibility to the bisporphyrin framework and also brings the two

porphyrin moieties close enough to distort the macrocycles significantly in the -hydroxo

complexes.

Scheme 3.1: Structures investigated in this work.

N N

N N

R R

R R

R

R

R

R Fe III

N N

N N

R R

R R

R R

R R

Fe III

O

N N

N N

R R

R R

R

R

R

R Fe III

N N

N N

R R

R R

R

R R

R Fe III

O H X OH

H X

1 X .

X : I 3 , BF 4 , ClO 4

1a : R = H

1b : R = C 2 H 5

Page 85: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

85

3.2 Methods

The work described here is focused on the computational description of -hydroxo

diiron(III)bisporphyrin 1•X with particular focus on the spin state ordering and relative

energies. Due to the size of the system, some of the calculations were done on the full model

(with ethyl side chains on both porphyrin rings), 1b, whereas other calculations use an

abbreviated model with all side chains replaced by hydrogen atoms, 1a. The initial work was

done on the crystal structure coordinates of the -hydroxo-diiron(III)bisporphyrin 1b•I3 with

the counter anion removed, Scheme 3.1, with a total of 176 atoms and overall charge +1.

Data was benchmarked against the data from ref (Bhowmik, S.; Ghosh, S. et al 2012, Ghosh,

S.; K. Rath, S. 2010), using a range of density functional theory methods and basis sets and

initially focused on the spin state ordering and energies.

Technically, a bridged -hydroxo diiron(III)bisporphyrin can exist in three possible spin

configurations on each metal center that are either ferromagnetically or antiferromagnetically

coupled. As such, there are a large number of possible spin-spin interactions between the

metal centers of which we took a representative sample. Experimental studies on 1b•I3

indicate a spin multiplicity of S = 5/2 on the iron center of core I and a gross multiplicity of

3/2

on the iron center of core II (Bhowmik, S.; Ghosh, S. et al 2012, Ghosh, S.; Rath, S. 2010). In

a first set of calculations we took the crystal structure coordinates of 1b•I3. As we were

interested in the free energy differences of the various spin states we attempted a frequency

calculation on this structure. However, as the coordinates were taken from the crystal

structure, this had a large number of imaginary frequencies. We, therefore, ran a subsequent

constrained geometry optimization, whereby the coordinates of all ethyl side-chains were

Page 86: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

86

minimized and the porphyrin cores kept fixed. This gave a frequency calculation with few

imaginary frequencies.

Structure 1b•I3 was calculated (without I3– counterion) with an overall charge of +1 and

total spin multiplicities of Stot (core I, core II) = 5 (5/2,

5/2), 4 (

5/2,

3/2), 3 (

3/2,

3/2) and 1 (½, ½).

All calculations were performed using the GAUSSIAN-09 suits of programs (Frisch, M.

2004) exploiting unrestricted density functional theory (DFT) methods with a variety of

hybrid and non-hybrid exchange-correlation functionals: B3LYP (Becke, A. 1993, Lee, C.;

Yang, W. et al 1988), B3LYP* (i.e. B3LYP with 15% HF exchange) (Reiher, M.; Salomon,

O. et al 2001), BP86 (Becke, A. 1988, Perdew, J. 1986), OLYP (Handy, N. C.; Cohen, A.

2001, Hoe, W.; Cohen, A. 2001), OPBE (Perdew, J.; Burke, K. 1996), and TPSSh (Tao, J.;

Perdew, J. et al 2003).

We tested three different basis set combinations labeled as BS1, BS2 and BS3. All initial

studies were done with a modest double- quality basis set on iron that contains an effective

core potential (LANL2DZ), in combination with a 6-31G basis set on the remaining atoms:

BS1 (Hay, P.; Wadt, W. 1985, Hehre, W.; Ditchfield, K. et al 1972). This basis set was

chosen to reduce computational cost as well as used as a comparison with a mixed valence

basis set of triple- quality consisting of DEF2-TZVP (Schaefer, A.; Huber, C. et al 1994) on

iron and the four nitrogen ligands in the first coordination sphere of each of the porphyrin

rings, whereas the remaining atoms were calculated using a 6-31G basis set: BS2. Finally, the

atoms of the complete complex 1b were also calculated with the triple- quality basis set

DEF2-TZVP throughout: BS3. The system was calculated with each of the basis sets BS1,

BS2 and BS3 coupled to all the density functional methods outlined above and the energetics

were tabulated for comparison.

Page 87: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

87

In a final set of calculations, we ran several full geometry optimizations whereby the

counter anion was included. Calculations were performed on 1b•X and 1a•X, with X = I3–,

BF4– or ClO4

–, whereby all degrees of freedom were minimized. These calculations were run

in Jaguar (Schrödinger LLC. 2010) at the B3LYP/BS1 level of theory.

To understand the differences in spin state ordering between the individual porphyrin cores,

we also split structure 1b•I3 into individual cores, whereby each monomer was capped with a

hydrogen atom at the trimmed ethyl bridge and each of the separate cores were calculated

using the same BS1, BS2 and BS3 basis sets together with the chosen density functional

method. In these monomer units, we tested OH– and H2O as the distal ligand mimicking the

bridged -hydroxo group in structure 1b•I3. The two separated mono-porphyrins were also

capped with methoxy (CH3O–) or methanol (CH3OH) in place of OH

– and H2O, respectively,

because they may be geometrically closer to the actual conformation in 1b•I3.

3.3 Results and Discussion

To try to understand the origin of the spin state ordering experimentally characterized

hydro-bridged diiron porphyrin complexes, a series of computational studies on 1•X with

several counter anions were performed. Each core has quartet and sextet spin states that are

very close in energy, whereby the relative ordering is dependent on the environmental

perturbations as well as the nature of ligands and substituents. This has analogy to enzymatic

systems, and, for instance, in the resting state of cytochrome P450, i.e. the water bound

iron(III) heme cysteinate complex, the ground state is a doublet spin state (Thomann, H.;

Bernardo, M. et al 1995), but upon release of the water molecule the spin state ordering

changes and the sextet spin state becomes the ground state (Auclair, K.; Moënne-Loccoz, P.

Page 88: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

88

et al 2001). In synthetic model complexes, such as cyanide ligated

iron(II)tetraphenylporphyrin, a rise of temperature leads to a spin crossover from low-spin to

high-spin (Li, J.; Lord, R. et al 2008). For [(py)2FeIII

(OEPO)] (OEPO = octaethyloxophlorin,

py = pyridine), the iron spin state is observed to be high-spin in solid which, however,

changes to a low-spin when dissolved in pyridine (Rath, S.; Olmstead, M. et al 2006).

Moreover, many cases have been reported on rate enhancements of oxygen atom transfer

reactions by axially ligated metal-porphyrin complexes (de Visser, S.; Ogliaro, F. et al 2002,

Nam, W.; Ryu, Y. et al 2004, Kumar, D.; de Visser, S. et al 2005, Song, W.; Ryu, Y. et al

2005, de Visser, S. 2006, de Visser, S.; Tahsini, L. et al 2009, Hessenauer-Ilicheva, N.;

Franke, A. et al 2009, Prokop, K.; de Visser, S. et al 2010, Takahashi, A.; Yamaki, D. et al

2012, Kumar, D.; Sastry, G. et al 2011).

In structure 1b•I3 there are two iron(III) centers, and both of these have three possible spin

states, which could give rise to 36 possibilities of coupling the spin states on core I with core

II. For instance, a spin of S = +5/2 on core I can couple with a spin of +

5/2, +

3/2, +

1/2, –

1/2, –

3/2

and –5/2 on core II. We did not test all these combinations, but decided to focus on the

ferromagnetic coupling between spin states with combinations (5/2,

5/2), (

5/2,

3/2), (

3/2,

3/2) and

(½, ½) only, which should give some idea on the relative stabilities of the resulting spin

states.

Figure 3.8 presents the spin state energies of the S = (5/2,

5/2), (

5/2,

3/2) and (

3/2,

3/2) states of

structure 1b as calculated with six different density functional methods. The S = (½, ½) states

were also calculated but found to be considerably higher in energy. We find dramatic

differences in spin state ordering and relative energies between the various DFT methods.

Thus, the experimental ground state of S = (5/2,

3/2) is only reproduced with the OPBE

Page 89: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

89

method, although with both B3LYP and OLYP the S = (5/2,

3/2) and S = (

3/2,

3/2) states are

close in energy within 1.0 kcal mol–1

. Therefore, within the error of the calculations, the

B3LYP and OLYP results can be considered a good match with experiment. The BP86 and

B3LYP* methods, by contrast give large deviations for the experimentally predicted spin

state ordering and seem to be less suitable to predict the spin state ordering of -hydroxo

diiron(III)bisporphyrin complexes. To test whether the basis set has an influence on the spin

state ordering and relative energies, we ran single point calculations using a triple- quality

basis set on all atoms for structure 1b: OPBE/BS3. However, this calculation with a larger

basis set only gave minor changes in relative energies and did not change the ordering with

respect to OPBE/BS2.

Figure 3.1: Relative spin state energies of (5/2,

5/2), (

5/2,

3/2) and (

3/2,

3/2) states of structure 1b

as calculated with various DFT/BS2 methods in Gaussian. Calculations done with basis set

BS3 labeled with superscript a. (a) E values relative to the S = (5/2,

5/2) state. (b) E+ZPE

values relative to the S = (5/2,

5/2) state.

Page 90: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

90

The results in Figure 3.8 show that changes in the density functional method affect the spin

state ordering and relative energies dramatically. In order to compare our work with

experimental data, zero point energies, thermal corrections, entropic and solvent corrections

were also included and each of these factors may have a different effect on the spin state

ordering and relative energies. Importantly, if a frequency calculation is run on crystal

structure coordinates of 1b, there are a large number of imaginary frequencies, mostly

connected to vibrations in the ethyl substituents. Therefore, a partial geometry optimization

with fixed cores and flexible ethyl side chains (distances, angles and dihedral angles

containing the C and H atoms of the ethyl groups) at the B3LYP/BS1 level of theory

preceded the frequency calculation. These structures with S = (5/2,

5/2), S = (

5/2,

3/2) and S =

(3/2,

3/2) spin states had no imaginary frequencies. ZPE corrections stabilize the S = (

5/2,

3/2)

state over the maximum spin state by 0.73 kcal mol–1

, whereas the S = (3/2,

3/2) state is

destabilized by 3.2 kcal mol–1

. Consequently, addition of ZPE corrections to the energies of

the data reported in Figure 3.8 improves the comparison between experiment and theory for

most methods, and now the correct spin state ordering is reproduced with B3LYP, OLYP,

TPSSh and OPBE. Despite the fact that the energy gap between the S = (5/2,

3/2) and the S =

(3/2,

3/2) states has narrowed considerably after addition of ZPE corrections, the BP86 and

B3LYP* methods still predict the wrong ground state.

We then investigated the effect of the zero-point energy, thermal and entropic corrections

on the spin state ordering and associated relative energies. It is found that free energy lowers

the S = (5/2,

3/2) state over the S = (

5/2,

5/2) state with respect to the E values by 1.4 kcal mol

1, while the S = (

3/2,

3/2) state is stabilized over the S = (

5/2,

5/2) state by 0.4 kcal mol

–1. This

Page 91: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

91

implies that at a free energy level, the experimental spin state ordering is reproduced using

the B3LYP, OLYP and OPBE methods only.

To understand the origin of the large variation in spin state ordering and relative energies

between the various DFT methods we decided to split structure 1b into two individual cores

and calculate those separately. Thus, we took structure 1b and separated it into two cores,

whereby the bridging ethyl group was included in both cores and capped with a hydrogen

atom. We initially investigated both cores with a hydroxo ligand in the distal position and

Figure 3.9 displays the spin state ordering and relative energies of core I (panel a) and core II

(panel b), respectively. Thus, all DFT methods predict an S = 3/2 ground state for core II, but

the spin state ordering for core I varies with the DFT methods. Moreover, the quartet-sextet

energy splitting varies by more than 13 kcal mol–1

for each individual core, whereby the

BP86 method gives a large stabilization of the quartet spin state over the sextet spin state.

Hence, the over-stabilization of lower spin states by pure density functionals such as BP86,

therefore, makes these methods unsuitable for the correct description of -hydroxo

diiron(III)bisporphyrin complexes. Also the reduced amount of HF exchange in the B3LYP*

method over-stabilizes the quartet spin states by a too large amount and therefore gives the

wrong spin state ordering of the complex. The B3LYP, OLYP, TPSSh and OPBE methods all

predict the correct spin state ordering for each of the individual cores and give a small

quartet/sextet spin state energy gap.

Page 92: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

92

Figure 3.2: Quartet/sextet spin state energies for (A) individual core I and (B) core II. All

energies are relative to the quartet spin state in kcal mol–1

. A negative value implies an S = 5/2

ground state.

A comparison of the quartet-sextet energy splitting in core I as compared to the energy

difference in 1b between the S = (5/2,

3/2) and S = (

3/2,

3/2) states is a measure for the

intramolecular interactions of core II on core I. In all cases, the dinuclear complex shows

extra stabilization of the S = 3/2 state on both core I and core II. On core I the quartet spin

state of the isolated core with respect to the dinuclear complex is stabilized over the sextet

spin state by 1.8 – 3.7 kcal mol–1

for the seven methods described in Figures 3.8 and 3.9,

whereas a stabilization of 2.1 – 6.2 kcal mol–1

is found for core II. This means that the

antiferromagnetic interaction between the two cores, the porphyrin ring deformation and -

stacking interactions lead to a stabilization of the quartet spin state on each core with respect

to the sextet spin state, which is independent of the density functional method and basis set.

Ring-deformation is known to stabilize intermediate spin states over high-spin states in

iron(III)-porphyrin complexes (Bhowmik, S.; Ghosh, S. et al 2012, Ghosh, S.; Rath, S. et al

2010, Ghosh, S.; Bhowmik, S. et al 2013). Indeed, in our particular system both porphyrin

Page 93: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

93

rings show considerable ruffling, but not by the same degree, hence leading to different spin

states on each iron center.

To test the effect of ZPE, thermal and entropic corrections on the spin state ordering of the

two individual cores, we ran a partial geometry optimization at B3LYP/BS1, where we fixed

the first and second coordination sphere of atoms around the iron center and reoptimized the

rest. This led to a structure without imaginary frequencies and almost negligible change in

ZPE for both spin states: The quartet spin state is stabilized over the sextet spin state by 0.2

and 0.1 kcal mol–1

for core I and II, respectively. Subsequently, we investigated the free

energy differences of these structures in the relevant spin states, and when the full

combination of ZPE, thermal and entropic corrections are included the sextet spin state is

favored by 0.9 and 0.6 kcal mol–1

for core I and core II, respectively. This is not surprising as

the entropy component contains a factor –RT ln (2S +1), which has a value of 4.4 kcal mol–1

for the sextet spin state and 3.4 kcal mol–1

for the quartet spin state at a temperature of 298 K.

As our spin state splittings are small, this extra stabilization of the free energy due to electron

spin may tip the balance in favor of the high-spin state.

We did a further set of test calculations on the individual cores I and II, where we replaced

the OH ligand by a water molecule. However, with all methods (except B3LYP*) a quartet

spin ground state was found for both core I and core II. This again shows that subtle changes

to the coordination system of iron(III)porphyrins can lead to significant changes in spin state

ordering and relative energies. To find out whether the difference between an OH ligand

versus a water ligand is an effect of the charge on the metal center, we also calculated the

same complexes with CH3O– and CH3OH bound in the distal ligand position. At OPBE/BS2

level of theory with a CH3O– ligand, we find core I to have a sextet spin ground state by 4.6

Page 94: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

94

kcal mol–1

, whereas core II has a quartet spin ground state by 2.0 kcal mol–1

. This spin state

ordering match those found for OH– as well as those observed experimentally. By contrast,

when we replace methoxide by methanol and redo these calculations at the same level of

theory, we find core I to have a quartet spin ground state by 2.6 kcal mol–1

, and core II has a

sextet spin ground state by 16.9 kcal mol–1

. In summary, binding of an anionic ligand to an

iron(III) center affects the spin state ordering with respect to binding a neutral (solvent)

molecule to an iron(III) center.

To further establish the environmental effects on the spin state ordering and relative

energies, we decided to include counter anions, i.e. I3–, BF4

– and ClO4

–, into the model. We

took the crystal structure coordinates of 1b•X with X = I3–, BF4

– and ClO4

– and did a full

geometry optimization without constraints for the three relevant spin states, i.e. S = (5/2,

5/2), S

= (5/2,

3/2) and S = (

3/2,

3/2), for 1a•X and 1b•X. The obtained structures of the complexes are

shown in Figure 3.10. Counter ions have small and virtually insignificant effects on the

optimized geometries and, as follows from Figure 3.10, we find very similar structures for

1a•X (X = I3–, BF4

– or ClO4

–). The same trends are found for the large system with ethyl

substituents attached to the porphyrin scaffold (1b•X). Of course, there are geometric

differences between the spin states as discussed above. A geometry optimization elongates

both Fe–O distances from 1.897/1.934 Å in the crystal structure to slightly over 2.0 Å. On the

other hand, very little changes are found for the porphyrin groups: the average Fe–N

distances in core I/core II are 2.051/2.006 Å in the crystal structure, which are values close to

those seen in Figure 3.10. Also the displacement of the metal from the plane through the four

nitrogen atoms (defined as ) is virtually the same in the optimized geometries as compared

to the crystal structures.

Page 95: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

95

Figure 3.3: Optimized geometries of the S = (5/2,

5/2), [S = (

5/2,

3/2)] and {S = (

3/2,

3/2)} states

of 1a•X and 1b•X with X = I3–, BF4

– or ClO4

– at the B3LYP/BS1 level of theory in Jaguar.

Bond lengths are given in angstroms and is the average deviation of the iron from the plane

through the four nitrogen atoms.

Table 3.1 provides relative energies of the S = (5/2,

5/2), S = (

5/2,

3/2) and S = (

3/2,

3/2)

optimized geometries of 1a•X and 1b•X (X = I3–, BF4

– or ClO4

–) as calculated with B3LYP

and OPBE. Energetically, in all cases the S = (3/2,

3/2) state is the ground state followed by the

S = (5/2,

3/2) state and the S = (

5/2,

5/2) state. Only very small changes in energies are obtained

when counter ions are added and no clear changes are observed, based on the nature of the

counter ions. The same trends are found for B3LYP and OPBE as density functional method.

Clearly, in the gas phase the effect of counter ions is small and does not induce changes in the

spin state ordering and relative energies. As discussed above, environmental effects can

rFeO = 2.013 [2.080] {2.050}

rFeO = 2.004 [1.978] {2.041}

rFeN,average = 2.071 [1.992] {1.998} = 0.456 [0.274] {0.300}

rFeN,average = 2.073 [2.080] {2.004} = 0.433 [0.458] {0.287}

rFeO = 2.020 [1.992] {2.006}

rFeO = 2.018 [2.084] {1.965}

rFeN,average = 2.074 [2.083] {2.073} = 0.453 [0.482] {0.447}

rFeN,average = 2.075 [2.000] {2.001} = 0.446 [0.267] {0.256}

rFeO = 2.006 [2.072] {2.042}

rFeO = 2.014 [1.988] {2.052}

rFeN,average = 2.073 [1.990] {1.999} = 0.441 [0.264] {0.291}

rFeN,average = 2.073 [2.076] {1.996} = 0.455 [0.470] {0.296}

rFeO = 2.018 [1.988] {2.056}rFeO = 2.014 [2.086] {2.052}

rFeN,average = 2.086 [2.088] {2.011} = 0.450 [0.490] {0.323}

rFeN,average = 2.083 [2.003] {2.013} = 0.458 [0.284] {0.309}

rFeO = 2.001 [2.079] {1.941}rFeO = 2.001 [1.975] {1.996}

rFeN,average = 2.068 [1.975] {1.976} = 0.479 [0.285] {0.275}

rFeN,average = 2.069 [2.078] {2.073} = 0.425 [0.469] {0.444}

rFeO = 2.005 [1.981] {2.055}

rFeO = 1.996 [2.065] {2.041}

rFeN,average = 2.065 [2.072] {1.984} = 0.470 [0.492] {0.314}

rFeN,average = 2.065 [1.982] {1.990} = 0.424 [0.270] {0.299}

Page 96: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

96

change the relative spin state ordering and in some cases a different electronic ground state is

observed experimentally in solution than in a crystal structure.

Table 3.1. Relative energies of optimized geometries in different spin states of 1a•X and

1b•X complexes.a

a Energies in kcal mol

–1 with ZPE corrected values in parenthesis.

E (E+ZPE) E (E+ZPE) E (E+ZPE) E (E+ZPE) E (E+ZPE)

1aX X = I3– X = BF4

– X = ClO4

B3LYP/BS1 B3LYP/BS1 OPBE/BS1 B3LYP/BS1 OPBE/BS1

S = (5/2,

5/2) 0.00 (0.00) 0.00 (0.00) 0.00 (0.00)

S = (5/2,

3/2) –6.07 (–5.23) –4.99 (–3.88) –5.83 (–4.70)

S = (3/2,

3/2) –9.87 (–7.79) –9.83 (–7.78)

1bX

S = (5/2,

5/2) 0.00 (0.00) 0.00 (0.00) 0.00 (0.00) 0.00 (0.00) 0.00 (0.00)

S = (5/2,

3/2) –5.90 (–4.84) –6.11 (–5.16) –2.83 (–1.66) –5.69 (–4.75) –3.00 (–1.61)

S = (3/2,

3/2) –11.08 (–8.74) –4.62 (–2.29) –9.12 (–7.09) –4.29 (–1.53)

Page 97: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

97

Interestingly, the relative energies in Table 3.3 are similar to those reported above in Figure

3.10a, where we reported single point DFT calculations on the crystal structure coordinates.

However, correction for ZPE by a partial geometry optimization lowered the S = (5/2,

3/2)

below the S = (3/2,

3/2) state, which is not seen when a full geometry optimization and

frequency calculation is performed. As such, minor geometric distortions to the structure

affect the spin state ordering and electronic ground state of -hydroxo-bridged diiron

complexes. We should note here that the relative energies can be further affected by spin-

orbital coupling also (Gupta, A.; Gupta, G. et al 1994, Hagen, W.; van den Berg, W. et al

1998). It may very well be that the degree of spin-orbit coupling is affected by nearby anions,

such as counter anions, and, thereby changes the spin state ordering as observed

experimentally.

The data shown in Table 3.3 indicate that the energetic differences between the three main

spin states are very small and that in the gas phase, there is a preference for an S = (3/2,

3/2)

spin state. As reasoned above, entropic corrections will favor higher spin states and

consequently, an increase of the temperature will strongly influence the spin state ordering.

We note, however, that in the optimized geometries, the counter anion is aligned with the

dipole moment vector of the diiron porphyrin complex as expected for an ion-dipole

interaction. The change in spin state upon addition of counter ions observed experimentally,

may, therefore, result from a combination of factors, of which some are not included in the

computational models. The hydrogen bonding interactions of solvent molecules to the

bridging -hydroxo group may play a key role in establishing the electronic ground state, and

this was not considered in the models used here. Furthermore, an induced electric field was

shown to strongly affect the spin state ordering, which may have an effect in the systems

Page 98: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

98

studied here (Shaik, S.; de Visser, S. et al 2004, de Visser, S. 2005). Finally, spin-orbit

coupling can influence the small energy differences and change the relative ordering between

the various spin states.

It would be interesting to compare the spin states of µ-hydroxo diiron(III)bisporphyrins

obtained out of experiment and theory. Experimentally, 1b•I3 is assigned to have nearly high-

spin and admixed-intermediate spins of iron in cores I and II, respectively, which is also

reproduced in theory using the crystal structure coordinates of the molecule utilizing a

variety of hybrid and non-hybrid exchange correlation functions. Energetically, the S = (3/2,

3/2) state is the ground state followed by the S = (

5/2,

3/2) state and then the S = (

5/2,

5/2) state

which happens to be the correct spin state ordering obtained experimentally for most of the

complexes described here. It is interesting to note here that two Fe centers in 1b•ClO4 which

are found to be high-spin (S = 5/2) in the solid state converts to the intermediate-spin (S =

3/2)

in solution. One probable reason could be the breaking of H-bonding interactions observed in

the solid once it dissolved in solution. Thus, subtle change in the environment can also affect

the relative spin state ordering and in some cases a different electronic ground state. The

relative energies of the spin states are also known to be affected by the extent of spin-orbital

coupling which can be influenced by the nearby counter-anions. Moreover, an induced

electric field can also affect spin state ordering and relative energies therein. Thus, the

different spin-state behavior of 1b•X, as observed in the experiment, can be attributed to the

properties of the counter anion X which are also known to operate under the influence of

variety of phenomena related to steric effects, charge polarization, stability of the ion-pair

formations etc. and were not considered in the computational models used here.

Page 99: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

99

3.4 Conclusion

I have provided a detailed computational studies on a diiron(III)bisporphyrin complex with

particularly focus on the spin state ordering. An earlier experimental work by Sankar and co-

workers shows a strong variation in electronic ground state for 1b•X (X = I3–, BF4

–, ClO4

–)

(Bhowmick. S.; Ghosh. S. et al, 2012). Crystallographic studies show considerable ruffling of

both porphyrin rings by unequal amounts, and the doming and structural features are

characteristic of a spin state. 1H NMR and Mössbauer studies confirmed this assignment.

Finally, an extensive computational study was performed on various diiron(III)bisporphyrin

complexes with a range of DFT methods, which established a variety of factors that affect

spin state ordering and relative energies in -hydroxo bridged diiron(III)bisporphyrin

complexes. The latter highlights the small energy gap between the S = (3/2,

3/2) and S = (

5/2,

3/2) spin states and how their ordering can interchange through external perturbations.

Page 100: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

100

CHAPTER 4

PROJECT TWO

Page 101: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

101

A comprehensive test set of epoxidation rate

constants by iron(IV)-oxo porphyrin complexes2

ABSTRACT

Cytochrome P450 enzymes are heme based monoxygenases that catalyse a range of oxygen

atom transfer reactions with various substrates, including aliphatic and aromatic

hydroxylation as well as epoxidation reactions. The active species is short-lived and difficult

to trap and characterize experimentally, moreover, it reacts in a regioselective manner with

substrates leading to aliphatic hydroxylation and epoxidation products, but the origin of this

regioselectivity is poorly understood. Crestoni and co-worker’s synthesized a model complex

and studied it with low-pressure Fourier transform-ion cyclotron resonance (FT-ICR) mass

spectrometry (MS). A novel approach was devised using the reaction of [FeIII

(TPFPP)]+

(TPFPP = meso-tetrakis(pentafluorophenyl)porphinato dianion) with iodosylbenzene as a

terminal oxidant which leads to the production of ions corresponding to [FeIV

(O)(TPFPP+•

)]+.

This species was isolated in the gas-phase and studied in its reactivity with a variety of

olefins. Product patterns and rate constants under Ideal Gas conditions were determined by

FT-ICR MS. All substrates react with [FeIV

(O)(TPFPP+•

)]+ by a more or less efficient oxygen

atom transfer process. In addition, substrates with low ionization energies react by a charge-

transfer channel, which enabled us to determine the electron affinity of [FeIV

(O)(TPFPP+•

)]+

2 Mala A. Sainna, Suresh Kumar, Devesh Kumar, Simonetta Fornarini, Maria Elisa Crestoni and Sam P. de Visser, “A comprehensive test

set of epoxidation rate constants by iron(IV)-oxo porphyrin complexes”. Chemical Science, 2015, 6, 1516-1529.

Page 102: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

102

for the first time. Interestingly, no hydrogen atom abstraction pathways are observed for the

reaction of [FeIV

(O)(TPFPP+•

)]+ with prototypical olefins such as propene, cyclohexene and

cyclohexadiene, which suggests that the competition between epoxidation and hydroxylation

– in the gas-phase – is in favour of substrate epoxidation. This notion further implies that

P450 enzymes will need to adapt their substrate binding pocket, in order to enable favourable

aliphatic hydroxylation over double bond epoxidation pathways. The MS studies yield a large

test-set of experimental reaction rates of iron(IV)-oxo porphyrin complexes, so far

unprecedented in the gas-phase, providing a benchmark for calibration studies using

computational techniques. My computational results presented here confirm the observed

trends excellently and rationalize the reactivities within the framework of thermochemical

considerations and valence bond schemes.

4.1 Introduction

The cytochromes P450 are part of the body’s natural defence mechanism in the liver and

perform vital functions for human health that include the biodegradation of xenobiotic and

drug molecules (Sono, M.; Roach, P. et al 1996, Groves, T. 2003, Ortiz de Montellano, R.

2004). Due to this broad chemical function the P450s can bind and activate a large range of

substrates with varying shapes and sizes. Generally, the P450s act as monoxygenases,

whereby they bind and utilize molecular oxygen via a heme centre and transfer one of the

oxygen atoms of O2 to a substrate, while the second oxygen atom leaves the process as a

water molecule. The P450s react with substrates activating aliphatic and aromatic

hydroxylation, epoxidation and sulfoxidation reactions, but have also been reported to

catalyse desaturation and N-dealkylation reactions (Groves, T. 2005, Watanabe. Y.;

Page 103: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

103

Nakajima, H.; et al 2007). There are many different P450 isozymes and until early 2014

thousands of different structures had been characterized (Nelson, R. 2006). All P450s share

common features which include a catalytically active heme group with a central iron atom

that is linked to the protein by the thiolate sulphur atom of a cysteinate side chain (Poulos, L,

Finzel, C. et al 1985, Poulos, L.; Finzel, C. et al 1986). Fig 4.1 displays the structure of a

typical P450 active site, namely the one belonging to the CYP124 isozyme as taken from the

2WM4 protein databank (pdb) file (Johnston, B.; Kells, M. et al 2009). As shown in Fig 4.1

the substrate (tyramine) is located in a cleft nearby the heme, the substrate binding pocket,

which is in a tight orientation with stabilizing hydrogen bonding interactions by several

residues. The vacant sixth coordination site of iron is the position where molecular oxygen

will bind during the catalytic cycle. The process includes two reduction and two protonation

steps to synthesize the active species of P450 called Compound I (Cpd I) (Meunier, B.; de

Visser, P. et al 2004, de Visser, P.; Kumar, D. 2010). Cpd I is highly reactive and therefore

difficult to study experimentally, however, a few reports on its spectroscopic properties have

appeared in the literature (J. Rittle, J.; Green, T. 2010).

Page 104: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

104

Figure 4.1: Active site of P450 as taken from the 2WM4 pdb file.

Due to the short lifetime of Cpd I, studying catalytic mechanisms and reaction rates of P450

catalysed reactions is challenging (Auclair, K.; Hu, Z. et al 2002, Lafite, P.; André, F. et al

2007, Dowers, S.; Rock, A. et al 2004, Cooper, R.; Groves, T. 2011, Roberts, M.; Jones, P.

2010); therefore, many studies have focused on biomimetic iron-porphyrin complexes instead

(Groves, T.; Myers, S. 1983, Groves, T.; Watanabe, Y. 1986, Ostović, D.; Bruice, C. 1989,

Groves, T.; Gross, Z. 1994, Stephenson, N.; Bell, T. 2006, Hessenauer-Ilicheva, N.; Franke,

A. et al 2007, Mas-Balleste, R.; Que Jr, L. 2007, Comba, P.; Rajaraman, G. 2008, Bruijnincx,

A.; Buurmans, C. et al 2008, Hull, F.; Sauer, O. et al 2009, McGown, A.; Kerber, D. et al

2009, Company, A.; Feng, Y. et al 2009, Franke, A.; Wolak, M. et al 2009, Hessenauer-

Ilicheva, N.; Franke, A. et al 2009). These studies gave detailed insight into the effect of axial

and equatorial ligands (Gross, Z.; Nimri, S. 1994, Gross, Z. 1996, Czarnecki, K.; Nimri, S. et

al 1996), but also on the local environment such as the substrate binding pocket. A

Page 105: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

105

particularly useful method to establish the properties and reactivity patterns of short-lived

complexes, such as catalytic intermediates, is Fourier transform – ion cyclotron resonance

(FT-ICR) mass spectrometry (MS) (Marshall, G.; Hendrickson, L. et al 1996, Roithová, J.;

Schröder, D. 2010, Nibbering, M. 2006). In FT-ICR MS, the charged species of interest

(either positive or negative ion) is trapped in a collision cell for a specific time during which

reactions with neutral gases can occur and be studied at the prevailing low pressure of the

instrument. FT-ICR MS allows one to measure the ion distributions and fragmentation

patterns at varying trapping time, thereby yielding insight into reactivities, and enabling one

to calculate rate constants and thermochemical properties. In recent work, Crestoni, Fornarini

and co-workers have succeeded in trapping and characterizing the Cpd I analogues of iron

and manganese porphyrin complexes and studied their reactivity with a selection of substrates

(Crestoni, E.; Fornarini, S. 2005, Crestoni, E.; Fornarini, S. 2007, Crestoni, E.; Fornarini, S.

et al 2010, Lanucara, F.; Crestoni, E. 2011). Thus, the [MnV(O)(TPFPP)]+ complex (TPFPP

= meso-tetrakis (pentafluorophenyl)porphinato dianion) was found to react with model

substrates through oxygen atom transfer (OAT), electron transfer (ET), hydride transfer and

ligand addition. However, no direct hydrogen atom transfer (HAT) with any tested substrate

took place.

In order to find out what drives the OAT reaction of Cpd I with olefins, Crestoni and co-

workers decided to investigate the properties and reactivities of [FeIV(O)(Por+•

)]+ (Por =

porphine dianion) and [FeIV(O)(TPFPP+•

)]+ with FT-ICR MS and relate the findings with

density functional theory (DFT) methods I did in order to rationalize the trends. These studies

represent the first comprehensive computational – study benchmarked against experiment on

olefin epoxidation by iron(IV)-oxo porphyrin cation radical models and allow correlations to

Page 106: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

106

be established between the OAT rate constant and the ionization energy (IE) of the olefin.

These correlations are further supported and rationalized by computational modelling.

Scheme 4.1: Models investigated in this work.

4.2 Methods

4.2.4 COMPUTATION

All calculations discussed here utilize density functional theory (DFT) methods as

implemented in the Jaguar and Gaussian-09 program packages (Schrodinger, 2011, Frisch, J.

2004). Two different models were investigated: (i) [FeIV

(O)(Por+•

)]+ (A) that includes a

porphyrin (Por) ring with all side-chains abbreviated to hydrogen atoms, and (ii)

[FeIV

(O)(TPFPP+•

)]+ (B), Scheme 4.1. Similar to previous work of ours in the field (Latifi, R.;

Sainna, A. et al 2013, Quesne, G.; Latifi, R. et al 2014), we use the unrestricted hybrid

density functional method UB3LYP19 as it was shown to reproduce the kinetics of

metal(IV)-oxo oxidants well (Vardhaman, K.; Sastri, V. et al 2011, Vardhaman, K.; Barman,

P. et al 2013, Kumar, S.; Faponle, S. et al 2014). Initial exploratory calculations employed a

Page 107: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

107

modest LANL2DZ basis set on iron and 6-31G on the rest of the atoms (basis set BS1) (Hay,

J.; Wadt, R. 1985) for geometry optimizations, analytical frequencies and geometry scans.

These studies explored the potential energy surfaces involving reactants, intermediates and

products on different spin states in detail and generated starting structures for the transition

state optimizations. All local minima reported here had real frequencies only and the

transition states were characterized by a single imaginary frequency for the correct mode. To

improve the energetics of these structures we did single point calculations in the gas-phase

with a triple- quality basis set on iron (LACV3P+) and 6-311+G* on the rest of the atoms,

basis set BS2. Subsequently, all geometries (local minima and transition states) were

reoptimized at the UB3LYP/BS2 and UB3LYP-D3/BS2 levels of theory (Becke, D. 1993,

Lee, C.; Yang, W. 1988) (Grimme, S.; Antony, J. et al 2010) and characterized by an

analytical frequency analysis. Barrier heights reported in this work were calculated relative to

isolated reactants, although using reactant complexes instead only minor changes are

observed. The effect of solvent on the rate constants was tested through single point

calculations using the self-consistent reactant field model as implemented in Gaussian with a

dielectric constant representing chloroform ( = 4.7113).

Ionization energies and bond dissociation energy (BDEOH) values were calculated as

before (de Visser, P. 2010) and represent adiabatic values for reaction 1 and 2, respectively

and report UB3LYP/BS2//UB3LYP/BS1 energies including ZPE and dispersion corrections.

A A+•

+ e– + IEA (1)

[FeIV

(OH)(Por)]+ [Fe

IV(O)(Por

+•)]

+ + H

• + BDEOH (2)

Page 108: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

108

4.3 Results

Scheme 4.2: Substrates investigated in this work.

Page 109: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

109

a Ionization energies (IE, eV) are from ref (Lias, G.; Bartmess, E. et al) N/A stands for not

available. b Second-order rate constants (kexp) in units of 10

–10 cm

3 molecule

–1 s

–1 are

measured at a temperature of 300 K in the FT-ICR cell. The estimated error in kexp is 30%,

although the internal consistency of the data is within 10%. c Collision rate constants (kADO)

evaluated with the parameterized trajectory theory. d Reaction efficiency (%), = kexp/kADO ×

___________________________________________________________________________

Table 4.1. Kinetic data and product distributions obtained for the gas phase reaction of

[FeIV

(O)(TPFPP+•

)]+ with selected olefins as determined by FT -ICR MS.

___________________________________________________________________________

Substrate IE a kexp

b kADO

c

d HT CT OAT Add

ethene 10.51 8.5 10–5

8.5 1 10–3

– – 100 –

propene 9.73 7.6 10–3

9.45 0.080 – – 100 –

1-butene 9.55 0.029 9.6 0.30 – – 100 –

E-2-butene 9.10 0.080 10.8 0.74 – – 100 –

2,3-dimethyl-1-butene9.07 0.145 9.5 1.5 – – 100 –

cyclohexene e 8.95 0.194–0.291 9.7 2–3 – – 75 25

1,4-cyclohexadiene 8.82 0.511 9.29 5.5 – – 90 10

2-methoxy-1-propene8.64 0.819 10.5 7.8 – – 100 –

1,3-pentadiene 8.60 0.826 9.6 8.6 – – 100 –

styrene 8.46 1.40 9.26 15 – – 100 –

1,3-cyclohexadiene 8.25 1.58 9.29 17 – – 100 –

trans--methylstyrene 8.1–8.2 2.97 11.9 25 4 – 96 –

indene 8.14 3.18 8.6 37 2 12 86 –

-pinene f N/A 4.32–4.7 9.4 46–50 – – 100 –

Page 110: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

110

100. e The reaction with cyclohexene-d10 gave a rate constant within experimental error of

that for cyclohexene-h10. f The IE for -pinene is 8.07 eV.

Scheme 4.3: Pathways observed for the reaction of [FeIV

(O)(TPFPP+•

)]+ ions (R = C6F5) with

selected substrates (Sub) as studied with FT-ICR MS.

Figure 4.2: Time dependence of relative ion abundancies for the reaction of

[FeIV

(O)(TPFPP+•

)]+ (m/z 1044) with indene. Product ions are [Fe

III(TPFPP)]

+ (m/z 1028),

0

20

40

60

80

100

0 5 10 15 20 25 30 35

m/z 116

m/z 1028

m/z 1044

m/z 1160

Abundance

[%]

time [seconds]

Page 111: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

111

[Fe(TPFPP)(C9H8)O]+ (m/z 1160) and C9H8

+• (m/z 116). Experiments were performed in the

presence of indene at 5.2 10–8

mbar in the FT-ICR cell.

4.4 Theoretically derived reaction paths, energetics and structures.

The experimental studies reported above present a comprehensive test set of model reactions

of iron(IV)-oxo porphyrins with olefins for the first time and enable extensive benchmarking

and calibration of computational methods and procedures against gas-phase (Ideal Gas

conditions) rate constants. We decided to take the opportunity and calibrate previously used

methods and procedures for DFT studies on these chemical systems and compare to the

results of the FT-ICR rates from Table 4.1. In addition, the computational studies were

performed to further understand the substrate activation patterns by [FeIV

(O)(TPFPP+•

)]+ with

olefins, and rationalize the obtained trends. Before we will discuss details of the reaction

mechanism and possible reactivity trends, let us start with a detailed analysis of the reactant

species, namely [FeIV

(O)(Por+•

)]+,

4,2A, and [Fe

IV(O)(TPFPP

+•)]

+,

4,2B.

Fig 4.3 displays the high-lying occupied and low-lying virtual orbitals of 4,2

A; the orbitals

for 4,2

B look very similar. These orbitals are dominated by the interactions of the metal 3d

orbitals with its ligands and several -type porphyrin orbitals. Lowest in energy are a pair of

-type orbitals (z2 and xy): z

2 represents the -interactions of the 3dz

2 orbital on iron

with the 2pz orbital on oxygen, whereas the xy gives the interactions of the 3dxy orbital on

iron with 2px,y orbitals on the four nitrogen atoms of the porphyrin ligand. The antibonding

combinations of these two orbitals (*z2 and *xy) are high in energy and virtual. Also

doubly occupied is the x2–y

2 orbital, which is a lone-pair orbital located in the plane of the

porphyrin ring. Finally, the interaction of the metal 3dxz/3dyz with the 2px/2py on the

Page 112: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

112

oxygen atom leads to a pair of xz/yz and a pair of *xz/*yz set of orbitals. The z2, xy,

xz and yz bonding orbitals are doubly occupied and low-lying in all calculations reported

here. In addition to the metal-type orbitals there are also two porphyrin-type -orbitals that in

D4h symmetry have the labels a1u and a2u. With a thiolate as axial ligand the a2u orbital

strongly mixes with a 3pz orbital on sulphur and hence is destabilized in energy (Green, T.

1999, Ogliaro, F.; de Visser, P. et al 2001), which strongly affects the electron affinity of the

oxidant and consequently is responsible for its push-effect (Dawson, H.; Holm, H. et al

1976).

Figure 4.3: Molecular valence orbitals of 4A.

2*z

xy*

ua2ua1

xz* yz*

yzxz

2z xy 22 yx

Page 113: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

113

a Relative energies in kcal mol

–1 with respect to the

2A1u state, ND stands for not determined.

b Energies obtained at UB3LYP/BS2//UB3LYP/BS1 level of theory.

c Energies and

geometries calculated at UB3LYP/BS2 level of theory. d Energies and geometries calculated

at UB3LYP-D3/BS2 level of theory.

The set of orbitals displayed in Fig 4.3 is occupied with 15 electrons and as several of these

orbitals are close in energy there a number of possibilities to distribute the electrons over the

orbitals. In addition, states can also be found in various spin states ranging from doublet to

quartet and sextet, where we identify the spin state with a superscript in front of the electronic

state label. Thus, the electronic state labelled as 4A2u has an overall quartet spin state and

___________________________________________________________________________

Table 4. 2 . Relative energies of several low - lying electronic states of [Fe(O) (Por +•

)] + (A). a

___________________________________________________________________________

State Configuration A b A

c A d

E+ZPE E+ZPE E+ZPE

2 A 1u 2 * xz * yz

a 1u 0.00 0.00 0.00

4 A 1u 2 * xz * yz

a 1u 0.71 0.19 0.21

2 A 2u 2 * xz * yz

a 2u 1.65 3 .75 3.80

4 A 2u 2 * xz * yz

a 2u 1.25 3.42 3.47

6 A 2u * xz

* yz * xy

a 2u 9.25 18.68 19.38

4 xy * xz * yz

* xy a 1u

9.69 ND ND

4 zz * xz

* yz * zz

a 2u 19.53 ND ND

Page 114: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

114

singly occupied a2u molecular orbital with overall electronic configuration: z22 xy

2 xz

2 yz

2

x2–y22 *xz

1 *yz

1 a1u

2 a2u

1 or in short [core] x2–y2

2 *xz

1 *yz

1 a1u

2 a2u

1. Similarly, we

calculated the doublet spin state (2A2u state), where the unpaired electron in the a2u orbital is

antiferromagnetically coupled to the unpaired electrons in the two * orbitals: 2A2u = [core]

x2–y22 *xz *yz a1u

2 a2u.

Previous studies with either imidazole, acetonitrile or thiolate as axial ligand (Green, T.

1999, Ogliaro, F.; de Visser, P. et al 2001, de Visser, P.; Shaik, S. et al 2003, Kamachi, T.;

Yoshizawa, K. 2003, Bathelt, M.; Zurek, J. et al 2005, de Visser, P. 2005, de Visser, P.;

Tahsini, L. et al 2009, Lonsdale, R.; Oláh, J. et al 2011, Isobe, H.; Yamanaka, S. et al 2011)

showed the 4,2

A2u states to be close in energy and well below alternative states. However, this

was due to considerable mixing of the a2u orbital with the axial ligand, which obviously is not

possible in our chemical system that lacks an axial ligand. However, in an isolated porphyrin

macrocycle, the a1u and a2u orbitals are degenerate (Ghosh, A. 1998); therefore, we decided to

investigate a range of possible electronic states for the pentacoordinated iron(IV)-oxo

porphyrin cation radical system, [FeIV

(O)(Por+•

)]+. Firstly, we tested the stability of the

4,2A2u

states and the alternative 4,2

A1u states with [core] x2–y22 *xz

1 *yz

a1u

1 a2u

2 orbital

occupation. In addition, we attempted to generate models with the iron in oxidation state

iron(V), i.e. 2xz state with occupation [core] x2–y2

2 *xz

1 a1u

2 a2u

2, or the iron in oxidation

state iron(III), i.e. the 4A state with orbital occupation [core] x2–y2

2 *xz

2 *yz

1 a1u

1 a2u

1 and

the 6xy,III state with [core] x2–y22 *xz

1 *yz

1 *xy

1 a1u

1 a2u

1 occupation. However, all our

attempts to calculate iron(III) or iron(V) states failed and converged back to lower lying

solutions with four un-paired electrons on the metal in a formal iron(IV) oxidation state,

Page 115: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

115

hence the 2xz,

4A and 6xy,III states are high in energy and inaccessible to our chemical

system.

Table 4.2 summarizes relative energies of optimized geometries of the various electronic

spin states as calculated with different DFT methods for 2,4,6

A. As follows from Table 4.2 all

calculations of A give a 2A1u ground state that is nearly degenerate with the corresponding

quartet spin state. In general, calculations done at UB3LYP/BS2 and UB3LYP-D3/BS2 give

almost identical spin state orderings and relative energies, which shows that dispersion is not

a critical component for these chemical structures. Nevertheless, the 4,2

A2u and 4,2

A1u states

are close in energy and all four states could have a finite lifetime.

Optimized geometries of the 4,2

A2u and 4,2

A1u states are given in Fig 4.4. Geometrically, no

dramatic changes in bond lengths are obtained between the three optimization techniques. A

small basis set gives slightly longer Fe–O distances than those found with a triple- basis set.

The effect of dispersion is negligible on the optimized geometries: UB3LYP/BS2 and

UB3LYP-D3/BS2 give virtually the same chemical structures. Addition of meso-substituents

to the porphyrin ring such as pentafluorophenyl groups is not expected to dramatically

change key bond lengths in the optimized geometries and relative energies of individual spin

states (Neu, M.; Quesne, G. et al 2014, Lanucara, F.; Chiavarino, B. et al 2011). Thus, recent

work of the Goldberg group showed that meso-substituted manganese-oxo porphyrinoid

complexes retained the spin state ordering and converged to a closed-shell singlet

manganese(V)-oxo state in all cases (Neu, M.; Quesne, G. et al 2014).

Page 116: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

116

Figure 4.4: Optimized geometries of the 4,2

A2u and 4,2

A1u states of 4,2

A as calculated

with UB3LYP/BS1 [UB3LYP/BS2] {UB3LYP-D3/BS2} with Fe–O bond lengths in

angstroms.

The calculations on the low-lying 4,2

A2u and 4,2

A1u states reported in Table 4.2 and Fig 4.4

show that geometrically there are very little differences between these states, but the spin

state ordering and relative energies are sensitive to the method and basis set. Recent,

complete active site (CASSCF) and restricted active site (RASSCF) calculations of Pierloot

and co-workers (Radoń, M.; Broclawic, E. et al 2011) calculated the 4,2

A2u and 4,2

A1u states of

A within 1 kcal mol–1

of each other with a small preference for the A1u states. However, they

also located two low-lying iron(V) states, which we were unable to characterize and for

which no experimental evidence exist. Unfortunately, our chemical systems (in particular

structure B) are too large to attempt calculations using the CASSCF and RASSCF methods;

therefore, we decided to continue with UB3LYP instead.Subsequently, we investigated the

substrate epoxidation by [FeIV

(O)(Por+

)]+, i.e.

4A. We find the lowest lying barriers to

proceed from the 4A2u state and will focus on those in the following. Previous studies on the

epoxidation of olefins by [FeIV

(O)(Por+

)(L)] with L = NCCH3 or Cl– showed that the same

2A1u: 1.630 [1.608] {1.608}4A1u: 1.637 [1.611] {1.608}2A2u: 1.626 [1.604] {1.604}4A2u: 1.627 [1.604] {1.604}

Page 117: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

117

trends in reactivity are observed when the Por ligand is replaced by TPFPP (D. Kumar, L.

Tahsini, et al 2009, Kumar, D.; Latifi, R. et al 2013), hence the smaller model was used in

this study. We investigated substrate epoxidation with a range of olefins: ethene (1), propene

(2), 1-butene (3), E-2-butene (4), cyclohexene (5), 1,3-cyclohexadiene (6), styrene (7), trans-

-methylstyrene (8), Z-2-butene (9) and 2-pinene (10). For all substrates we calculated the

full potential energy profile from reactants to epoxide products, but for space restrictions we

will focus on the rate determining C–O bond formation transition states (TSCO) only. All

reactions are concerted with a single C–O activation barrier leading to epoxide product

complexes PE. This is unusual as previous calculations on substrate epoxidation by Cpd I

models gave a stepwise mechanism via a radical intermediate that via a ring-closure barrier

was separated from epoxide product complexes (de Visser, P.; Ogliaro, F. et al 2001, de

Visser, P.; Ogliaro, F. et al 2001, de Visser, P.; Ogliaro, F. 2002, Kumar, D.; Karamzadeh, B.

et al 2010, Kumar, D.; de Visser, P. et al 2005). The orientation of the substrate and the

strong displacement of the metal from the porphyrin plane are the likely reason for the fact

that radical intermediates are saddlepoints here. Thus, the ring-closure barrier on the quartet

spin state surface involves an electron transfer from substrate into *z2. The latter orbital in

iron-porphyrin complexes with axial ligand, e.g. thiolate, contains a strong contribution from

axial ligand orbitals (3pz) and therefore is high in energy. Since our particular system lacks

an axial ligand, the *z2 orbital is considerably lower in energy and as a consequence the

lifetime of the radical intermediate is reduced and the reaction to form products is now

concerted.

Fig 4.5 gives the optimized geometries of the C–O activation transition states (TSCO) for

all substrates. Generally, the transition states occur early with a long C–O distance and

Page 118: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

118

relatively short Fe–O distance that has not dramatically changed from what it was in the

iron(IV)-oxo porphyrin cation radical state. As expected the metal is considerably displaced

from the plane through the four nitrogen atoms of the porphyrin ring by as much as 0.287 –

0.299 Å. These transition states bear resemblance to substrate epoxidation barriers calculated

previously for P450 Cpd I reactions with olefins (de Visser, S.P.; Ogliaro, F. et al 2001, de

Visser, S.P.; Ogliaro, F. et al 2001, de Visser, S.P.; Ogliaro, F. 2002, Kumar, D.;

Karamzadeh, B. et al 2010, Kumar, D.; de Visser, S.P. et al 2005). Electronically, all

transition states are accomplished by single electron transfer from the substrate into the a2u

orbital and the formation of an [FeIV

(OSub)(TPFPP)]+ transition state with orbital occupation

[core] x2–y22 *xz

1 *yz

1 a1u

2 a2u

2 Sub

1 with Sub a radical on the substrate group.

Figure 4.5: UB3LYP/BS1 optimized geometries of epoxidation transition states with

bond lengths in angstroms.

1.687

1.908

1.390

4TS(1)

i435.9 cm–1

rFeN,average = 2.011average = 0.287

1.7242.041

1.414

4TS(4)

i242.4 cm–1

rFeN,average = 2.003average = 0.292

1.644

2.1931.379

4TS(5)

i91.5 cm–1

rFeN,average = 2.015average = 0.288

1.6492.187

1.380

4TS(7)

i167.5 cm–1

rFeN,average = 2.013average = 0.295

1.663

2.171 1.404

4TS(6)

i214.7 cm–1

rFeN,average = 2.006average = 0.290

1.648

2.237

1.391

4TS(8)

i134.6 cm–1

rFeN,average = 2.012average = 0.299

1.6662.072

1.380

4TS(2)

i303.5 cm–1

rFeN,average = 2.014average = 0.296

1.665

2.073

1.381

4TS(3)

i299.9 cm–1

rFeN,average = 2.015average = 0.298

1.6542.145

1.387

4TS(9)

i248.3 cm–1

rFeN,average = 2.002average = 0.268

1.642

2.3231.380

4TS(10)

i59.2 cm–1

rFeN,average = 2.015average = 0.293

Page 119: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

119

4.5 Discussion

The present work gives a detailed and extensive overview on the reactivity of iron(IV)-oxo

porphyrin systems with a test-set of olefins. Crestoni and co-workers determined rate

constants and measured product ion distributions in the gas phase using FT-ICR MS. This

comprehensive set of transition metal containing reactivities is unique and will enable

computation to benchmark and calibrate its methods effectively. This is particularly

important for transition metal complexes, such as iron(IV)-oxo species, where the

reproducibility of the computational (DFT) methods sometimes varies strongly depending on

the density functional method used, the basis set, environmental perturbations, dispersion

effects etc (A. Ghosh and P. R. Taylor, 2003, M. Swart, A. R. Groenhof, et al 2004, de

Visser, P.; Quesne, G. et al 2004). In this work, we supplemented the experimental studies

with a series of preliminary DFT calculations for two reasons: (i) to validate and calibrate

computational methods against experiment; (ii) to establish the physicochemical properties

that influence the rate constant of the chemical reaction.

Let us first start with a comparison of the experimental and computational reaction rates.

As FT-ICR MS experiments are being performed at very low pressures, these experimental

conditions are close to Ideal Gas conditions with very few molecular collisions per second. In

the kinetic study of ion-molecule reactions in the gas phase one needs to consider that

thermal equilibration of the reacting system with the environment is in general not granted.

On the contrary, when small species react at low pressures, the absence of thermalizing

collisions leads to non-equilibrium energy distributions. In the absence of solvation, the

double-well potential model first proposed by Brauman in 1977 to account for the kinetic

behaviour of displacement reactions by anionic nucleophiles predicts that the energy of the

Page 120: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

120

intermediate and transition state must lie below the energy of the combined reactants

(Chabinyc, L.; Craig, L. et al 1998). Because at low pressure the intermediates cannot be

stabilized by unreactive collisions, determining the transition state energy is less

straightforward than in solution. However, the kinetics results presently reported deal with a

relatively large reactant ion that effectively establishes thermal equilibrium with the

environment through coupling with the background radiation field allowed by the several low

frequency IR modes of the iron(IV)-oxo macrocyclic ligand complex. This condition is

responsible, for example, for the consistency between the kinetics of NO ligand addition

kinetics to iron(II/III) porphyrin complexes and the equilibrium data independently

established through equilibrium measurements (Angelelli, F.; Chiavarino, B. et al 2005).

Because of these considerations, the notion can be adopted that the presently investigated

systems are in prevailing thermal equilibrium with the environment and reaction kinetics can

be interpreted within the framework of transition state theory. Consequently, reaction rates

represent bimolecular reactions and as such they should compare to computationally

determined reaction rates well.

Page 121: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

121

Figure 4.6: Correlation between experimental and computational barrier heights.

0

5

10

15

20

0 5 10 15 20RT ln kexp [kcal mol‒1]

Hcalc

[kcal mol‒1]

Page 122: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

122

The experimental rate constants were converted into free energy units via RT ln kexp, with R

the gas constant and T the temperature, using transition state theory and plotted against the

calculated enthalpy of activation for the same substrates, see Fig 4.6. Although only a limited

computational study is reported here, when we calculate the deviation between experiment

and theory for each data point, we find an average difference between experiment and theory

of 1.5 kcal mol–1

with a standard deviation of 3.4 kcal mol–1

. As such, the DFT methods used

here reproduce the trends obtained from experimental enthalpies of activation well and the

linearity and reproducibility of the calculations is well within the typical error reported for

DFT calculations using this method of about 5 kcal mol–1

(Y. Zhao and D. G. Truhlar, et al

2008, Schwabe, T.; Grimme, S. 2008, Weymuth, T.; Couzijn, A. et al 2014). There is,

however, a large systematic error as well as a relatively large standard deviation that require

further studies. Note that the experimental data in Figure 4.6 refers to free energies of

activation, whereas the computational results are enthalpy changes instead. The systematic

error between experiment and theory contains entropic corrections to the energy.

Figure 4.7: (a) Correlation between experimentally determined RT ln kexp (for raw data,

see Table 4.1) versus known ionization energies (IE). (b) Correlation between

calculated epoxidation activation enthalpy (in kcal mol–1

) and experimental ionization

energy for the substrates in Fig 4.5.

y = 0.15x + 7.89

R² = 0.82

6

8

10

12

0 5 10 15 20

IE

[eV]

Hcalc [kcal mol‒1]

(b)

y = -0.382x + 8.55

R² = 0.96

6

8

10

12

-6 -4 -2 0 2

IE

[eV]

RT ln kexp [kcal mol‒1]

(a)

Page 123: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

123

Subsequently, we investigated the origin of the rate constant, and in particular, the physical

and chemical properties of the substrate and oxidant that determine the reaction mechanism

and the enthalpy of activation of an epoxidation reaction. Previous studies on heteroatom

oxidation and double bond epoxidation by P450 enzymes implicated a correlation between

the natural logarithm of the rate constant with the ionization energy of the substrate (Crestoni,

E.; Fornarini, S. et al 2009, Lanucara, F.; Crestoni, E. 2011, de Visser, S.P.; Ogliaro, F. et al

2002, Watanabe, Y.; Iyanagi, T. et al 1980). To find out whether the data in Table 4.1 follow

these trends as well, we plot RT ln kexp versus experimentally known ionization energies

(Lias, G.; Bartmess, E. et al), see Fig 4.7. The set of data shown in Table 4.1 and Fig 4.7

gives a linear correlation between the natural logarithm of the rate constant and the ionization

energy of the substrate with an R2 = 0.96. Fig 4.7(b) displays the correlation between the

DFT calculated enthalpy of activation of the reaction of [FeIV

(O)(Por+•

)]+ with olefins. In

agreement with the experimental trends given in part (a) of Fig 4.7 also the computational

trends link the natural logarithm of the rate constant to the ionization energy of the substrate.

Clearly, the key physicochemical property that drives the reaction mechanism and affects the

rate constant of substrate epoxidation by iron(IV)-oxo complexes is the ionization energy of

the substrate.

In order to explain the experimental and computational trends in the reaction mechanisms,

we devised a valence bond (VB) curve crossing diagram, which is schematically depicted in

Fig 4.8. This diagram starts bottom left with the reactant configuration of

[FeIV

(O)(TPFPP+•

)]+ in electronic configuration xz

2 yz

2 x2–y2

2 *xz

1 *yz

1 a2u

1. The and *

electrons along the FeO bond are identified with dots in the VB diagram and due to

occupation of xz2 *xz

1 there are three dots on the left-hand-side of the Fe–O bond. In

addition, there are three electrons in the yz and *yz orbitals, which are identified with the

other three dots on the right-hand-side of the Fe–O bond. Furthermore, the oxidant has a

Page 124: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

124

radical on the porphyrin ring for single occupation of the a2u molecular orbital. The substrate

double bond is also highlighted with four electrons spread out over the interaction. Upon

approach of the substrate on the iron(IV)-oxo species a radical intermediate is formed that has

a single bond between the oxygen and carbon atoms and a doubly occupied a2u orbital.

Figure 4.8: VB curve crossing diagram for the C–O bond formation step in olefin

epoxidation (R2C=CH2) by [FeIV

(O)(TPFPP+•

)]+. Valence electrons are identified with

a dot and lines (curved and straight) in the VB structures represent bonds.

In VB theory the electronic configuration in the reactant complex (R) connects to an excited

state in the product geometry (P*) as shown with the blue line in Fig 4.8. At the same time

the product electronic configuration (P) connects to an excited state in the reactant geometry

(R*), black line in Fig 4.8. These two VB curves cross and lead to an avoided crossing and a

transition state for the C–O bond formation with barrier E‡. The barrier height is linearly

proportional to the curve crossing energy, which in its own right is a fraction of the excitation

energy (GH) from the reactant wave function to the product wave function in the geometry of

Page 125: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

125

the reactants, i.e. for R R*. The difference in VB structures for R and R* thereby

should give a reflection of the key electron transfer/migrations upon product formation.

Moreover, based on the excitation energy, the factors that determine the barrier height can be

predicted.

An analysis of the differences between the reactant and product wave functions in the

reactant geometry reveals the following information: First of all, a comparison of the VB

structures of R and R* shows that the electrons in the -bond of the olefin are singlet

paired in the ground state and triplet coupled in the excited state, hence the excitation energy

GH includes the -* electron excitation in the substrate, Eex,Sub. Generally, the first

ionization potential of an olefin corresponds to the removal of an electron from a -orbital,

and, hence, is proportional to the * excitation energy. Indeed, our experimentally and

computationally determined barrier heights correlate linearly with the ionization energy of

the olefin, and therefore support the VB model.

One of the electrons originating from the -bond of the olefin forms a bond with the *xz

electron along the FeO bond, to create the C–O bonding pair of electrons. This means that the

xz/*xz pair of orbitals during the reaction splits back into individual atomic orbitals namely

3dxz(Fe) and 2px(O). The 2px(O) electron pairs with the electron from the substrate, while one

of the electrons of the 3dxz(Fe) orbital is transferred into the a2u orbital through internal

excitation/rehybridization of the oxidant, Eex,ox. The promotion gap, GH, therefore, will be

proportional to the * excitation in the substrate and the 3dxz to a2u electron transfer in the

oxidant: GH = Eex,Sub + Eex,ox. Obviously, since the ionization energy represents the energy to

remove an electron from a -type orbital of an olefin, this will imply a linear correlation

between the first excited state and the ionization energy of the substrate (de Visser, S.P.;

Ogliaro, F. et al 2002). The VB diagram, therefore, confirms a linear correlation between the

Page 126: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

126

ionization energy of the substrate and the C–O bond formation enthalpy of activation as

shown above.

Although, the oxygen atom transfer reaction between [FeIV

(O)(TPFPP+

)]+ and an olefin

could lead to either epoxide or hydroxylated products, unfortunately the FT-ICR MS

experiments cannot distinguish the two. Thus, several substrates Scheme 4.2 and Table 4.1

contain aliphatic groups that in a reaction with an iron(IV)-oxo group can be converted into

an alcohol. A correlation between the rate constant of oxygen atom transfer and the ionization

energy, Fig 4.7a, of the olefin provides indirect experimental evidence that all reactions lead

to epoxidation products. In fact, hydrogen atom abstraction reactions should not correlate

with the ionization potential of the substrate, but were shown to be proportional to the

strength of the C–H bond of the substrate that is formed (Lanucara, F.; Crestoni, E. 2011,

Watanabe, Y.; Iyanagi, T. et al 1980). To test that the rate constants do not correlate with the

bond dissociation energy (BDECH) of the C–H bond of the substrate that is broken, we plot in

Fig 4.9 calculated BDECH and barrier heights of selected olefins, namely propene, Z-2-

butene, E-2-butene, cyclohexene and 1,3-cyclohexadiene. As can be seen from Fig 4.9, no

correlation between BDECH and barrier height exists, and, therefore, hydrogen atom

abstraction is not the rate determining step in the reaction mechanism. Further evidence that

hydrogen atom abstraction reactions can be ruled out here comes from kinetic isotope effect

(KIE) studies. We measured the rate constant of oxygen atom transfer with cyclohexene and

cyclohexene-d10 and determined a KIE = kH/kD ~ 1 (Table 4.1). Consequently, the oxygen

atom transfer is unlikely to proceed with an initial hydrogen atom abstraction and double

bond epoxidation will be the dominant pathway.

The experimental trends, therefore, provide the first indirect experimental evidence that in

the gas-phase the regioselectivity of double bond epoxidation versus aliphatic hydroxylation

will be in favour of the epoxidation pathway. This implies that in enzymatic systems, such as

Page 127: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

127

the cytochromes P450, the shape and size of the substrate binding pocket will influence the

regioselectivity of hydroxylation over epoxidation and can change the natural preference

away from epoxidation.

Finally, the calculations presented in this work obviously refer to gas-phase results and

hence correlate well with gas-phase mass spectrometric data. In order to further establish that

the work can be extrapolated to solution phase, we did a series of single point calculations

using the polarized continuum model with a dielectric constant of = 4.7 to mimic a solution.

The obtained correlation between solvent corrected free energies of activation of epoxidation

reactions by [FeIV

(O)(Por+

)]+ is plotted against the solvent corrected ionization energy of all

substrates although it is not shown in here. Even in solvent, the linear trend in the correlation

between free energy of activation and ionization energy is retained, therefore, we expect to be

able to extrapolate our results to the solution phase as well.

Figure 4.9: Correlation between calculated epoxidation activation enthalpy (in kcal

mol–1

) and BDECH for the substrates.

y = -0.52x + 49.12

R² = 0.34

0

5

10

15

20

60 65 70 75 80 85 90

BDECH [kcal mol–1]

Hcalc

[kcal mol‒1]

Page 128: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

128

4.6 Conclusion

In this work we report a comprehensive combined mass spectrometric and computational

study on substrate epoxidation by iron(IV)-oxo porphyrin complexes in the gas phase. We

present a novel method to synthesize [FeIV

(O)(TPFPP+

)]+ in the gas phase at low pressure.

Furthermore, we report a large set of experimentally derived rate constants and product

distributions. All olefins undergo oxygen atom transfer, whereas compounds with low

ionization energy also give a certain degree of hydride transfer and charge transfer reactions.

Our experimentally determined reaction rates correlate linearly with the ionization potential

of the substrate and show that the electron transfer from substrate to oxidant is rate

determining. A thorough computational survey has confirmed the suggested mechanism and

provides a rationale for the observed trend in the rate constants.

Page 129: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

129

CHAPTER 5

PROJECT THREE

Page 130: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

130

Rationalization of the barrier height for para-Z-

styrene epoxidation by iron(IV)-oxo porphyrins

with variable axial ligands.3

ABSTRACT

A versatile class of heme monoxygenases involved in many vital functions for human health

are the cytochromes P450, which react via a high-valent iron(IV)-oxo heme cation radical

species called Compound I. One of the key reactions catalyzed by these enzymes is C=C

epoxidation of substrates. We report here a systematic study into the intrinsic chemical

properties of substrate and oxidant that affect reactivity patterns. To this end, we investigated

the effect of styrene and para-substituted styrene epoxidation by Compound I models with

either an anionic (chloride) or neutral (acetonitrile) axial ligand. We show, for the first time,

that the activation enthalpy of the reaction is determined by the ionization potential of the

substrate, the electron affinity of the oxidant as well as by the strength of the newly formed

C–O bond (approximated by the bond dissociation energy, BDEOH). We have set up a new

valence bond model that enables us to generalize substrate epoxidation reactions by iron(IV)-

oxo porphyrin cation radical oxidants and make predictions of rate constants and reactivities.

We show here that electron withdrawing substituents lead to early transition states, whereas

electron donating groups on the olefin substrate give late transition states. This affects the

barrier heights in such a way that electron withdrawing substituents correlate the barrier 3 Devesh Kumar, Reza Latifi, Suresh Kumar, Elena V. Rybak-Akimova, Mala A. Sainna, and Sam P. de Visser, “Rationalization of the barrier

height for para-Z-styrene epoxidation by iron(IV)-oxo porphyrins with variable axial ligands” Inorganic Chemistry, 2013, 52 (14), 7968–

7979

Page 131: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

131

height with BDEOH, while the electron affinity of the oxidant is proportional to the barrier

height for substrates with electron donating substituents.

5.1 Introduction.

An important class of enzymes in human physiology are the cytochromes P450, which are

a large set of heme enzymes involved in the biodegradation and metabolism of toxic

compounds in the liver (Sono, M.; Roach, M. et al 1996, Groves, J. 2003, Ortiz de

Montellano, P. 2004, Kadish, K.; Smith, K. et al 2010, Ortiz de Montellano, P. 2010, de

Visser, S.; Kumar, D. 2011). These enzymes utilize molecular oxygen on a heme center in a

catalytic cycle that uses two electrons and two protons to generate an iron(IV)-oxo heme

cation radical active species called Compound I (CpdI) (Denisov, I.; Makris, T. et al 2005,

Rittle, J.; Green, M. 2010). This species is the active oxidant of the P450 enzymes and reacts

with substrates via, for instance, aliphatic and aromatic hydroxylation, double bond

epoxidation, N-dealkylation and sulfoxidation (Groves, J.; Shalyaev, K. et al 2000, Nam, W.

2007). Because of its versatility in substrate activation the group of enzymes has attracted

interest from the biotechnological and pharmaceutical industries although the intricate details

of its catalytic mechanism and reactivity with substrates are still poorly understood. The

epoxidation of olefins is a common reaction in P450 enzymes for a variety of important

bioprocesses in the body including the activation of unsaturated fatty acids (Ruettinger, R.;

Fulco, A. 1981, Guengerich, F. 2003, McLean, K.; Munro, A. 2008). As a result, substrate

epoxidation by P450 isozymes is well studied for a wide range of (non)natural substrates

(Groves, J.; Avaria-Neisser, G. et al 1986, Alcalde, M.; Farinas, E. et al 2004, Dansette, P.;

Bertho, G. et al 2005, Yuan, X.; Wang, Q. et al 2009, Shinkyo, R.; Xu, L. et al 2011). For

instance, the reactions were shown to be highly enantioselective, whereby cis--methyl

styrene substrate gave an 89:11 epoxide product ratio of the 1S,2R over the 1R,2S form (Ortiz

de Montellano, P.; Fruetel, J. et al 1991).

Page 132: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

132

Since CpdI is a very versatile oxidant, many studies investigated synthetic analogues and

models in substrate oxidation (Groves, J.; Myers, R. et al 1983, Groves, J.; Watanabe, Y.

1986, Collman, J.; Kodadek, T. et al 1986, Ostović, D.; Bruice, T. 1989, Collman, J.;

Brauman, J. et al 1990, Groves, J.; Gross, Z. et al 1994, Stephenson, N.; Bell, A. 2006,

Collman, J.; Zeng, L. et al 2006, Hessenauer-Ilicheva, N.; Franke, A. et al 2007, Mas-

Balleste, R.; Que Jr., L. 2007, Comba, P.; Rajaraman, G. 2008, Bruijnincx, P.; Buurmans, I.

et al 2008, Hull, J.; Sauer, E. et al 2009, McGown, A.; Kerber, W. et al 2009, Company, A.;

Feng, Y. et al 2009, Franke, A.; Wolak, M. et al 2009, Hessenauer-Ilicheva, N.; Franke, A. et

al 2009, Leeladee, P.; Goldberg, D. 2010, Lanucara, F.; Crestoni, M. 2011). A number of

these biomimetic studies focused on double bond epoxidation mechanisms and one of the

most common substrates in these studies is styrene. Thus, a series of styrene epoxidation

studies with a CpdI mimic with varying axial ligand gave rate constants that were

proportional to its electron donating ability (Gross, Z.; Nimri, S. 1994, Gross, Z. 1996,

Czarnecki, K.; Nimri, S. et al 1996). Green and co-workers (Green, M.; Dawson, J. et al

2004) characterized the axial ligand effect as arising from changes in pKa values of the oxo

group, which was confirmed with density functional theory (DFT) calculations on model

complexes (Kumar, D.; Karamzadeh, B. et al 2010).

Experimental studies reported rate constants for styrene epoxidation by synthetic iron(IV)-

oxo complexes: [FeIV

(O)(TPFPP+•

)X]0/+

with TPFPP = meso-

tetrakis(pentafluorophenyl)porphyrinato and X = Cl– or NCCH3 (Song, W.; Ryu, Y. et al

2005). In an intriguing set of experiments, it was shown that the reaction of

[FeIV

(O)(TPFPP+•

)NCCH3]+ with ethylbenzene gave aromatic hydroxylation products,

whereas the one using [FeIV

(O)(TPFPP+•

)Cl] as an oxidant led to benzylic hydroxylation

instead, hence the axial ligand in iron(IV)-oxo porphyrins seems to affect the product

distributions in the reaction processes and the regioselectivity of aromatic over aliphatic

Page 133: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

133

hydroxylation (Song, W.; Ryu, Y. et al 2005). The two oxidants with chloride versus

acetonitrile as axial ligand also gave differences in styrene epoxidation using para-substituted

styrene derivatives. The work identified a correlation between the rate constant of styrene

epoxidation with the + Hammett factor, but the slopes were different for the CpdI models

with anionic versus neutral axial ligands. The fundamental nature of this axial ligand effect

and how it affects reaction mechanisms, rate constants and product distributions of iron(IV)-

oxo porphyrins is unknown and, therefore, warrant a computational study. Thus, to gain

insight into the effect of axial ligands on the reactivity of iron(IV)-oxo porphyrins with

olefins we have done a density functional theory study of the activation barriers using a

selection of para-substituted styrene as substrates and Compound I models with meso-

substituted porphyrin.

Fujii and co-workers did a systematic investigation into the redox potentials of

[FeIV

(O)(TMP)X]n+

with TMP = 5,10,15,20-tetramesitylporphyrinate and X an axial ligand

that is either anionic or neutral (Takahashi, A.; Kurahashi, T. et al 2011, Cong, Z.; Kurahashi,

T. et al 2011). They found a positive E½ shift upon binding of an anionic axial ligand that was

virtually constant for a range of ligands. By contrast, binding of neutral ligands, such as

imidazole gave redox potentials in line with those found for peroxidases corresponding to a

negative E½ shift. They then replaced the TMP group with a meso-substituted porphyrin

ligand with electron-withdrawing groups, which led to an increase of the E½ values.

Currently, it is not clear what factors determine the redox potentials of iron-porphyrins and

whether there is a relationship with catalysis. Therefore, to gain insight into the effect of

meso-substitution on the intrinsic chemical properties of the oxidant and the subsequent

catalysis of substrates we decided to include this study in the present work.

The model we chose for our studies presented here is given in Figure 5.1 and is based on

the system used in Ref (Song, W.; Ryu, Y. et al 2005). The basic features of the oxidant are

Page 134: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

134

an iron(IV)-oxo group embedded in a protoporphyrin IX (Por) without side chains and with

either chloride or acetonitrile as the axial ligand: [FeIV

(O)(Por+•

)X]0/+

with X = Cl– or

NCCH3. In addition, we also investigated a more elaborate model that uses pentafluorophenyl

substituents on the meso-position of the porphyrin ring, i.e. meso-

tetrakis(pentafluorophenyl)porphyrinato or TPFPP. Substrate epoxidation was studied using a

range of para-Z-substituted styrenes with Z = H, F, Cl, CH3, t-Bu, CN, NO2, OCH3, NH2 and

N(CH3)2. We show that para-substitution affects the ionization potential of the substrate and

its electron donating ability, which leads to changes in the activation barrier of oxygen atom

transfer. Although we reported a detailed analysis of substrate epoxidation by iron(IV)-oxo

porphyrins before (Kumar, D.; Karamzadeh, B. et al 2010), the systematic study described

here goes beyond that of the previous work and highlights the differences in reactivity of

substrates with electron donating versus electron withdrawing substituents. Moreover, a new

model is presented that correlates with the obtained reactivity trends.

Figure 5.1: Oxidants and substrates used in this work.

5.2 Methods

The study presented here uses density functional theory methods as implemented in the

Jaguar, Gaussian-03 and Gaussian-09 program packages (Schrödinger, 2011). All geometries

are the result of a full geometry optimization, whereby all degrees of freedom are minimized.

Z

Substrates:Z = H, F, Cl, CH3, t-Bu, CN,NO2, OCH3, NH2, N(CH3)2

FF

F

FF

Y = H,

NH N

HNN

Y

Y

Y

Y

Por: Y = H; TPFPP: Y =

FF

F

FF

X = Cl- or NCCH3

X

Page 135: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

135

A subsequent analytical frequency calculation characterized the structures as local minima

(with real frequencies only) or first order saddle points with one imaginary frequency for the

correct mode. The hybrid B3LYP method (Becke, A. 1993, Lee, C.; Yang, W. et al 1988) was

employed throughout in combination with the Los Alamos-type LACVP basis set on iron and

6-31G on the rest of the atoms (BS1) for the geometry optimizations and frequencies (Hay, P.

J.; Wadt, W. 1985, Hehre, W. J.; Ditchfield, R. et al 1972). Energies were then improved by

single point calculations with a triple- type basis set on iron (LACV3P+) and 6-311+G* on

the rest of the atoms; BS2. Benchmark studies against experimental data reproduced free

energies of activation using these methods to within 3 kcal mol–1

(de Visser, S.; Oh, K. et al

2007, Vardhaman, A.; Sastri, C. et al 2011). Previously, we calculated a full potential energy

profile for substrate hydroxylation by an iron(IV)-oxo porphyrin cation radical system at

UB3LYP/BS2 and obtained relative energies within a few tenth of a kcal mol–1

for local

minima and first order saddle points along a reaction mechanism as compared to those

obtained at UB3LYP/BS2//UB3LYP/BS1, hence the latter method was used here (de Visser,

S. 2010). Single point calculations in a dielectric constant with = 37.5 mimicking

acetonitrile with a probe radius of 2.1 Å were performed in Jaguar using basis set BS2. Free

energies reported here were calculated at 298K temperature and 1 bar pressure and are based

on the UB3LYP/BS2//UB3LYP/BS1 energies and corrected with ZPE, thermal and entropic

corrections from the frequency file and with solvent corrections from the single point solvent

calculation. Further corrections to the energy were made by performing single point

calculations with dispersion corrected B3LYP as implemented in Jaguar (Schwabe, T.;

Grimme, S. 2007). For selected structures we also did geometry optimizations and

frequencies using UB3LYP-D/BS1 in Jaguar, although little changes in the optimized

geometries were obtained.

Page 136: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

136

We used two synthetic iron-porphyrin models that are distinguished by the choice of the

axial ligand, which was either chloride or acetonitrile. Note that [FeIV

(O)(Por+•

)Cl]0 is overall

charge neutral, while [FeIV

(O)(Por+•

)NCCH3]+ is positively charged. In our initial

calculations we used a bare porphyrin ring, whereby all side chains were replaced by

hydrogen atoms. In a second set of calculations we studied a TPFPP ligand system, TPFPP =

tetra(pentafluorophenyl)porphyrin. We studied styrene epoxidation using a selection of para-

substituted styrene substrates as described in Figure 5.1.

To ascertain that the results are not influenced by the density functional method chosen

here, we ran a selection of single point calculations using dispersion corrected DFT

(Schwabe, T.; Grimme, S. 2007) and B3LYP with 15% HF exchange (designated B3LYP*)

(Reiher, M.; Salomon, O. et al 2001). As before (Kumar, D.; Thiel, W. et al 2011), these test

calculations reproduced the trends obtained with the B3LYP method so that the results give a

systematic error, which does not affect our discussion and analysis as we are dealing with

trends only here. Nevertheless, for comparison we give here the results obtained for E+ZPE

(UB3LYP), G + Esolv (UB3LYP) and G + Esolv + Edisp (UB3LYP-D).

5.3 Results

We started the work with a detailed study into the electronic properties of iron(IV)-oxo

complexes, the effects of axial versus equatorial ligands and finally the reactivity patterns

with a selection of para-substituted styrenes. Our models include [FeIV

(O)(Por+•

)X]0/+

,

designated 1X, and [FeIV

(O)(TPFPP+•

)X]0/+

, designated 2X, with X = Cl– and NCCH3, Figure

5.1. Similar to previous studies on CpdI of P450 and biomimetic iron(IV)-oxo porphyrin

complexes (Green, M. 1999, Ogliaro, F.; de Visser, S. et al 2001, Kamachi, T.; Yoshizawa,

K. 2003, Bathelt, C.; Zurek, J. et al 2005, Lonsdale, R.; Oláh, J. et al 2011, Isobe, H.;

Yamanaka, S. et al 2011), all complexes have the same electronic ground state with four

Page 137: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

137

electrons in metal 3d-type orbitals and a radical on the porphyrin macrocycle. High lying

occupied molecular orbitals include the x2–y2 orbital that is nonbonding, doubly occupied and

located in the plane of the porphyrin. Slightly higher in energy are two *FeO orbitals (*xz,

*yz) for the antibonding interaction of the metal 3dxz/yz with 2px/y atomic orbitals on oxygen.

Higher lying and virtual are the *z2 and *xy orbitals for the antibonding interactions along

the Fe–O axis and between the Fe–N groups in the plane of the porphyrin ring. In addition

there is a radical on a porphyrin type orbital that in D4h symmetry has the label a2u. The

orbital occupation gives close lying doublet and quartet spin states with configuration x2–y22

*xz1 *yz

1 a2u

1, whereby the two * electrons are either ferromagnetically or

antiferromagnetically coupled to the a2u electron in the quartet and doublet spin states. As a

consequence the two spin states are close in energy and the oxidant reacts via two-state-

reactivity patterns on close-lying doublet and quartet spin state surfaces (Shaik, S.; Kumar, D.

et al 2005).

Page 138: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

138

Figure 5.2: UB3LYP/BS1 optimized geometries of 4,2

1X and 4,2

2X in the gas phase with bond

lengths in angstroms. Group spin densities are obtained at UB3LYP/BS2//UB3LYP/BS1 and

are reported in atomic units.

Figure 5.2 gives optimized geometries of 4,2

1X and 4,2

2X with X = Cl–/NCCH3 as calculated

with DFT. Optimized geometries of 4,2

1X (X = Cl–/NCCH3) are almost identical to those

reported before (de Visser, S.; Tahsini, L. et al 2009). The Fe–O distances are short and

typical for an iron(IV)-oxo species and shorten somewhat with an axially ligated acetonitrile

molecule as compared to chloride. This is due to mixing of the a2u and * orbitals in 4,2

1Cl

41Cl (21Cl)

42Cl (22Cl)

41NCCH3 (21NCCH3)

42NCCH3 (22NCCH3)

rFeO = 1.660 (1.657)

rFeCl = 2.419 (2.433)

rFeO = 1.660 (1.658)

rFeCl = 2.424 (2.436)

rFeO = 1.646 (1.645)

rFeNax = 2.126 (2.129)

Fe = 1.00 (1.12)

O = 1.02 (0.96)

Cl = 0.19 (–0.17)

TPFPP = 0.80 (–0.92)

Fe = 1.03 (1.14)

O = 0.99 (0.95)

Cl = 0.15 (–0.13)

Por = 0.83 (–0.95)

Fe = 0.97 (1.01)

O = 1.05 (1.05)

NCCH3 = 0.00 (–0.02)

Por = 0.97 (–1.04)

rFeO = 1.648 (1.647)

rFeNax = 2.112 (2.121)

Fe = 0.95 (0.99)

O = 1.07 (1.07)

NCCH3 = 0.00 (–0.01)

TPFPP = 0.98 (–1.05)

Page 139: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

139

with 3p atomic orbitals on the chloride ligand (de Visser, S.; Tahsini, L. et al 2009), which

brings the metal more inside the plane through the porphyrin ring. This type of mixing has

been identified before as the key reason for the intrinsic electronic differences of CpdI in

monoxygenases and peroxidases (de Visser, S.; Shaik, S. et al 2003). We did additional

geometry optimizations of 21X (X = Cl

–/NCCH3) at UB3LYP-D/BS1 level of theory, which

reproduced the UB3LYP/BS1 structures within 0.010 Å.

As can be seen, meso-substitution of the porphyrin ring with pentafluorophenyl groups has

little effect on the optimized geometries and the Fe–O, Fe–Cl and Fe–Nax distances for

structures 1 and 2 are almost the same. Note also that very little saddling is found for the

[FeIV

(O)(TPFPP+•

)X] structures. Not surprisingly, since the same molecular orbitals are

singly occupied in the doublet and quartet spin state structures, their geometries are virtually

identical for each CpdI set of data. To establish whether the meso-pentafluorophenyl groups

give electronic changes to the iron(IV)-oxo species we also show in Figure 5.2 the group spin

densities (), but only minor differences are observed between the data for structures 1 and 2.

Subsequently, we calculated styrene epoxidation using a range of para-substituted styrene

derivatives. Before we discuss the results on the substituted styrenes, let us focus on the

general overall mechanism first. All epoxidation reactions described in this work proceed

with the same stepwise mechanism that starts with the formation of reactant complex

between CpdI and substrate, R. As an example, we give the reaction profile for

[FeIV

(O)(Por+

)Cl] with para-H-styrene in Figure 5.3 on the lowest doublet and quartet spin

states. The spin multiplicity is given in superscript and in subscript we give the axial ligand

(X) and the para-Z-substituent of styrene. The mechanism resembles that found in earlier

studies of substrate epoxidation of olefins by metal(IV)-oxo oxidants (de Visser, S.; Tahsini,

L. et al 2009, Linde, C.; Åkermark, B. et al 1999, Kamachi, T.; Shiota, Y. et al 2003,

Quiñinero, D.; Musaev, D. et al 2003, Sharma, P.; de Visser, S. et al 2003, Hirao, H.; Kumar,

Page 140: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

140

D. et al 2005, Kumar, D.; Derat, E. et al 2005, Bassan, A.; Blomberg, M. et al 2005, de

Visser, S.; Ogliaro, F. et al 2001, de Visser, S.; Ogliaro, F. et al 2001, de Visser, S.; Ogliaro,

F. et al 2002, de Visser, S.; Ogliaro, F. et al 2002, Kumar, D.; de Visser, S. et al 2005) and

starts with an initial electrophilic addition of the oxo group to the double olefinic bond of

styrene via a transition state TSX,Z to form a radical intermediate (IX,Z), whereby the subscript

X refers to the axial ligand, i.e. X = Cl– or AN (acetonitrile) and the subscript Z to the para-

substituent of styrene. In a final reaction step a ring-closure transition state (TSrc,Z) leads to

epoxide product complexes (PX,Z). Optimized geometries obtained at UB3LYP-D/BS1 show

little difference with those obtained at UB3LYP/BS1, Figure 5.3.

Figure 5.3: Potential energy profile of styrene epoxidation by 4,2

1Cl as calculated with

UB3LYP/BS2//UB3LYP/BS1. All energies are in kcal mol–1

relative to isolated reactants in

the doublet spin state and include ZPE corrections. Also shown are optimized geometries of

critical points with bond lengths in angstroms and the imaginary frequency in the transition

state in wave numbers. Free energies are given in parenthesis and include UB3LYP/BS2

21Cl+styrene-H

41Cl+styrene-H

2ICl,H

4ICl,H

4TSCl,H

2TSCl,H

0.0 (0.0)

–0.1 (–0.4)

–15.0 (–3.5)

–9.6 (3.8)

7.4 (18.7)

[–1.2 (10.8)]

8.4 (19.4)

4TSrc,H

2TSrc,H

4PCl,H

2PCl,H

2TSCl,H (4TSCl,H)

rFeO = 1.682 (1.706)

[1.674]

rFeCl = 2.387 (2.371) [2.373]

rCO = 2.121 (2.072)

[2.197]

i563.2 (i437.1) [i424.4] cm–1

2ICl,H (4ICl,H)

rFeO = 1.784 (1.804)

rFeCl = 2.322

(2.340)

rCO = 1.472 (1.461)

–15.0

–8.2

–36.6 (–23.7)

–44.6 (–33.6)

CH2 CH

C6H5

C6H5

CH

CH2

C6H5

CHCH2

FeIV

O

X FeIV

O

X

FeIII

O

4,2I4,2P

X

Page 141: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

141

energies corrected with thermal and entropic corrections at 298 K. Data in square brackets

obtained after a UB3LYP-D/BS1 geometry optimization.

Although the ring-closure barrier was calculated for a selection of chemical systems,

in all cases its barrier was much smaller than the C–O bond formation barrier via TSX,Z,

therefore, we will focus here on the rate determining barrier only. In particular, in the quartet

spin state ring-closure barriers were located and found to be small, whereas in the doublet

spin state the ring-closure was virtually barrierless. The origin of this difference is due to

differences in electron transfer processes in the ring-closure step, whereby on the doublet spin

state surface the *xz orbital is filled with an extra electron, while on the quartet spin state a

higher lying and virtual *z2 orbital is filled with one electron (de Visser, S.; Ogliaro, F. et al

2001). Energies obtained for the reaction of [FeIV

(O)(Por+•

)Cl] with para-H-styrene are very

similar to those reported before using [FeIV

(O)(Por+•

)SH] as an oxidant (de Visser, S.;

Tahsini, L. et al 2009, de Visser, S.; Ogliaro, F. et al 2001). That is not surprising, since the

pKa and electron affinity of [FeIV

(O)(Por+•

)X], X = Cl–/SH

– are virtually the same.

Geometries are typical for epoxidation structures from previous calculations for P450

catalyzed reaction mechanisms (Linde, C.; Åkermark, B. et al 1999, Kamachi, T.; Shiota, Y.

et al 2003, Quiñinero, D.; Musaev, D. et al 2003, Sharma, P.; de Visser, S. et al 2003, Hirao,

H.; Kumar, D. et al 2005, Kumar, D.; Derat, E. et al 2005, Bassan, A.; Blomberg, M. et al

2005, de Visser, S.; Ogliaro, F. et al 2001, de Visser, S.; Ogliaro, F. et al 2001, de Visser, S.;

Ogliaro, F. et al 2002, de Visser, S.; Ogliaro, F. et al 2002, Kumar, D.; de Visser, S. et al

2005). In the transition state the Fe–O bond elongates slightly with respect to reactants and

further extends in the intermediate to 1.784 Å. At the same time considerable shortening of

the C–O distance occurs to a formally single bond in 4,2

ICl,H of 1.472 Å.

Subsequently, we investigated the reaction mechanisms of [FeIV

(O)(Por+•

)X]0/+

, X = Cl– or

NCCH3 with para-substituted styrene and the optimized geometries of the transition states

Page 142: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

142

are given in Figure 5.4. Interestingly, structural differences are found between the two axially

ligated oxidants, whereby the Cl– bound TSX,Z structures are with styrene in an upright

position, whereas the acetonitrile bound ones are more sideways bound. Test calculations

using an upright starting structure and acetonitrile axial ligand, however, converged back to

the sideways bound structure instead, which implicates that the upright structures are higher

in energy for the axially ligated acetonitrile TSs. The differences in substrate orientation

affect the bond distances of the respective transition states. For instance, the bond forming C–

O distances vary from 2.017 – 2.241 Å for the transition states with X = Cl–, whereas for X =

NCCH3 distances between 2.083 and 2.439 Å are found. This implies that the barriers for the

acetonitrile ligated structures are somewhat earlier than the ones for the chloride bound

systems. Although the variation in Fe–O and Fe–X distance is considerably smaller than

those for the C–O distance variations between the oxidant with chloride and acetonitrile

ligands are also found.

Figure 5.4: Optimized geometries of rate determining transition states 2TSX,Z for the reaction

of para-Z-styrene with 21X (X = Cl

–/NCCH3). Geometries optimized at UB3LYP/BS1 with

bond lengths given in angstroms and the imaginary frequency in wave numbers. Also given

Z H F Cl CH3 t-Bu CN NO2 OCH3 NH2 N(CH3)2

X = Cl–

rC–O 2.121 2.107 2.090 2.142 2.143 2.053 2.017 2.163 2.232 2.241

rO–Fe 1.682 1.684 1.685 1.680 1.680 1.689 1.692 1.679 1.674 1.674

rFe–Cl 2.387 2.386 2.385 2.390 2.389 2.383 2.382 2.392 2.399 2.401

imag i563.2 i600.4 i641.0 i500.3 i500.1 i693.3 i697.8 i421.5 i269.4 i245.7

E+ZPE 7.40 7.01 7.06 6.79 6.85 7.30 7.28 5.97 4.80 3.92

G+Esolv+Edisp

13.8 11.2 11.9 9.4 11.3 12.7 11.8 10.1 6.5 –0.3

X = NCCH3

rC–O 2.187 2.178 2.160 2.244 2.272 2.117 2.083 2.305 2.430 2.439

rO–Fe 1.668 1.673 1.677 1.659 1.653 1.692 1.703 1.653 1.648 1.646

rFe–NCCH3 2.103 2.099 2.096 2.115 2.128 2.088 2.084 2.130 2.150 2.163

imag i209.8 i220.3 i234.3 i181.1 i238.2 i274.4 i310.5 i147.8 i75.3 i66.0

E+ZPE 3.55 4.15 4.66 1.39 1.25 8.30 9.44 -1.94 -7.11 -10.56

G+Esolv+Edisp

6.3 8.7 10.7 5.7 7.7 24.8 27.0 4.9 –1.0

Z

Cl

Z

Page 143: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

143

are barrier heights (E+ZPE) for 2TSX,Z with energies calculated at

UB3LYP/BS2//UB3LYP/BS1+ZPE relative to isolated reactants in kcal mol–1

and free

energies of activation in solvent (G+Esolv+Edisp) relative to 2RCX,Z.

The epoxidation barriers (TSCl,Z) range from 3.92 kcal mol–1

for para-N(CH3)2-styrene to

7.40 kcal mol–1

for para-H-styrene, i.e. vary by 3.48 kcal mol–1

upon para-substitution. With

acetonitrile as axial ligand the lowest barrier is 10.56 kcal mol–1

below isolated reactants for

para-N(CH3)2-styrene, although it should be noted that the reactant complex is more stable

than isolated reactants by 14.75 kcal mol–1

. Thus, the epoxidation barrier height varies by 20

kcal mol–1

between para-N(CH3)2-styrene and para-NO2-styrene for TSAN,Z, hence the

substituent located at a distance of over 4 Å from the reaction center causes a rate constant

change by a factor of ca. 1015

(estimated using transition state theory for an enthalpy-derived

free energy change of 20 kcal mol–1

; entropy of activation was not included in this estimate).

The ordering of the barrier heights is virtually the same for the two oxidants studied, namely

Z = N(CH3)2 < NH2 < OCH3 < CH3/t-Bu < F/Cl < CN < H. A plot of the barriers 4TSCl,Z

versus those calculated for 4TSAN,Z with the same methods gives a linear correlation.

Furthermore, calculations using a solvent model included give free energies of activation that

follow the same trends as values found in the gas-phase. Inclusion of dispersion corrections

on the solvent corrected free energies of activation gives a further systematic change of the

energetics. This highlights the fact that for the assignment of reactivity trends it does not

matter whether E+ZPE, G, G+Esolv or G+Esolv+Edisp energies are used. As the enthalpies

in a reaction mechanism are determined by the electronic changes of the reactants and we aim

to establish the intrinsic chemical properties that determine the reactivity patterns, we will

focus in the following on establishing trends through the E+ZPE sets of data.

Page 144: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

144

To understand the changes in transition state geometry between the para-substituted

styrene substrates, we plot in Figure 5.5 a selection of bond distances versus the barrier

height (E‡+ZPE) for the two CpdI models with either chloride or acetonitrile axial ligand.

Figure 5.5(a) shows the distance of the C–O bond that is formed in the process versus the

barrier height. An almost perfect linear correlation (R2 = 0.98) is obtained for the acetonitrile

ligated TSs and a satisfactory linear correlation is found for the chloride ligated system. Thus,

a drop in C–O bond length of 0.224 Å is found between the weakest epoxidating substrate

(para-NO2-styrene) and the strongest one (para-N(CH3)2-styrene) using a chloride axial

ligand, whereas the difference between these substrates is 0.256 Å for an axially ligated

acetonitrile molecule. This implies that the para-substituent of styrene has a single possibly

electrostatic effect on the C–O bond formation step and consequently the barrier height of the

reaction. To further ascertain that the optimized geometries and barrier heights are

reproducible we reoptimized a selection of TS structures with UB3LYP-D/BS1. Thus,

2TSAN,CN and

2TSCl,CN give Fe–O distances of 1.683 and 1.680 Å and C–O distances of 2.131

and 2.107 Å, respectively. These values are close to those given in Figure 5.4 calculated at

UB3LYP/BS1 level of theory.

y = -14.653x + 37.661

R² = 0.7678

y = -51.372x + 116.46

R² = 0.9767

-15

-10

-5

0

5

10

15

2.0 2.1 2.2 2.3 2.4 2.5

Cl

AN

E‡+ZPE

[kcal mol–1]

C–O distance [angstroms]

(a)

y = 0.0068x + 2.9263

R² = 0.8647

y = 0.0771x - 13.783

R² = 0.9426

-15

-10

-5

0

5

10

15

0 200 400 600 800

Cl

AN

E‡+ZPE

[kcal mol–1]

imaginary frequency[cm–1]

(c)

Page 145: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

145

Figure 5.5: Analysis of structural features of the transition states TSX,Z calculated at

UB3LYP/BS1 as a function of the height of the epoxidation barrier with respect to the: (a) C–

O distance, (b) Fe–X distance, and (c) imaginary frequency in the transition state. Data given

for X = Cl– axial ligand (diamonds) and X = acetonitrile (squares).

We also investigated the correlations of epoxidation barrier height with the Fe–O and Fe–X

distances in the transition states, and indeed these correlations are linear as well. With a

chloride axial ligand modest Fe–O and Fe–Cl changes are seen throughout the series of –

0.018 and +0.019 Å, respectively, between the outer ranges for para-NO2-styrene and para-

N(CH3)2-styrene. Much larger differences are found for the weakly bound acetonitrile

system, where the Fe–NCCH3 distance is elongated from 2.084 to 2.163 Å (+0.079 Å shift)

between para-NO2-styrene and para-N(CH3)2-styrene. Thus, the para-substituent of styrene

affects bond distances well over 5 Å from the catalytic center and weakens the metal to axial

ligand distances Fe–Cl and Fe–NCCH3 bonds dramatically.

Interestingly, even the value of the imaginary frequency gives a linear correlation with

barrier height, which implies that the para-substituent of styrene affects the height as well as

the width of the potential energy curve around the transition state. Since, both height and

width of the barrier change linearly for our series of substrates, this means that the area under

the curve could stay the same for these substrates. However, changes to the width of the

y = -176.72x + 428.7

R² = 0.9089

y = -233.96x + 496.28

R² = 0.9647

-15

-10

-5

0

5

10

15

2.0 2.1 2.2 2.3 2.4 2.5

Cl

AN

E‡+ZPE

[kcal mol–1]

Fe–X distance [angstroms]

(b)

Page 146: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

146

potential energy curve may have important effects on, for instance, kinetic isotope effects for

a reaction (Kumar, D.; de Visser, S. et al 2003, Kumar, D.; de Visser, S. et al 2004, Kumar,

D.; de Visser, S. et al 2004, de Visser, S. 2006). Thus, a small imaginary frequency correlates

with a broad and wide potential energy surface, whereas a large imaginary frequency

implicates a narrow and high peak. Tunnelling through a narrow and high peak should be

easier than through a broad peak, therefore, kinetic isotope effects associated with this

reaction may be affected as well. It is interesting to note that the imaginary frequencies are

substantially larger for [FeIV

(O)(Por+•

)Cl] than for [FeIV

(O)(Por+•

)NCCH3]+, hence the

barriers for the latter oxidant will be much broader in shape. If this trend also applies to

hydrogen atom abstraction reaction then this would imply that the calculations predict

[FeIV

(O)(Por+•

)Cl] to react with considerably larger kinetic isotope effect for the replacement

of hydrogen by deuterium atoms than [FeIV

(O)(Por+•

)NCCH3]+ with substrates. Indeed, Nam

and co-workers reported differences in KIE values for the two oxidants in aliphatic hydrogen

atom abstraction reactions (Takahashi, A.; Kurahashi, T. et al 2011, Cong, Z.; Kurahashi, T.

et al 2011).

In a final set of calculations, we investigated the effect of meso-substitution of the

porphyrin ring on the electronic properties of CpdI and the epoxidation of para-substituted

styrenes. We selected the TPFPP ligand system with pentafluorophenyl groups on the meso-

position of the porphyrin and calculated styrene epoxidation using the TPFPP oxidant:

[FeIV

(O)(TPFPP+•

)X]0/+

with X = Cl– and NCCH3 or 2X. We studied styrene epoxidation for

para-N(CH3)2-styrene and para-NO2-styrene and compared the mechanism and energy

profiles with those observed above for 1X. Figure 5.6 displays optimized geometries of the

epoxidation transition states TS′X,Z for the reaction of styrene with 2X,Z with Z = NO2 or

N(CH3)2 as calculated with DFT methods.

Page 147: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

147

Figure 5.6: Optimized geometries of epoxidation transition states 2TS′X,Z for the reaction of

2X (X = Cl–/NCCH3) with para-Z-styrene. Bond lengths are in angstroms and the value of the

imaginary frequency in wave numbers.

Geometrically, there are striking differences between TSX,Z, on the one hand, with TS′X,Z

on the other hand. Thus, the epoxidation barriers with TPFPP ligand system are characterized

with long C–O and short Fe–O distances, which implicates much earlier transition states

along the potential energy surface. Furthermore, both acetonitrile and chloride ligated TS′X,Z

structures are in upright configuration with a structure similar to the TSCl,Z geometries

reported in Figure 5.4. Probably, the stereochemical interactions with meso-substituents

prevent a substrate orientation analogous to TSAN,Z with a sideways attack on the oxo group.

In addition, the structures are stabilized with hydrogen bonding interactions of substrate C–H

groups with the fluoride atoms from the TPFPP ligand, and specifically those located in the

ortho-position of the meso-substituent. As a consequence, both the reactant complexes as

well as the transition states for epoxidation are considerably stabilized with respect to isolated

reactants. Each transition state structure in Figure 5.6 is stabilized by at least two or three C–

H---F hydrogen bonding interactions with distances between 2.4 – 2.6 Å. F–H hydrogen

bonding interactions can be dramatic and, for instance, trifluoromethanol as a solvent has

2TS′Cl,N(CH3)2 (2TS′AN,N(CH3)2)

rFeO = 1.667 (1.660)

rFeX = 2.394 (2.156)

rCO = 2.574 (2.188)

i53.9 (i162.8) cm–1

2TS′Cl,NO2 (2TS′AN,NO2)

rFeO = 1.676 (1.651)

rFeX = 2.380 (2.118)

rCO = 2.205 (2.305)

i399.3 (i322.8) cm–1

rHF = 2.551 (2.589)

rHF = 2.530 (2.524)rHF = 2.478

rHF = 2.455 (2.470)

E‡+ZPE = –3.8 (–19.8) kcal mol–1E‡+ZPE = 4.5 (7.0) kcal mol–1

Page 148: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

148

been shown to lead to considerable lowering of epoxidation and halogenation barriers heights

due to hydrogen bonding and charge transfer interactions (Ben-Daniel, R.; de Visser, S. et al

2003, de Visser, S.; Kaneti, J. et al 2003).

Recent computational studies of dehydrogenation of cyclohexadiene by

[FeIV

(O)(TPCPP+•

)Cl] with TPCPP = meso-tetrakis(pentachlorophenyl)porphyrin also

showed substrate stabilization due to weak hydrogen bonding interactions of chloride atoms

of the TPCPP ligand with C–H groups of the approaching substrate (Kumar, D.; Tahsini, L.

et al 2009). These interactions were shown to be particularly strong for epoxidation barrier

heights due to closer approach to the oxidant as compared to hydrogen atom abstraction

reactions and lowered epoxidation barriers significantly (de Visser, S. 2006). In line with this

it is not surprising that we find considerable lowering of the epoxidation barriers from 7.3 to

4.5 kcal mol–1

for X = Cl/Z = NO2, from 9.4 to 7.0 kcal mol–1

for X = AN/Z = NO2, from 3.9

to –3.8 kcal mol–1

for X = Cl/Z = N(CH3)2 and from –10.6 to –19.8 kcal mol–1

for X = AN/Z

= N(CH3)2 upon replacement of the equatorial ligand from Por to TPFPP. The big changes in

geometry for TSX,Z as compared to TS′X,Z also affect the shape of the potential energy

surface and the values of the imaginary frequencies. Both TS′Cl,NO2 and TS′Cl,N(CH3)2

structures have lower imaginary frequencies than TSCl,NO2 and TSCl,N(CH3)2, whereas this is

not the case for the axially ligated acetonitrile structures.

Thus, we report here the effect of meso-substitution on the electronic and reactivity

properties of iron(IV)-oxo porphyrin cation radical models. We find little changes in

electronic configuration of the various CpdI reactants. However, fluorine substituents on the

meso-position of TPFPP entice favorable electrostatic interactions with the approaching

substrate and stabilize the epoxidation transition states. In the following we will try to

rationalize these observations.

Page 149: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

149

5.4 Discussion

In this work we report a systematic set of epoxidations of substituted styrenes using four

different oxidants, namely 1Cl, 1NCCH3, 2Cl and 2NCCH3. During the substrate epoxidation

process two electrons are transferred from substrate to oxidant, which is reduced from

[FeIV

(O)(Por+•

)X] to [FeIII

(Por)X]. As a consequence, the electron affinity (EA) of the

oxidant and the ionization energy (IE) of the substrate should reflect the electron transfer

processes that take place. Indeed several experimental studies found a correlation between the

ionization energy of substrates and the rate constant of substrate epoxidation or sulfoxidation

(Lanucara, F.; Crestoni, M. 2011, Watanabe, Y.; Iyanagi, T. et al 1980, Watanabe, Y.;

Numata, T. et al 1981, Watanabe, Y.; Iyanagi, T. et al 1982). So far, no experimental studies

have been reported on the activation parameters of substrate sulfoxidation by synthetic

biomimetic iron(IV)-oxo porphyrin cation radical models, but the rate constants are expected

to correlate with the activation enthalpies. In earlier work we reported a systematic study of

olefin epoxidation by iron(IV)-oxo complexes and set up a model that predicts barrier heights

from known ionization potentials. Although that correlation still applies, we generalize it

further in this work using the data described here.

To understand the individual contributions of oxidant and substrate onto the reaction

process, i.e. transition states, we will investigate those separately. Let us start with a

discussion of the differences and comparison between 1Cl/2Cl and 1NCCH3/2NCCH3 in styrene

epoxidation reactions and their corresponding electron affinities. Thus, as shown in Figure

5.2 above the reactant complexes give little differences in geometric and electronic features

upon changing the equatorial ligand from Por to TPFPP.

We calculated EA values of 79.4 kcal mol–1

for 1Cl, 101.4 kcal mol–1

for 2Cl, 148.7 kcal

mol–1

for 1NCCH3 and 163.8 kcal mol–1

for 2NCCH3 in the gas-phase. Note here that despite the

fact that the valence orbital occupations and orbital shapes seem little influenced by the

Page 150: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

150

nature of the meso-substituents of the porphyrin ring, actually the EA values increase by as

much as 15.1 – 22.0 kcal mol–1

. This observation is in excellent agreement with the

electrochemical studies of Fujii and co-workers who found similar trends (Takahashi, A.;

Kurahashi, T. et al 2011).

Since, the electrophilic reaction mechanism results in the electron transfer from substrate

into the oxidant orbitals, these differences in electron affinity of the four oxidants also affect

the subsequent reaction mechanisms and reactivities with substrates. The transition state

structures for these reactions, however, show dramatic differences in group spin densities as

displayed in Figure 5.7. Thus, generally the reaction proceeds with electron transfer from the

substrate into the half-filled a2u porphyrin orbital, whereby a decrease of spin density of the

porphyrin group is found. The electronic configuration of the TSCl,Z structures is very much

product-like with Por in the range from –0.29 to –0.47, whereas the TSAN,Z values are

reactant-like with considerably more spin density on the porphyrin ring.

Figure 5.7: UB3LYP/BS2//UB3LYP/BS1 calculated group spin density ranges for para-Z-

styrene epoxidation by 1Cl (top) and 1AN (bottom).

Z

ClZ

Fe = 1.43 – 1.64

O = 0.21 – 0.54

Sub = –0.37 – –0.53

Por = –0.47 – –0.29

Cl = –0.08 – –0.12

Fe = 1.07 – 1.40

O = 0.51 – 0.78

Sub = –0.02 – –0.85

Por = –0.16 – –0.79

NCCH3 = –0.03 – –0.04

Page 151: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

151

Similarly, a polarization of the FeO spin density from almost equal oxygen and iron spin

densities in the reactants to dominant iron radical in TSCl,Z occurs, whereby Fe ranges from

1.43 – 1.64 and O ranges from 0.21 – 0.54. By contrast, Fe varies from 1.07 – 1.40 and O

ranges from 0.51 – 0.78 in the set of TSAN,Z structures. The axial ligand, therefore, has an

electronic effect on the charge and spin distributions in the rate determining transition states

and, in particular, an anionic axial ligand, such as chloride, polarizes the FeO biradical

towards the metal. Consequently, the anionic ligand will incur a “push”-effect on the metal-

oxo group and formally change it from an FeVO

–• configuration in the reactant state to

FeIV

O2–

in the transition state. By contrast, an iron(IV)-oxo oxidant with a neutral axial ligand

retains much larger radical character on the oxygen atom in the electrophilic transition states

and keeps the metal-oxo group in a formal FeVO

–• configuration. This is important as it

reduces the barrier heights of electrophilic reaction mechanisms.

Let us in the following look into the barrier heights of styrene epoxidation and elucidate the

enthalpic contributions to the relative values of the associated rate constants. Experimental

studies on substrate sulfoxidation by P450 enzymes indicated a correlation between rate

constant and ionization energy (IE) of the selected substrates (Watanabe, Y.; Iyanagi, T. et al

1980, Watanabe, Y.; Numata, T. et al 1981, Watanabe, Y.; Iyanagi, T. et al 1982). More

recent mass spectrometric and computational studies showed that this correlation also applies

to substrate epoxidation reactions (Lanucara, F.; Crestoni, M. 2011, Kumar, D.; Karamzadeh,

B. et al 2010). To test this relationship for the set of data studied here, we plot in Figure 5.8

the calculated barrier heights 2TSCl,Z and

2TSAN,Z as a function of the ionization energies of

the para-Z-styrene substrates. We calculated the trends using two sets of relative energies: In

the first we take the isolated reactants as a reference point, whereas in the second set a

reactant complex (RC) is used. DFT calculated ionization potentials at the same level of

Page 152: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

152

theory as the transition states for the epoxidation reactions were used for consistency. Our

calculated IE values are close to those reported in the literature (Lide, D. 1996), however, as

not all IE values are experimentally known, we will use the calculated data in our analysis

here. As can be seen both series of para-Z-styrene epoxidation reactions give barriers that

correlate linearly with ionization potential, whether energies relative to isolated reactants or a

reactant complex is used. Interestingly, the two correlations that use energies relative to

isolated reactants have different slope and intercept, and indeed the two curves cross at an

ionization energy of about 9.40 eV. Therefore, [FeIV

(O)(Por+•

)NCCH3]+ in the gas-phase is a

better oxidant than [FeIV

(O)(Por+•

)Cl] in epoxidation reactions with substrates with ionization

energy below 9.40 eV, whereas it is reversed for substrates with higher ionization potential,

i.e. for para-CN-styrene and para-NO2-styrene. This observation is in excellent agreement

with experimental studies of Nam et al (Song, W.; Ryu, Y. et al 2005) who also found

improved reactivity of [FeIV

(O)(Por+•

)NCCH3]+ over [Fe

IV(O)(Por

+•)Cl] for the epoxidation

of para-substituted styrenes with large +

p Hammett value of the substituent. The study that

uses a reactant complex as a reference point, by contrast, does not find this change in

reactivity but instead produces almost parallel trends (Figure 5.8b). It may very well be,

therefore, that the reactant complexes are unrealistic models and do not compare well with

experiment. Thus, in a reactant complex the structure of oxidant and substrate is solvated by

solvent molecules and the interactions between all particles. In our reactant complex no

solvent molecules were included and as a consequence full freedom of optimization was

possible, which may have resulted in an experimentally unrealistic structure.

Page 153: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

153

Figure 5.8: Styrene epoxidation barrier heights (E‡+ZPE) of

2TSX,Z plotted against the

ionization energy of the corresponding substrate. (a) Energies relative to isolated reactants.

(b) Energies relative to a reactant complex (RC).

To further understand the reactivity trends and the effect of substrate as well as oxidant on

the obtained barrier heights and, by extension, rate constants we set up a valence bond (VB)

curve crossing diagram for styrene epoxidation by iron(IV)-oxo porphyrin cation radical

systems, Figure 5.9. The VB curve crossing diagram displayed here is analogous to that

presented before for hydrogen atom abstraction reactions by iron(IV)-oxo oxidants (Shaik, S.

1981), but is further generalized to accommodate substrate epoxidation reactions.

This diagram starts on the bottom-left with the reactant configurations, i.e. styrene and

[FeIV

(O)(Por+•

)X]. The latter appears in two VB configurations, where we highlight the

valence -orbital electrons with dots. Thus, the electronic ground state of [FeIV

(O)(Por+•

)X] is

x2–y22 xz

2 yz

2 *xz

1 *yz

1 a2u

1 in both the quartet and doublet spin states and its wave

y = 1.33x - 4.94

R² = 0.88

y = 7.52x - 63.15

R² = 0.98

-15

-10

-5

0

5

10

4 5 6 7 8 9 10

Cl

ANE‡+ZPE

[kcal mol–1]

IE [eV]

(a)

y = 2.11x - 16.39

R² = 0.905

y = 2.42x - 13.97

R² = 0.96

-5

0

5

10

15

6 7 8 9 10

AN

ClE‡+ZPE

[kcal mol–1]

IE [eV]

(b)

Page 154: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

154

function is labelled as r in Figure 5.9. The x2–y2 orbital is a nonbonding orbital in the plane

of the porphyrin ring. The perpendicular set of orbitals, xz/*xz and yz/*yz, are the bonding

and antibonding combinations of the metal 3dxz/3dyz with the oxygen 2px/2py atomic orbitals

and contain three electrons in each pair.

Figure 5.9: Valence bond curve crossing diagram for para-Z-styrene epoxidation by

iron(IV)-oxo porphyrin cation radical oxidants. Lewis structures give relevant valence-

orbitals with a dot.

In the C–O bond formation step in the epoxidation mechanism, one electron is transferred

from the substrate to the oxidant and it fills the a2u orbital with a second electron to give the

radical intermediate wave function I. In VB theory (Shaik, S. 1981), the reactant (r) and

product (I) wave functions cross each other and connect to excited states in the product and

reactants conformations, respectively. Thus, I* is an excited reactant state with an electronic

Page 155: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

155

configuration that represents a C–O bond pair, a radical on the styrene group and a closed

shell a2u orbital. The two VB curves give an avoided crossing and a barrier height for C–O

bond formation, E‡. Based on a series of hydrogen abstraction reactions (Shaik, S. 1981,

Shaik, S.; Kumar, D. et al 2008, Latifi, R.; Bagherzadeh, L. et al 2009) it was shown that E‡

is proportional to the curve crossing energy (Ec) minus the resonance energy B via E‡ =

Ec – B. However, the curve crossing energy is proportional to the promotion gap (GH,r) or

excitation energy from the reactant to product configuration in the geometry of the reactants

so that the barrier height can be written as:

E‡ = f GH,r – B (1)

In a recent study we showed that the relationship in Equation 1 is also valid for substrate

epoxidation and sulfoxidation reactions and that the promotion gap GH,r reflects the excitation

energy or ionization potential of the substrate (Kumar, D.; Karamzadeh, B. et al 2010,

Kumar, D. Sastry, G. et al 2011).

A comparison of the VB structures for the r and I* configurations show that there is an

electron transfer from the iron(IV)-oxo group into the porphyrin a2u orbital reflecting an

electron excitation in CpdI (Eex,CpdI), which can be approximated with the electron affinity of

the oxidant. In addition, it can be seen that the electrons in the C=C bond are singlet paired in

the reactant ground state but triplet coupled in the excited state in the reactant geometry,

which implies a singlet-triplet energy gap in the C=C bond, EST. Thus, a singlet-triplet

energy gap in the C=C bond reflects the excitation energy of an electron in the -bond of the

C=C moiety, namely Eex,Z, which in its turn can be approximated with the ionization energy

of the substrate. Consequently, GH,r is proportional to EACpdI(X) + Eex,Z and so should the

barrier height of the epoxidation reaction. Of course, electron excitation from the -orbital of

Page 156: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

156

the substrate, which is the HOMO, is also proportional to the ionization energy of the

substrate. Indeed, the plot in Figure 5.8 confirms a relationship between barrier height and

ionization potential as predicted by the VB diagram in Figure 5.9.

For the reverse reaction, that is from radical intermediates to reactants, the epoxidation

barrier height (E‡

rev) is equal to E‡ plus the exothermicity to form radical intermediates

(Eri). In VB the reverse barrier is proportional to the promotion gap in the radical

intermediates, i.e. I to p* or GH,p. The VB structures on the radical intermediate side of

Figure 5.9 show that the excitation energy GH,p corresponds to the formation energy of the C–

O bond plus the electron transfer from the substrate into the a2u orbital. The VB structures in

Figure 5.9 indicate that the electrons in the C–O bond are singlet coupled in I but triplet

coupled in p*, which implies that a singlet-triplet excitation in the C–O bond has occurred.

The singlet-triplet excitation in the C–O bond refers to the bond breaking of the C–O bond. In

earlier work we showed that the C–O bond formation energy is proportional to a H–O bond

formation energy, so that it can be mimicked with BDEOH as defined as the reaction enthalpy

for Equation 3 (Kumar, D.; Karamzadeh, B. et al 2010).

To test whether the VB curve crossing diagram can predict barrier heights from empirical

values, we calculated the gas-phase electron affinities of all four oxidants from Figure 5.1 and

the -* excitation energies and ionization energies of all para-Z-styrene substrates and the

results are given in Table 4.1. We calculated the electron affinity of the oxidant, the -*

excitation energy in the para-Z-styrene substrates and the promotion gap via Equation 2.

Subsequently, we used these values to estimate E‡

VB for all substrates. Although these VB

calculated barrier heights are on average within –0.05 kcal mol–1

of the DFT calculated ones,

actually the standard deviation is quite large (about 2 kcal mol–1

). This implies that the model

has short-comings, which we address with a slightly modified model as described below.

Page 157: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

157

GH,r = 2 (EACpdI(X) + Eex,Z) (2)

[FeIV

(O)(Por+•

)X] + H• → [Fe

IV(OH)(Por)X] (3)

Thus, the forward reaction barrier is dependent on the intrinsic chemical properties of the

substrate, i.e. the ionization energy, whereas the reverse reaction barrier depends on the

variables of the oxidant that is the BDEOH value. We calculated gas-phase values of BDEOH =

87.5 kcal mol–1

for [FeIV

(O)(Por+•

)Cl] and BDEOH = 82.8 for [FeIV

(O)(Por+•

)NCCH3]+,

respectively (de Visser, S.; Tahsini, L. et al 2009). This energy difference narrows to almost

equal values in a dielectric constant of = 5.7: BDEOH = 82.3 and 81.3 kcal mol–1

for

[FeIV

(O)(Por+•

)X], X = Cl– or NCCH3, respectively.

The VB diagram of Figure 5.9 shows that the transition state described from reactants to

intermediates, i.e. forward, is proportional to EACpdI(X) + Eex,Z, whereas the transition state in

the reverse reaction, i.e. backwards, is proportional to the BDEOH value of the oxidant. Thus,

the location of the barrier on the potential energy surface will determine whether the barrier

correlates with EACpdI(X) +Eex,Z or BDEOH. So, for a substrate epoxidation reaction where the

transition state has a very reactant-like electronic configuration, i.e. an early transition state,

very little electron transfer from substrate to oxidant has taken place and consequently the

reactant wave function dominates in the transition state. As a result, an early transition state

should be proportional to BDEOH. On the other hand, a late transition state has an electronic

configuration closely resembling the radical intermediate state due to a significant amount of

electron transfer that has taken place already. Therefore, late transition states, based on the

VB diagram in Figure 5.8 above should correlate with the electron affinity of the oxidant.

Indeed the group spin density of the epoxidation transition state with para-NO2-styrene as a

substrate shows a only a small amount of electron transfer from substrate to oxidant and

consequently the charge transfer from substrate to oxidant (QCT) is almost zero. This implies

Page 158: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

158

that the epoxidation barriers for para-NO2-styrene will be early on the potential energy

surface and little electron transfer has taken place. As reasoned above, these barrier heights

should, therefore, correlate with BDEOH and not with EACpdI(X) + Eex,Z. To test this we display

in Figure 5.10 the correlation of para-NO2-styrene epoxidation with BDEOH for 1X/2X (X =

Cl– and NCCH3). As can be seen for this set of four data points a linear correlation is found

between the epoxidation barrier of para-NO2-styrene and the corresponding BDEOH value of

the oxidant.

Figure 5.10: Correlations of (a) Epoxidation barrier height of para-NO2-styrene with

BDEOH. (b) Epoxidation barrier height of para-N(CH3)2-styrene with EACpdI(X).

The group spin densities of 2TSX,N(CH3)2, by contrast to those found for

2TSX,NO2 show

considerably larger amount of spin density on the substrate part of the structure. At the same

time the spin density on the porphyrin ring has dropped due to electron transfer from

substrate to oxidant. Thus, the 2TSX,N(CH3)2 structures are electronically late and resemble

0

2

4

6

8

10

80 85 90 95 100

E‡+ZPE

[kcal mol–1]

BDEOH [kcal mol–1]

Z = NO2(a)

-25

-20

-15

-10

-5

0

5

10

75 95 115 135 155 175

E‡+ZPE

[kcal mol–1]

EA [kcal mol–1]

Z = N(CH3)2

(b)

Page 159: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

159

product type conformations. Because of that, in 2TSX,N(CH3)2 the electron transfer has taken

place already (or at least most of it) and, consequently, the barrier height will be driven by the

differences in electron affinity of the individual oxidants. To test this, we plot in Figure

5.10(b) the correlation between epoxidation barrier height of para-N(CH3)2-styrene

epoxidation by 1X/2X oxidants as a function of the electron affinity of the oxidant. We find a

linear correlation between electron affinity and barrier height as predicted by the VB diagram

in Figure 5.8.

As follows from the correlations depicted in Figure 5.10, the amount of electron transfer

from substrate to oxidant determines whether a transition state correlates with either the sum

of EACpdI(X) + Eex,Z or with BDEOH instead. Thus, we extracted the degree of charge transfer

from substrate to oxidant (QCT) from the UB3LYP/B2//UB3LYP/B1 calculations and

summarize these values for the 2TSX,Z transition states in Table 5.1. We now define the

variable that describes the position of the transition state on the potential energy profile

through Equation 4 and link to the degree of charge transfer (QCT), BDEOH and EACpdI(X) +

Eex,Z.

= QCT BDEOH + (1 – QCT) (EACpdI(X) + Eex,Z) (4)

Essentially, is proportional to BDEOH when the charge-transfer is maximal, i.e. an early

transition state, but it is proportional to EACpdI(X) + Eex,Z for a late transition state with a small

value of QCT. We calculated for all transition states 2TSX,Z, X = Cl

–/NCCH3 and para-Z-

styrene as substrates, and a plot of all data is given in Figure 5.11. Thus, the full set of

transition states, irrespective of the axial ligand of the oxidant, fits a linear correlation.

Consequently, we described the trend in epoxidation barriers for early as well as late

Page 160: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

160

transition states as a function of intrinsic chemical properties of the oxidant and substrate and

the amount of charge transfer in the TS during the reaction process.

Figure 5.11: Correlation between epoxidation barrier height of all data for 1X (X = Cl–,

NCCH3) with parameter

Table 5.1. Substrate chemical properties and charge-transfer (QCT) values in the

transition states.

Z IEZ a Eex,Z

a QCT,Cl

b QCT,AN

c

H 9.05 3.82 0.10 0.25

F 9.05 3.81 0.07 0.24

Cl 8.87 3.70 0.08 0.16

CH3 8.59 3.75 0.12 0.32

t-Bu 8.43 3.76 0.12 0.42

CN 9.46 3.46 0.01 0.10

NO2 9.70 2.88 –0.03 –0.01

OCH3 8.03 3.73 0.16 0.42

NH2 7.55 3.61 0.22 0.55

N(CH3)2 7.05 3.57 0.27 0.63

a In eV.

b Charge transfer in

2TSCl,Z.

c Charge transfer in

2TSAN,Z.

y = 2.47x + 136.46

R² = 0.86

0

50

100

150

200

-15 -10 -5 0 5 10 15E‡+ZPE [kcal mol–1]

[kcal mol–1]

Page 161: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

161

In summary, we show here that the barrier height of an epoxidation reaction by a metal-oxo

oxidant is dependent on the electron donating/withdrawing character of the substituents of the

substrate. Thus, substrates with high electron donating power such as para-N(CH3)2-styrene

result in low barrier and rapid electron transfer from substrate to oxidant well before the

epoxidation barrier height has overcome. On the other hand, electron withdrawing groups

such as para-NO2-styrene result in much later electron transfer from substrate to oxidant and

now the height of the barrier is determined by the strength of the C–O bond that is formed.

5.5 Conclusion

In this work we report a systematic computational study into substrate epoxidation by four

iron(IV)-oxo porphyrin cation radical models. We investigated the effect of the axial ligand,

chloride versus acetonitrile, as well as the equatorial ligand, Por versus TPFPP. It is shown

that the substituents on the porphyrin ring can guide substrate binding through electrostatic

interactions with halide atoms, which lowers the barrier heights. A neutral axial ligand leads

to displacement of the metal from the plane through the porphyrin ring and results in different

orbital interactions between metal and porphyrin ring as compared to systems with an anionic

ligand. This has profound effects on the electron affinity of the oxidant and the subsequent

reactivity patterns. Finally, we investigated para-Z-styrene epoxidation by four iron(IV)-oxo

porphyrin cation radical models and highlight the differences in reactivity for olefins with

electron withdrawing versus electron donating substituents. In particular, it is shown that

electron donating substituents give early transition states and lower reaction barriers than

substrates with electron withdrawing substituents.

Page 162: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

162

CHAPTER 6

PROJECT FOUR

Page 163: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

163

Drug metabolism by cytochrome P450 enzymes:

What distinguishes the pathways leading to

substrate hydroxylation over desaturation?4

ABSTRACT

Cytochrome P450 enzymes are highly versatile biological catalysts in our body that react

with a broad range of substrates. Key functions in the liver include the metabolism of drugs

and xenobiotics. However, these processes often lead to reactive (toxic) metabolites that are

harmful. Understanding the biodegradation of molecules by P450 enzymes is therefore

important for pharmaceutical purposes and may assist in developing better medication. One

particular metabolic pathway that is poorly understood relates to the P450 activation of

aliphatic groups leading to either hydroxylation or desaturation pathways. We have done the

first combined density functional theory and quantum mechanics/molecular mechanics study

on the factors that determine the regioselectivity of aliphatic hydroxylation over desaturation

of compounds by P450 isozymes. The calculations establish multistate reactivity patterns on

competing doublet and quartet spin states, whereby the product distributions differ on each of

the spin state surfaces; hence we find spin-selective product formation. We analyzed the

electronic and thermochemical factors that determine the bifurcation pathways and have

established a model that predicts the regioselectivity of aliphatic hydroxylation over

4 Li Ji, Abayomi S. Faponle, Matthew G. Quesne, Mala A. Sainna, Jing Zhang, Alicja Franke, Rudi van Eldik, Weiping Liu and Sam P. de

Visser. “Drug metabolism by cytochrome P450 enzymes: What distinguishes the pathways leading to substrate hydroxylation over

desaturation?”Chem. Eur. J. 2015, 21, 1-11.

Page 164: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

164

desaturation pathways from valence bond and molecular orbital theories. Thus, the difference

in energy of the C–H bond orbital versus that of the substrate radical orbital in the radical

intermediate determines the degree of desaturation products. In addition, we have identified

environmental effects of the substrate binding pocket that affect the regioselectivities. These

studies imply that bioengineering P450 isozymes for desaturation reactions will have to

include modifications in the substrate binding pocket to restrict the hydroxylation rebound

reaction.

6.1 Introduction

The cytochrome P450s (P450s) are key drug metabolism enzymes in the human body

(Sono, M.; Roach, M. P.; et al 1996, Groves, J. T. 2003, Ortiz de Montellano, P. R., 2004,

Denisov, I. G.; Makris, T. M.; et al 2005, Kadish, K. M.; Smith, K. M.; et al 2010, Ortiz de

Montellano, P. R. 2010), and are widely distributed in living organisms, such as eukaryotes,

but also in plants, fungi and bacteria. As of January 2014, more than 20,000 different

sequences had been identified, making the P450s one of the major biological enzyme classes

(Nelson, D. R. 2009). In the human body, P450s are predominantly found in the liver, where

they catalyze the biodegradation of xenobiotics but are also involved in the biosynthesis of

hormones (Guengerich, F. P. 2001, Posner, G. H.; O’Neill, P. M. 2004, Munro, A. W.;

Girvan, H. M.; et al 2007). As such the P450s have multiple functions in the body, which also

shows in their versatility on substrate activation. For instance, the P450BM3 subclass

hydroxylates long-chain fatty acids (Gelb, M. H.; Heimbrook, D. C.; et al 1982, Atkins, W.

M.; Sligar, S. G. 1987), and has a substrate binding pocket that accommodates long and thin

substrates. By contrast, P450cam is a bacterial isozyme that regioselectively hydroxylates

camphor at the carbon-5 position and has a fit-for-purpose substrate binding pocket

(Ruettinger, R. T.; Wen, L. P.; et al 1989, Davydov, D. R.; Bon Hoa, G. H.; et al 1999). Key

drug metabolizing P450 isozymes include P4502D6, which has a more open substrate

Page 165: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

165

binding pocket and can activate a large range of different substrate shapes and sizes. As P450

isozymes are highly efficient in substrate activation, they have gained interest from the

biotechnology industry for the regioselective and/or stereospecific activation of substrates

(O'Reilly, E.; Koehler, V.; et al 2011, Grogan, G. 2011). Despite decades of extensive

scientific research into P450 chemistry, there are still major gaps in our understanding of its

reactivity and activity, in particular related to the metabolism of drug molecules and

xenobiotics, which is important for drug development and the pharmaceutical industry.

The name cytochrome P450 derives from the fact that the CO bound heme complex has an

absorption band at 450 nm (Omura, T.; Sato, R. 1962). Their heme active site is linked to the

protein via a thiolate linkage of a cysteinate residue covalently bound to the iron center

(Poulos, T. L.; Finzel, B. C.; et al 1985, Schlichting, I.; Berendzen, J.; et al 2000, Auclair, K.;

Moënne-Loccoz, P.; et al 2001), while the substrate binding pocket varies in shape and size

for the different P450 isozymes. As an example we show in Figure 5.1 the active site of

P4502C19 with substrate warfarin bound as taken from the 4GQS protein databank (pdb) file

(Reynald, R. L.; Sansen, S.; et al 2012). The heme is linked to the protein through an iron-

thiolate linkage with Cys435 and trans to this ligand the substrate (warfarin) binds.

Page 166: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

166

Figure 6.1: Extract of the active site of P4502C19 as taken from the 4GQS pdb file.

The active species of P450 enzymes is an iron(IV)-oxo heme cation radical species called

Compound I (Cpd I) (Rittle, J.; Green, M. T. 2010). Due to its transient nature, Cpd I is short-

lived and difficult to study with experimental means and as a consequence many details of its

catalytic mechanism with substrates are unknown. The most common reaction mechanism

performed by the P450s is aliphatic hydroxylation of substrates (Groves, J. T. 2005,

Watanabe, Y.; Nakajima, H.; et al 2007), which was experimentally revealed as a stepwise

process leading to alcohol product complexes (Groves, J. T.; McClusky, G. A. 1976).

Computational modelling confirmed the general mechanism and showed that the oxidation

heme

Cys435

warfarin

Page 167: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

167

actually contains two close-lying spin states that lead to two-state reactivity patterns, each

with an individual rate constant of the reaction (Kamachi, T.; Yoshizawa, K. 2003, Shaik, S.;

Kumar, D.; et al 2005). Moreover, the two-state reactivity pattern also predicted that the

lifetime of the radical intermediate could lead to a bifurcation of the potential energy profile,

thereby generating, for instance, rearrangement patterns on one spin-state surface but not on

the other (De Visser, S. P.; Ogliaro, F.; et al 2001).

A key reaction mechanism in P450 chemistry with big relevance to drug metabolism is the

desaturation of aliphatic groups to form olefins. As shown in Scheme 6.1 desaturation is a

reaction mechanism that shows similarities to aliphatic hydroxylation, whereby both

processes start with an initial hydrogen atom abstraction. However, thereafter the reaction

bifurcates into two possible product channels: one giving OH rebound to form alcohol

products (aliphatic hydroxylation) and the other leading to a second hydrogen atom

abstraction to give an olefin and water (desaturation). The origin and consequences of the

bifurcation pathways in P450 catalysis are currently unknown and no predictive model exists

that gives product distributions as a function of substrate chemical properties.

Several drug molecules have been shown to produce desaturation metabolites after a

reaction with P450 enzymes. Namely, valproic acid, an antiepileptic drug, is converted by

P450 enzymes to form both alcohol and olefin products (Rettie, A. E.; Rettenmeier, A. W.; et

al 1987, Rettie, A. E.; Boberg, M.; et al 1988, Sadeque, A. J. M.; Fisher, M. B.; et al 1997,

Wen, X.; Wang, J.-S.; et al 2001, Gunes, A.; Bilir, E.; et al 2007). Another example is

ethylcarbamate (Loch, J. M.; Potter, J.; et al 1995, Forkert, P.-G.; Lee, R. P. 1997, Lee, R. P.;

Parkinson, A.; et al 1998, Guengerich, F. P. 2008), that in a reaction with P4502E1 enzymes

leads to a small amount of vinyl carbamate as a precursor to its epoxide (Guengerich, F. P.;

Kim, D. H. 1991). The latter is considered the ultimate carcinogen (Leithauser, M. T.; Liem,

A.; 1990) and its biosynthesis requires an essential desaturation step. Over the years a range

Page 168: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

168

of substrates of P450 isozymes have been identified that are desaturated to form olefins,

including lauric acid (Guan, X.; Fisher, M. B.; et al 1998), but also a step in the biosynthesis

of ergosterol includes a desaturation (Kelly, S. L.; Lamb, D. C.; 1997).

Clearly, the mechanistic pathways surrounding drug metabolism by the P450s produce a

large versatility of product patterns and currently it is not clear why certain substrates lead to

desaturation, whereas others solely give hydroxylation products. To gain insight into the

origin of the regioselectivity of desaturation versus hydroxylation processes by P450

enzymes, we decided to do a comprehensive computational study that includes density

functional theory (DFT) modelling, quantum mechanics/molecular mechanics (QM/MM) and

valence bond (VB) theory. The overall study gives an electronic structure description on the

metabolism of a variety of drug molecules and enables us to provide a predictive pattern that

anticipates desaturation versus hydroxylation product distributions.

Scheme 6.1: (a) Competitive hydroxylation and desaturation metabolism pathways of drug

molecules by P450 Cpd I. (b) Reaction products observed for valproic acid and

ethylcarbamate.

Page 169: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

169

Our studies explain the experimentally observed bifurcation reactions and may be used to

develop computer-aided toxic risk assessment of potential carcinogens resulting from

desaturation reactions catalyzed by P450 isozymes.

6.2 Methods

Our work uses two main computational approaches, namely DFT on model complexes and

QM/MM on a full P450 isozyme.

6.2.1. DFT model calculations

As in previous studies of our groups (de Visser, S. P.; Tan, L.-S. 2008, Ji, L.; Zhang, J.; et

al 2014), Cpd I of P450 was modeled as a six-coordinated iron(IV)-oxo porphyrin (Por)

cation radical complex, [FeIV

(O)(Por+

)(SH)]. For simplicity, the heme is replaced by

protoporphyrin IX, whereby all side chains are abbreviated with hydrogen atoms and the

axial cysteinate ligand is changed to thiolate. All geometries were optimized using the

unrestricted hybrid B3LYP density functional (Becke, A. D. 1993, Lee, C.; Yang, W.; et al

1988) in combination with an LACVP basis set on iron and 6-31G on the rest of the atoms

(Hay, P. J.; Wadt, W. R. 1985, Hehre, W. J.; Ditchfield, K.; et al 1972); basis set BSI. A

subsequent analytical frequency calculation was run to confirm that all ground states had no

imaginary frequencies, and all transition states had a single one for the correct mode. The

computed vibrational frequencies were then used further for quantifying the zero-point

energy (ZPE), and enabled us to calculate thermal and entropic corrections to the free energy

at 298.15 K and 101.325 kPa. More accurate energies were determined by single-point

calculations with the SDD basis set on iron coupled to 6-311++G** on the rest of the atoms;

basis set BSII. To test the effect of the basis set on the optimized geometries, we calculated a

full potential energy profile of 1-butene epoxidation by Cpd I using different basis sets.

However, a single point calculation with a triple- quality basis set on all atoms performed on

a geometry that was optimized with either a double- or a triple- quality basis set did only

Page 170: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

170

give minor differences in the relative energies to within 0.1 kcal mol–1

of each other (Kumar,

D.; Karamzadeh, B.; 2010), hence the former procedure was used in this work.

Bulk protein environment was simulated with the polarized continuum-solvation model

with a dielectric constant of = 5.62. Dispersion interactions were added to the energy at the

UB3LYP-D3 level of theory with zero damping (Grimme, S. 2006). All calculations were

carried out using the Gaussian 09 program package (Frisch, M. J. 2010).

As density functional theory occasionally struggles with the correct description of spin

state ordering of transition metal complexes (de Visser, S. P.; Quesne, M. G.; et al 2014), we

tested the effect of the density functional method on the spin state ordering and relative

energies of the rebound versus desaturation reactions. We did full geometry optimizations at

the UBLYP/BSI (Lee, C.; Yang, W.; et al 1988, Becke, A. D. 1988), UB3PW91/BSI (Becke,

A. D. 1993, Perdew, J. P.; 1992), and TPSS/BSI (Tao, J.; Perdew, J. P.; et al 2003) levels of

theory. However, no change in spin state ordering was observed and only minor differences

in the relative energies of the barrier heights were obtained. Therefore, the same qualitative

picture is obtained with alternative DFT methods and we will focus on the B3LYP results

only.

6.2.2. QM/MM calculations

To support the DFT calculations and gain insight into whether the model would be

applicable to realistic systems, we did a further set of calculations using an actual P450

isozyme that is known to desaturate its natural substrate. OleTJE was recently identified as a

P450 isozyme that converts long chain fatty acids into terminal olefins (Rude, M. A.; Baron,

T. S.; et al 2011). We investigated the pathways of -hydroxylation and desaturation of

eicosanoic acid by the P450 isozyme OleTJE using QM/MM methods. The starting

coordinates for the calculations were taken from crystal structure data (Belcher, J.; McLean,

K. J.; et al 2014). We followed the set-up procedures of the QM/MM system as reported and

Page 171: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

171

benchmarked before, which we will briefly summarize here (Kumar, D.; Thiel, W.; et al

2011, Quesne, M. G.; Latifi, R.; et al 2014).

Starting from the crystal structure coordinates, hydrogen atoms and oxo group were added

to the structure using the PDB2PQR program package (Dolinsky, T. J.; Czodrowski, P.; et al

2007) and the active site was manually modified from a resting state iron(III)-water(heme)

complex into an iron(IV)-oxo heme cation radical (Cpd I) structure. We did a careful analysis

of all acidic and basic amino acid side chains for their protonation states, whereby we chose

all glutamic and aspartic acid residues to be deprotonated at the carboxylic acid group and all

arginine and lysine residues protonated. All histidine groups were visually inspected for

proton donor and acceptor possibilities. The histidine residues His120 and His210 were doubly

protonated, whereas the other six (His85, His92, His222, His259, His325 and His363) were singly

protonated. The overall structure was neutralized with counter-ions (Mg2+

and Cl–). Solvent

water (with sphere of radius of 35 Å) was added to this structure, and equilibrated, followed

by a molecular dynamics minimization and heating procedure to 298 K of the full structure

using the CHARMM force field (Brooks, B. R.; Bruccoleri, R. E.; et al 1983). The thus

obtained QM/MM model had a total number of 32,992 atoms including 8,739 TIP3P water

molecules.

Three snapshots from the MD simulation were selected as starting points for the QM/MM

calculations at time intervals of 300, 400 and 500 ps (Sn300, Sn400 and Sn500). In these

QM/MM calculations the inner core including all residues and water molecules within a

radius of 8 Å from the heme iron were fully geometry optimized, whereas all other atoms

were fixed in the MD simulated coordinates. The iron(III)-hydroxo species was geometry

optimized in the doublet and quartet spin states and a frequency calculation on the QM region

established these structures as stationary points.

Page 172: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

172

QM/MM calculations employed the Turbomole program package (Ahlrichs, R.; Bär, M.;

et al 1986) for the calculations of the QM region and Charmm as implemented in DL-Poly

(Smith, W.; Forester, T. R. 1996) for the MM region all linked through the ChemShell

interface (Sherwood, P. et al 2003). For consistency with the DFT studies reported above, we

selected the UB3LYP (Becke, A. D. 1993, Lee, C.; Yang, W.; et al 1988) density functional

for the QM region in combination with the Turbomole built-in basis set SVP, BSIII

(Ahlrichs, R.; Bär, M.; et al 1989). The QM region contained the iron(IV)-oxo group, the

heme (without side chains), thiolate for the axial cysteinate group and the butanoic acid group

of the substrate. To improve the energetics, single point calculations were performed with a

Wachters all electron basis set on iron and 6-311+G** on the rest of the atoms; basis set

BSIV (Wachters, A. J. H. 1970). Electronic embedding procedures were implemented

(Bakowies, D.; Thiel, W. 1996), thereby taking into consideration to electronic interactions of

electrons in the QM region with charges in the MM region.

Scheme 6.2 displays the QM region selected in the QM/MM calculations. As the substrate

has a terminal carboxylate group that is linked through a salt bridge with an arginine residue,

we kept the methylguanidinium component of Arg245 in the QM region and terminated the

substrate at the sixth carbon atom. All side chains of the heme were in the MM region and

cysteinate was abbreviated by methylmercaptane. Subsequently, we investigated hydrogen

atom abstraction from the C position of the substrate and the pathways leading to

desaturation and hydroxylation.

Page 173: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

173

Scheme 6.2: Atoms in the QM region of the QM/MM calculation. Wiggly lines represent the

cuts between the QM and MM regions.

6.3 Results and Discussion

6.3.1. DFT model reactions

Before going into detail of the mechanism of drug metabolism by P450 enzymes, let us

first go through a typical reaction mechanism of substrate activation leading to desaturation

and hydroxylation products by taking ethylcarbamate (EC) as an example. Scheme 6.3

displays the calculated reaction mechanisms of EC activation by Cpd I of P450 and the

bifurcation mechanism leading to desaturation and hydroxylation products.

Scheme 6.3: Reaction Mechanism of Ethyl Carbamate Activation by Cpd I of P450.

Page 174: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

174

Figure 6.2: Potential energy profile of ethylcarbamate activation by 4,2

Cpd I of P450 as

calculated with DFT. Energies are given in kcal mol–1

and are calculated at

UB3LYP/BSII//UB3LYP/BSI level of theory with ZPE and solvent corrections included.

Values in parenthesis are free energies in solvent. Optimized geometries give bond lengths in

angstroms and the imaginary frequency in the transition states in wave numbers.

We calculated the reaction between Cpd I and ethylcarbamate, as shown in Figure 6.2,

starting from a reactant complex (4,2

REC), which exists in close-lying doublet and quartet spin

states, both with electronic configuration x2–y22 *xz

1 *yz

1 a2u

1. Attack of the oxo group of

Cpd I on the substrate leads to a hydrogen atom abstraction via a transition state 4,2

TSH,EC to

form a radical intermediate (4,2

IH,EC) with electronic configuration x2–y22 *xz

1 *yz

1 a2u

2

Rad1. During the lifetime of the radical intermediate it can isomerize to structures

4,2ID,EC,

whereby the C–H group of the substrate forms a hydrogen bonding interaction with the

oxygen atom of the OH group as a precursor to the second hydrogen atom abstraction step via

4TSH,EC [2TSH,EC]

i1179 [i1208] cm–1

rCH = 1.471 [1.370]rOH = 1.112 [1.157]

4ID,EC [2ID,EC]

4IR,EC [2IR,EC]4IH,EC [2IH,EC]

rHO = 2.420 [2.410]rOH = 0.982 [0.982]

rCH = 2.588 [2.479]rOH = 0.984 [0.985]

rCO = 3.565 [3.568]rOH = 0.982 [0.982]

Page 175: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

175

transition state 4,2

TSD,EC to form desaturated products 4,2

PD,EC. The alternative mechanism

from 4,2

IH,EC gives an isomerization to form the pro-hydroxylation radical intermediate

4,2IR,EC, where the radical is aligned to the oxygen atom for rebound via transition state

4,2TSreb,EC to give the alcohol product complexes

4,2PA,EC. Our calculations were focused on

establishing whether isomerization barriers exist from IH to ID and IR, and what the factors

are that determine the height of these barriers and consequently the regioselectivity of the

reaction.

As shown in Figure 6.2, the hydrogen atom abstraction step is endergonic by G = 19.4

(17.9) kcal mol–1

in the quartet (doublet) spin states, respectively. Addition of dispersion

correction to the calculated relative energy lowers 4,2

TSH,EC in energy by about 4 kcal mol–1

,

which is in accord with analogous previous findings for transition metal complexes

(Lonsdale, R.; Harvey, J. N.; et al 2010). The hydrogen atom abstraction transition states are

characterized by an almost linear O--H--C configuration and a large imaginary frequency

(HS: i1179 cm–1

; LS: i1208 cm–1

), and consequently a considerable kinetic isotope effect

(KIE) for the reaction may be expected (De Visser, S. P. 2006).

We then investigated the OH rebound to form alcohol products and the second hydrogen

atom abstraction to give desaturation products. We propose this to happen (Scheme 6.2) via

an initial isomerization of the radical intermediate from 4,2

IH,EC to 4,2

IR,EC, which sets up the

molecular orientation for the OH rebound leading to 2-hydroxyethyl-carbamate product.

Similarly to previous studies (Shaik, S.; Kumar, D.; et al 2008), a small but significant

rebound barrier (G = 4.9 kcal mol–1

) is found on the quartet spin surface, while the doublet

spin pathway is barrier free. Dispersion corrections lower the rebound barriers by about 1

kcal mol–1

, which is a significant lesser change than that found for the hydrogen atom

abstraction step. The radical rebound pathway from 4,2

IH,EC is exergonic by G = 58.8 (55.8)

kcal mol–1

on the quartet (doublet) spin state surfaces as expected for the formation of a

Page 176: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

176

stable alcohol product complex. The alternative isomerization pathway from 4,2

IH,EC, leads to

a substrate rotation to form 4,2

ID,EC, whereby the oxygen atom of the OH group is aligned

with a hydrogen atom of substrate to enable a second hydrogen atom abstraction. Note that

the difference in orientation between 4,2

IR,EC and 4,2

ID,EC are similar to those reported for a

synthetic manganese-hydroxo porphyrin complex (Hull, J. F.; Balcells, D.; et al 2010). The

doublet spin radical intermediate 2ID,EC collapses to the desaturation product complex

virtually barrier-free, whereas the quartet spin complex 4ID,EC encounters a large activation

barrier (4TSreb,EC) of G = 17.8 kcal mol

–1 instead. Similarly to the above mentioned rebound

process, the activation barrier only decreases by 1.5 kcal mol–1

after dispersion corrections

are added. The large 4TSD,EC barrier implicates that desaturation is an unlikely process on the

quartet spin state surface. Moreover, the difference in kinetics between the quartet and

doublet spin desaturation routes suggests that the doublet spin state surface is the only viable

pathway responsible for vinyl carbamate formation. Consequently, the regioselectivity of

desaturation versus hydroxylation is a spin-selective reaction process, whereby the high-spin

pathway gives solely hydroxylation products whereas a mixture of products may be expected

on the low-spin surface instead.Experimental reports on ethylcarbamate activation by P450

isozymes reveal only a small proportion of biotransformation leading to desaturation

reactions (Guengerich, F. P.; Kim, D. H. 1991). The DFT calculations presented in Figure

6.2, therefore, are in line with these experimental product distributions of hydroxylation

versus desaturation, with a small thermodynamic and kinetic preference in favor of substrate

hydroxylation. Moreover, when we look at the driving force for desaturation versus

hydroxylation, both pathways on the low-spin surface have similar exergonicity, i.e. G =

−48.0 vs −47.7 kcal mol–1

. As such, P450s should exhibit mixed hydroxylase/desaturase

activity toward EC as indeed observed experimentally. To find a general trend in desaturation

versus hydroxylation regioselectivity preferences, we did a further set of calculations using

Page 177: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

177

alternative substrates that should give the chemical extremes of the two processes. In Figure

6.3 we display studies on desaturation versus hydroxylation mechanisms using valproic acid

(VA), ethane (ET) and dihydroanthracene (DHA) as substrates. Thus, experimentally, DHA

in a reaction with P450 enzymes is known to lead exclusively to anthracene products,

whereas ethane is leads solely to alcohol products instead. For all chemical systems the full

potential energy profile like the one shown in Figure 6.2 is calculated on both spin state

surfaces and mechanisms are analogous to those reported above. Thus, for valproic acid as a

substrate the DFT calculations predict mixed desaturase/hydroxylase activity on a low-spin

surface. Key optimized geometries of the hydrogen atom abstraction transition states are

given in Figure 6.3. In general, all substrates reveal the same catalytic mechanism as

calculated above and previously reported in the literature (Kamachi, T.; Yoshizawa, K. 2003,

Shaik, S.; Kumar, D.; et al 2005, Ogliaro, F.; Harris, N.; et al 2000, de Visser, S. P.; Kumar,

D.; et al 2004, Ji, L.; Schüürmann, G. 2013) with a rate determining hydrogen atom

abstraction leading to a radical intermediate. Differences, however, are found for the relative

energies for radical rebound versus desaturation.

Figure 6.3: Optimized geometries of hydrogen atom abstraction transition states of VA, ET

and DHA by 4,2

Cpd I of P450 as calculated with DFT. Bond lengths are in angstroms, the

imaginary frequency is in wave numbers and (free) energies are given in kcal mol–1

and are

4TSH,VA [2TSH,VA]

i1560 [i1273] cm–1

rCH = 1.358 [1.248]rOH = 1.184 [1.284]

4TSH,ET [2TSH,ET] 4TSH,DHA [2TSH,DHA]

i1683 [i1099] cm–1

rCH = 1.402 [1.322]rOH = 1.150 [1.193]

i1505 [i1655] cm–1

rCH = 1.282 [1.246]rOH = 1.300 [1.346]

E+ZPE = 13.1 (12.1)G = 15.0 (14.4)

E+ZPE = 15.8 (13.7)G = 19.0 (17.1)

E+ZPE = 7.8 (6.9)G = 10.0 (9.3)

Page 178: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

178

calculated at UB3LYP/BSII//UB3LYP/BSI level of theory with ZPE and solvent corrections

included.

Table 6.1. Free energies of activation of hydrogen atom abstraction, rebound and

desaturation barriers.a

Substrate TSH b TSreb

c TSD

d

LS:

Ethylcarbamate 17.9 0.0 0.0

Valproic acid 14.4 0.0 0.0

Ethane 17.1 0.0 14.7

DHA 9.3 0.0 0.0

HS:

Ethylcarbamate 19.4 4.9 17.8

Valproic acid 15.0 1.3 23.0

Ethane 19.0 1.7 35.6

DHA 10.0 0.8 4.2

a G values at 298 K in kcal mol

–1.

b G relative to R.

c G relative to IH.

d G relative to

ID.

On the quartet spin state surface a small but significant radical rebound barrier exists (0.8 –

4.9 kcal mol–1

), whereas all low-spin rebound pathways are essentially barrier free. Details of

the rebound and desaturation barriers for all substrates are given in Table 6.1.

By contrast to the radical rebound process, the desaturation barriers (TSD) show

considerably more variation in relative energies. Similarly to the ethylcarbamate mechanism

from Figure 6.2, a substantial desaturation barrier of 4.2 – 35.6 kcal mol–1

is required for

valproic acid, ethane and dihydroanthracene desaturation on the quartet spin state. Note as

well that there is a significant desaturation barrier on the low-spin pathway for ethane, which

is further evidence that desaturation is an unlikely process for a substrate like ethane.

Page 179: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

179

Although, the barrier heights displayed in Table 6.1 implicate competitive hydroxylation

and desaturation on the doublet spin state surface, there appears to be a clear preference for

radical rebound on the high-spin surface. This is similar to previous studies on the

regioselectivity of substrate epoxidation versus suicidal complex formation in olefin

activation by P450 isozymes that was predicted to only lead to by-products on the high-spin

surface (De Visser, S. P.; Ogliaro, F.; et al 2001). It should be noted, however, that the

driving force for hydroxylation is larger for EC, VA and ethane as substrates, whereas it is

favorable for desaturation for DHA. Thermodynamically, therefore, DHA should

preferentially undergo the desaturation pathway, in accord with experimental observation that

anthracene is the major product (Yeong, Y. J.; Kang, Y.; et al 2008, Ji, L.; Franke, A.; et al

2014). By contrast, VA shows thermodynamically the dominant hydroxylase as the same as

EC. It seems the precursor leading to aromatic species on desaturation gives the highest

desaturase activity because of the increased resonance stabilization of the benzene product

(Kumar, D.; Tahsini, L.; et al 2009). Regardless of the external factors, we conclude that the

P450-catalyzed desaturation reaction can occur, but is competitive with hydroxylation

reactions on the low-spin state.

To further assess the calculated results, we estimated kinetic isotope effects for the

replacement of the transferring hydrogen atom by deuterium using the semi-classical Eyring

equation as well as the tunneling-corrected Wigner isotope effect (Heyes, D. J.; Sakuma, M.;

et al 2008).

Page 180: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

180

Figure 6.4: Potential energy profile of eicosanoic acid activation by 2Cpd I of P450 as

calculated with QM/MM. Energies are given in kcal mol–1

and are calculated at

UB3LYP/BSIV//UB3LYP/BSIII level of theory with ZPE corrections included. Optimized

geometries give bond lengths in angstroms and angles in degrees.

We reevaluated the free energy and imaginary frequency for the hydrogen atom abstraction

step for VA-4,4-d2 and based on a comparison with the fully hydrogenated system establish a

KIE = 6.4 (8.4) using the Eyring model and a KIE = 7.2 (12.1) at the doublet (quartet) spin

states with tunneling corrections included. These values compare well with the

experimentally determined KIE value of 5.58 for the same substrate (Rettie, A. E.; Boberg,

M.; et al 1988).

6.3.3. QM/MM studies of desaturation

As follows from the DFT model calculations in the previous section, both the radical

rebound and second hydrogen atom abstraction processes are barrierless on the doublet spin

2TSH,EA

rCH = 1.333rOH = 1.178

rFeO = 1.730rFeS = 2.420

aFeOC = 123.5

2TSD,EA

Page 181: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

181

state surface, whereas a substantial barrier on the quartet spin state surface is found. This

implies that the mechanistic switch from substrate hydroxylation to desaturation by P450-

catalyzed drug metabolism reactions is not kinetically controlled by the properties of the

oxidant and/or substrate but must be affected – at least in part - by external influences, such

as the substrate binding into the substrate binding pocket. Therefore, we followed our study

up by a set of QM/MM calculations on a system that is known to desaturate its natural

substrate. These calculations take the size and shape of the substrate binding pocket into

consideration and should give a realistic picture of the regioselectivity switch from

hydroxylation to desaturation.

The OleTJE P450 isozyme binds long chain fatty acids, such as eicosanoic acid (EA), and

converts them to olefins, CO2 and water using one molecule of oxygen. Figure 6.4 displays

the potential energy landscape as obtained by QM/MM for the hydrogen atom abstraction by

Cpd I from the C–H group and the pathway leading to desaturation. The hydrogen atom

abstraction barrier is low: E+ZPE = 16.4 kcal mol–1

for 2TSH,EA, and of the same order of

magnitude as those reported above for the DFT models as well as to previous hydrogen atom

abstraction barriers from the literature (Kamachi, T.; Yoshizawa, K. 2003, Shaik, S.; Kumar,

D.; et al 2005, Shaik, S.; Kumar, D.; et al 2008, Ogliaro, F.; Harris, N.; et al 2000, de Visser,

S. P.; Kumar, D.; et al 2004, Ji, L.; Schüürmann, G. 2013).

Page 182: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

182

Figure 6.5: Geometry scans for the rotation along the Fe–O bond from 2I as calculated

with QM/MM. Energies are given in kcal mol–1

and each data point represents a full

geometry optimization with fixed H–O–Fe–Nheme dihedral angle. Also shown are the maxima

of the scans with key hydrogen bonding interactions identified. The atom labelled with a

yellow star is C.

The hydrogen atom abstraction leads, as above, to an iron(IV)-hydroxo intermediate and

substrate radical. In the case of 2IH,EA the only mechanism we managed to locate is

desaturation to form CO2 and olefin products. Structures of the hydrogen atom abstraction

and desaturation transition states are given in Figure 6.4. The geometry of 2TSH,EA is

analogous to the hydrogen atom abstraction barriers of ethylcarbamate, valproic acid and

ethane, where long C–H and short H–O distances are found to be of 1.33 and 1.18 Å,

respectively. At the same time, the Fe–O bond elongates from about 1.602 Å in 2REA to 1.730

Å in 2TSH,EA, which is indicative of electron rearrangement within the Fe–O bond, see

below.

* = C

** W2 W1

W3

-60

-40

-20

0

20

0 50 100 150 200 250 300

2IH,EA

E

[kcal mol‒1]

dihedral angle H‒O‒Fe‒N [degrees]

11.5

3.8

W2

W1 W3

Page 183: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

183

In order to establish pathways leading to desaturation and hydroxylation, we did a series of

geometry scans starting from 2IH,EA, whereby the geometry was stepwise modified along one

degree of freedom, while the rest of the structure was energy minimized. In this particular

case, we searched for pathways from 2IH,EA to either

2IR,EA or

2ID,EA, and as such ran

geometry scans for the dihedral angle rotation (H–O–Fe–Nheme) from 172.6 in 2IH,EA in

gradual steps of 10 in forward and reverse directions for 180. Figure 6.5 gives the obtained

energies for the geometry scans as well as structures of the optimized geometries of the

maxima along the scan in each direction. The clockwise scan (pathway from 2IH,EA to the

right in Figure 6.5) has a small barrier of about 3.8 kcal mol–1

and collapses to alcohol

product complexes by rebound of the OH group to the substrate radical. The structure shows

that this pathway is assisted by hydrogen bonding interactions, whereby the FeOH group

hydrogen bonds to the carboxylic acid group of substrate but also receives a hydrogen bond

from a nearby water molecule in the active site. In this particular P450 isozyme we identified

a large water channel penetrating into the substrate binding pocket and three of these water

molecules are highlighted as W1, W2 and W3 in Figure 6.5. They form clusters of hydrogen

bonding interactions and form a tight network that includes the salt bridge between Arg245

with the carboxylate group of the substrate as well as the channel leading to the propionate

chains of the heme. By contrast, the substrate binding pocket of P450cam is very tight and has

few water molecules present (Poulos, T. L.; Finzel, B. C.; et al 1985, Schlichting, I.;

Berendzen, J.; et al 2000, Auclair, K.; Moënne-Loccoz, P.; et al 2001).

In the anticlockwise direction the scan encounters a much larger barrier of 11.5 kcal mol–1

.

The structure of the maximum of the scan (left-hand-side of Figure 6.5) identifies the key

reason for this high barrier. Thus, upon rotation the OH group of the iron(IV)-hydroxo moiety

clashes with protons of water molecule W1 and hence this pathway is disfavored over the

alternative rotation.

Page 184: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

184

Figure 6.6: Valence bond curve crossing diagrams for product formation from radical

intermediates. (a) Radical rebound leading to hydroxylation products. (b) Hydrogen atom

transfer to give desaturation products.

It appears that the subtle network of hydrogen bonding interactions affects the

regioselectivity of the reaction. Note that also this pathway leads to OH rebound and

formation of alcohols. A further geometry scan for the C–COO dissociation in the substrate

leads to release of CO2 and the formation of olefin as shown in Figure 6.4. This pathway has

a lower barrier than the alternative OH rotation of Figure 6.5 and therefore, the desaturation

degree of freedom will be C–C cleavage here rather than OH rotation. The regioselectivity of

hydroxylation versus desaturation, therefore, in OleTJE is guided by the hydrogen bonding

network in the substrate binding pocket that destabilizes the low-spin rebound pathways that

are barrierless in the gas-phase and in model complexes.

6.3.4. Valence Bond rationalization of rebound and desaturation mechanisms

To understand the origin of the regioselectivity of desaturation over hydroxylation we set

up valence bond (VB) curve crossing diagrams that rationalize the kinetics of the reactions.

Figure 6.6 show the VB curve crossing diagrams for the radical rebound and desaturation

Page 185: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

185

pathways leading to the competitive hydroxylation and desaturation products. These curve

crossing diagrams start on the left-hand-side with the radical intermediates, i.e. 4,2

IH, which

are described as an iron(IV)-hydroxo (porphyrin) complex with a nearby radical. The wave

function I represents structure 4,2

IH and corresponds to an electronic configuration x2–y22

*xz1 *yz

1 Rad

1 with two unpaired electrons along the Fe–O bond and a third on the substrate

in the Rad orbital. On the bottom right-hand-side of each Figure the product complexes are

drawn, which are either alcohol products or olefin products. The product complexes 4,2

PA and

4,2PD have wave function P and D, respectively, whereby the doublet spin states have

configuration x2–y22 *xz

2 *yz

1 and the quartet spin states x2–y2

2 *xz

1 *yz

1 *z2

1. In a

curve-crossing diagram, as displayed in Figure 6.6, the reactant wave function (I) connects

with an excited state wave function (P* in part a and D* in part b) in the product

geometry. At the same time, the ground state product wave function connects to an excited

state in the reactant geometry (I,reb* in part a and I,D* in part b). These two curves cross

and lead to an avoided crossing and hence a transition state for the reaction pathway from

intermediate to products.

Mathematically, the height of the barrier (E‡) can be determined from a VB curve

crossing diagram by estimating the excitation energy (or promotion gap G) in the reactant

geometry from I to I*. Thus, as has been shown previously (Shaik, S.; Hiberty, P. C.

2007), the barrier height is equal to the curve crossing energy (EX) minus the resonance

energy (B). Since, EX is proportional to a fraction (f) of the promotion gap, the barrier

height can be estimated from:

E‡ = EX – B = fG – B (1)

Page 186: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

186

A careful look at the VB structures in the reactant geometry for the hydroxylation pathway

shows that upon excitation the two electrons on the carbon of the substrate and the donor

oxygen atom in I,reb* have formed a bond and at the same time an electron has transferred

from the OH group to iron. The rebound excitation energy Greb, therefore, is dependent on the

strength of the CO bond formed, i.e. the bond dissociation energy of the CO bond (BDECO),

but also on the excitation energy (Eex,FeOH) within the iron-hydroxo complex from 2pz(OH)

into the iron 3d-block, Eq 2. This excitation energy will fill the *xz orbital with a second

electron in the doublet spin state and the *z2 orbital in the quartet spin state.

Greb = BDECO + Eex,FeOH (2)

The desaturation pathway, by contrast, shows somewhat different electron transfer

mechanisms as compared to the hydroxylation pathway, Figure 6.6b. Thus, in the excited

state I,D* the secondary hydrogen atom forms a chemical bond with the hydroxo moiety,

whereas this bond is missing but it forms a bond with the carbon atom of the substrate

instead. Therefore, the excitation energy is proportional to the difference in strength of the O–

H bond that is formed (BDEOH) and the C–H bond that is broken (BDECH). In addition, the

two radicals on the adjacent carbon atoms in I,D* have formed a -bond with energy E.

Obviously, -conjugation will have a major effect on the excitation energy and lower it

dramatically in value. Similarly to the radical rebound also in the desaturation pathway there

is an internal electron transfer from the hydroxo group to the metal (Eex,FeOH), Eq 3.

GD = BDEOH – BDECH + E + Eex,FeOH (3)

In order to stabilize a desaturation pathway over a hydroxylation pathway the barrier for

the latter should be higher in energy than that for the desaturation mechanism, i.e. Ereb‡ >

ED‡. Considering Eq 1 above, the difference in barrier height of radical rebound versus

Page 187: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

187

desaturation is then dependent on the promotion gap values G of the two processes and the

resonance energies B. Generally, f is found to have a constant value of 0.30 (Usharani, D.;

Lai, W.; et al 2014). Therefore, if both pathways incur similar resonance energy values B, the

change in product distributions will only be dependent on the relative values of Greb versus

GD, Eq 4.

Greb – GD = BDECO – BDEOH + BDECH – E (4)

The BDEOH value was calculated from the iron(III)-water complex and iron(IV)-hydroxo

intermediate to be 70.8 kcal mol–1

for the iron-porphyrin model used here (Kumar, D.;

Tahsini, L.; et al 2009). The corresponding C–H bond strength of the substrate that is broken

ranges from 36.7 kcal mol–1

for C2H5 to 21.8 kcal mol

–1 for cyclo-C6H7

. By contrast, the C–

O bond strength of the alcohol product that is formed ranges from BDECO = 85.8 kcal mol–1

for ethane to 59.7 kcal mol–1

for cyclohexadiene. Furthermore, DFT calculations predicted a

-conjugation energy of about 36 kcal mol–1

for a benzene molecule (Fernández, I.; Frenking,

G. 2006). Based on this, we calculate a difference Greb – GD = +16 kcal mol–1

for a molecule

like ethane, whereas a value of –25 kcal mol–1

is found for cyclohexadiene. This implies that

a regioselective hydroxylation of ethane may be expected, whereas the desaturation will be

favorable for cyclohexadiene. These values, therefore, implicate strong effects of the nature

of the substrate on the regioselectivity of hydroxylation over desaturation, which is

determined by a much stronger C–O bond formed in ethanol (by 26 kcal mol–1

) than the

corresponding alcohol of cyclohexadiene and in addition also has a stronger C–H bond for

the second hydrogen atom that is abstracted in the desaturation process (by another 15 kcal

mol–1

).

Page 188: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

188

As the difference Greb – GD for ethane is calculated to be +16 kcal mol–1

in favor of

substrate hydroxylation, this implies that perturbations through binding of the substrate in the

enzyme binding pocket will have to overcome 16 kcal mol–1

in order to change the

regioselectivity of hydroxylation over desaturation. As seen in Figures 6.4 and 6.5, an

aliphatic substrate like eicosanoic acid can be desaturated by P450 enzymes due to a tight

substrate binding pocket that through hydrogen bonding networks prevent the OH rotation

and, thereby, blocks the OH rebound process efficiently.

Subsequently, we calculated the difference Greb – GD for the drug molecules EC and VA as

well as for DHA and find values of 10, 13 and –4 kcal mol–1

, respectively. This implies that

the dihydroanthracene reaction will naturally be driven towards desaturation and the

formation of anthracene products. However, in the case of EC and VA the balance in the

isolated system is in favor of hydroxylation rather than desaturation. The P450 isozymes

involved in detoxification of EC and VA, therefore, have a specific substrate binding pocket

that will change the balance in favor of desaturation rather than hydroxylation due to

environmental perturbations affecting the rebound process. Consequently, subtle

perturbations in the substrate binding pocket can have a major effect on the regioselectivity of

substrate activation by P450 isozymes and are necessary to change the preference from

hydroxylation to desaturation. This may be an important point to consider into bioengineering

P450s for commercial purposes.

6.3.5. Molecular orbital rationalization of rebound and desaturation mechanisms

To further understand the intricate details of desaturation and hydroxylation reactions by

P450 enzymes, we did a molecular orbital analysis of the relevant orbitals in the reaction

process. The bifurcation originates in the radical intermediate IH, and, therefore, we looked

into the orbital levels of IH as compared to those in PA and PD. The change in orbitals should

reflect the height of the barrier as seen also above in the VB discussion.

Page 189: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

189

Thus, the radical intermediate has configuration x2–y22 *xz

1 *yz

1 Rad

1 and upon product

formation the xz2 *xz

1 pair of orbitals splits back into atomic orbitals to form 3dxz,Fe

2 2px,O

1.

Radical rebound via TSreb will lead to the formation of a O–C bond through the mixing of the

2px,O and Rad orbitals leading to the new pair O–C and *O–C, Scheme 6.4. On the other

hand, the desaturation reaction will lead to the mixing of 2px,O, Rad and C–H orbitals of the

radical intermediate into new O–H/*O–H and C=C orbitals. The possibilities and easiness of

this mixing will affect the kinetics of the rebound and desaturation barriers.

Scheme 6.4: Orbital mixing patterns for the pathways from radical intermediates to products.

Therefore, the relative stability of the desaturation versus hydroxylation product complex

will depend on the relative energies of the O–H2 C=C

2 orbitals with respect to the C–H

2 O–C

2

ones. Obviously if the energy of the O–H and C=C orbitals is lower than that of the sum of

the C–H and O–C pair then desaturation is more exothermic.

In a similar way, we could look at the orbitals of the radical intermediate IH. In both

processes the 2px,O and Rad orbitals are involved and are being rehybridized during the

Page 190: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

190

pathway leading to products. Along the desaturation, however, also the C–H orbital in IH

comes into play. We, therefore, analyzed the energy gap (E-) between Rad and C–H in the

radical intermediates for several substrates at the B3LYP/6-31+G** level of theory. We find

values of E– of 94.0, 99.7, 103.0 and 132.2 kcal mol–1

for ethylcarbamate, valproic acid,

dehydroanthracene and ethane, respectively. Clearly, the large energy gap between the Rad

and C–H orbital in a molecule like ethane will drive the reaction towards hydroxylation as

indeed also observed experimentally.

In addition to the above mentioned drug molecules, a range of substrates have been

identified as leading to desaturation products from P450 metabolized reactions. In particular,

further examples include the desaturation of testosterone to 17-hydroxy-4,6-androstadiene-

3-one (Nagata, K.; Liberato, D. J.; et al 1986), warfarin to dehydrowarfarin (Kaminsky, L. S.;

Fasco, M. J.; et al 1980), and lindane to 1,2,3,4,5,6-hexachlorocyclohexene (Chadwick, R.

W.; Chuang, L. T.; et al 1975). To test the hypothesis that E– can predict a desaturation

versus hydroxylation product ratio, we then calculated the radical intermediates for

testosterone, warfarin and lindane and find values of 99.7, 106.2 and 106.0 kcal mol–1

,

respectively. These three molecules also give low E– values and are expected to give

desaturation products at least in competition with hydroxylation.

Page 191: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

191

6.4 Conclusion

We present here a detailed set of computational results on the potential regioselectivity of

aliphatic hydroxylation versus desaturation. We identify key steps in the mechanisms that

both are initiated by a hydrogen atom abstraction step followed by OH isomerization to form

either the pro-hydroxylation radical intermediate or the pro-desaturation radical intermediate.

A VB and molecular orbital analysis identifies the electron transfer processes for both

reaction steps and highlights the substrate specific factors and orbitals that drive the

regioselectivity to hydroxylation or desaturation. Further evidence from QM/MM

calculations shows that the regioselectivity can be modified through external perturbations

that block one of the isomerization pathways.

Page 192: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

192

CHAPTER 7

PROJECT FIVE

Page 193: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

193

A trimetal carbene with reactivity Reminiscent of

Fischer-Tropsch Catalysis.5

ABSTRACT

Metal-carbenes are common reaction intermediates in chemical catalysis, however, very few

catalysts are known with a trimetal carbene. In this work we describe the chemical and

catalytic properties of a unique trimetal carbene containing an Ru2Pt-carbene core:

[(CpRu)2(2-CH2)(3-NCH3)Pt(P(CH3)3)2]. The chemical structure and individual orbital

interactions are identified from a natural bond orbital analysis and the reaction pathways for

consecutive intramolecular CH3 and H transfer (or vice versa) are investigated. It is shown

that this trimetal carbene can catalyze alkyl chain growth efficiently through an initial CH3

transfer followed by H migration. These reaction processes are rationalized by valence bond

models and thermochemical bond strength studies. The work leads to suggestions on how to

improve this chemical system for enhanced alkane synthesis.

5 Mala A. Sainna, Devendra Singh, Deve sh Kumar, Sam P. de Visser. “A trimetal carbene with reactivity Reminiscent of Fischer-Tropsch

Catalysis”Organomettalics, 2015, (34), 1651-1660

Page 194: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

194

7.1 Introduction

Ruthenium catalysts are very efficient and powerful and are being applied to a range of

chemical processes, including, for instance, olefin metathesis (Fürstner, A. 2000, Monsaert,

S.; Lozano Vila, A.; et al 2009, Vougioukalakis, G. C.; Grubbs, R. H. 2010, Lin, Y. A.;

Davis, B. G. 2010), water oxidation (Hurst, J. K. 2005, Sala, X.; Romero, I.; et al 2009,

Romain, S.; Vigara, L.; et al 2009, Concepcion, J. J.; Jurss, J. W.; et al 2009), the

enantioselective C–C bond formation and oxygen atom transfer reactions (Ikariya, T.;

Murata, K.; et al 2006, Fukuzumi, S.; Ohkubo, K. 2010, Man, W.-L.; Lam, W. W. Y.; et al

2014) . Of these processes, olefin metathesis is a carbon-carbon double bond rearrangement

reaction, where two olefins in a chemical reaction convert to two new olefins (Fürstner, A.

2000, Monsaert, S.; Lozano Vila, A.; et al 2009, Vougioukalakis, G. C.; Grubbs, R. H. 2010,

Lin, Y. A.; Davis, B. G. 2010). This reaction generally happens on an organometallic reaction

center often with Ru as the central catalyst, via the binding of a carbene group to the metal.

The catalytic mechanism of olefin metathesis involves a C–C bond breaking and the

formation of another C–C bond in a subsequent step. In most of these olefin metathesis

catalysts, the metal binds to a carbene either via a single or double bond, whereas carbon

centers ligated to multiple metal atoms are rarely found.

Over the past 30 years quite a number of ruthenium complexes have been identified that are

able to oxidize water to molecular oxygen (Hurst, J. K. 2005, Sala, X.; Romero, I.; et al 2009,

Romain, S.; Vigara, L.; et al 2009, Concepcion, J. J.; Jurss, J. W.; et al 2009, Huynh, M. H.

V.; White, P. S.; et al 2001, Meyer, T. J.; Huynh,M. H. V. 2003, Wang, L.-P.; Wu, Q.; et al

2010, Duan, L.; Bozoglian, F.; et al 2011, Barnett, S. M.; Goldberg, K. I.; et al 2012). These

complexes can either be mononuclear or binuclear ruthenium centers and reduce water via

proton-coupled electron transfer (PCET) to molecular oxygen, for instance, as a means to

generate electricity. In addition, ruthenium catalysts have been investigated extensively in

Page 195: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

195

porphyrin chemistry as a replacement of iron. These porphyrin complexes are based on

enzymatic systems including the cytochromes P450 that catalyze substrate hydroxylation and

epoxidation processes on an iron-heme center (Sono, M.; Roach, M. P.; et al 1996, Groves, J.

T. 2003, Ortiz de Montellano, P. R., 2004, Denisov, I. G.; Makris, T. M.; et al 2005, Kadish,

K. M.; Smith, K. M.; et al 2010, O'Reilly, E.; Koehler, V.; et al 2011). Biomimetic studies on

iron-porphyrin complexes investigated the reactivity and chemical properties of the active

site of P450 enzymes and gave insight into the fundamental properties of the catalyst that

drives the reaction. In early studies, Groves and co-workers(Groves, J. T.; Quinn, R. 1985,

Groves, J. T.; Ahn, K.-H. 1987, Groves, J. T.; Ahn, K.-H.; et al 1988, Groves, J. T.; Roman,

J. C. 1995, Groves, J. T.; Bonchio, M.; et al 1996) replaced iron by ruthenium in these

biomimetic porphyrin scaffolds and did a detailed comparison that focused on the reactivity

changes upon metal-replacement. They found ruthenium-porphyrin complexes to react more

efficiently with olefines than the corresponding iron analogues and, moreover, produced

stereospecific reaction products. Subsequent computational studies (Sharma, P. K.; de Visser,

S. P.; et al 2003) showed that this was due to the fact that ruthenium is stabilized in a higher

oxidation state than the corresponding iron porphyrin (Por) complex, i.e. RuV(O)(Por) versus

FeIV

(O)(Por+

), which resulted in differences in electron transfer processes and radical

rebound barriers. In these chemical systems, the active oxidant reacts with aliphatic groups

via proton-coupled electron transfer (PCET), whereby the proton is transferred to the oxygen

atom and the electron to the metal in a formal hydrogen atom abstraction step. PCET is a

common reaction type in transition metal catalysis and many examples have been identified

in the literature over the years (Crevier, T. J.; Mayer, J. M. 1998, Crevier, T. J.; Lovell, S.; et

al 1998, Bryant, J. R.; Mayer, J. M. 2003, Matsuo, T.; Mayer, J. M. 2005, Wu, A.; Mayer, J.

M. 2008, Maestri, A. G.; Cherry, K. S.; et al 2001).

Page 196: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

196

Recently, the Matsuzaka group (Takemoto, S.; Morita, H.; et al 2009) reported the

synthesis and spectroscopic characterization of [(Cp*Ru)2(2-CH2)(3-NPh)Pt(P(CH3)3)2],

which was treated with CH3OTf in diethylether and after heating in toluene produced a

unique structure (labelled as 3 in Ref 9) with a trigonal sp2 hybridized carbon atom linked to

three transition metal atoms. It was found that structure 3 reacted via alkyl chain growth via

H and CH3 migration from the RuH and PtCH3 ligands to form 2-CHCH3 products, Scheme

7.1. Therefore, structure 3 catalyzes the C–C and C–H bond formation reactions for the

synthesis of alkanes, which are thermodynamically difficult processes. It may very well be

that the reaction displayed in Scheme 7.1 represents a novel catalytic reaction for the

synthesis of alkanes on carbene centers, but in order to explore its potential and possibilities

further it is important to fully understand the origin of the catalysis and the chemical

properties that drive the reaction mechanisms. Currently, however, there is limited knowledge

on the chemical system displayed in Scheme 7.1 and experimental studies failed to identify

intermediates along the reaction mechanism. Hence it is unclear how and why this catalyst

operates and how it can be further improved. Moreover, there are very few examples in the

literature of analogous processes and this chemical system may pose a novel synthetic route

for the generation of alkanes on a carbene center.

Scheme 7.1: Catalyst investigated in this work for alkyl formation on a carbide center.

Page 197: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

197

The aim of the studies in this work, therefore, is focused on establishing the reaction

mechanism of alkyl chain growth on the carbene center and the chemical properties of the

catalyst shown in Scheme 7.1. Firstly, it is unclear whether there is an initial H-transfer

followed by CH3-transfer or the reverse. Moreover, it may very well be that both mechanisms

are kinetically and thermodynamically possible. Secondly, as shown in iron-porphyrin

chemistry for many examples the reaction can take place via proton transfer, hydrogen atom

transfer or hydride transfer (Mayer, J. M. 1998, Weinberg, D. R.; Gagliardi, C. J.; et al 2012).

Furthermore, hydrogen atom and hydride transfer processes may happen via PCET, whereby

the electrons move to a different donor site than the protons involved in the reaction. The

reaction described in Scheme 7.1 shows reminiscence with the ruthenium-porphyrin

reactivity described above, where either hydrogen atom transfer or PCET takes place from

substrate to metal-oxo or metal-nitrido oxidant. Consequently, there are many possibilities for

a chemical reaction here, which warrants a detailed computational investigation. Theory is

the ideal tool to distinguish between these mechanistic pathways and to establish the

electronic features that drive this important mechanism in chemical catalysis. In addition to

the reactivity questions and the synthesis of alkyl groups on this carbon center, there are also

questions related to the chemical structure of the central carbon atom in 3 itself. Thus, an sp2

hybridized carbon atom would make 3 a carbocation with planar conformation, which is

unlikely given the fact that the carbon atom is surrounded by positively charged metal ions.

On the other hand, if the carbon atom is negatively charged, i.e. a carbanion, it would have

pyramidal structure because of its sp3 hybridization, which is in disagreement with the

experimental crystal structure coordinates. Clearly, the structural and reactivity features of

this complex are poorly understood and theory may provide insight into the electronic and

structural features of the catalyst. Moreover, as it appears this is a novel catalytic system that

may be exploited further for the synthesis of linear alkanes from smaller components, i.e.

Page 198: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

198

methane and hydrogen, it may serve as a template for the design and development of future

catalysts.

The Ru2PtC core of structure 3 is an unusual structure in chemical catalysis but shows

some resemblance to Grubbs catalyst (Love, J. A.; Sanford, M. S.; et al 2003, Alcaide, B.;

Almendros, P.; et al 2009, Bernal, M. J.; Torres, O.; et al 2013), where the metal is covalently

linked to a carbon atom through either a single, a double or a triple bond. However, as far as

we know, there are no reported catalysts in the literature, where the carbon atom is ligated to

two Ru atoms and a Pt atom. There are several examples in the literature of carbido centered

structures, where the carbon atom is bound to three or more metal atoms (Miller, R. L.;

Wolczanski, P. T.; et al 1993, Su, C.-J.; Su, P.-C.; et al 1996, Peters, J. C.; Odom, A. L.;

1997, Caselli, A.; Solari, E.; et al 2000, Buchowicz, W.; Herbaczyńska, B.; et al 2012,

Harding, D. J.; Kerpal, C.; et al 2013, Borren, E. S.; Hill, A. F.; et al 2013), but very few of

these show catalytic activity. It, therefore, remains unclear why structure 3 is catalytically

active for the formation of alkyl groups from individual methyl, hydrogen and carbon groups.

In order to gain insight into the peculiar structure and catalytic reactivity of 3 we decided to

do a computational study and establish the electronic and thermochemical features. As we

will show here our studies give fundamental insight into the nature of the metal-carbon bond

and characterizes structure 3 as a trimetal carbene that can react via sequential methyl and

hydrogen atom transfer to form 2-CHCH3. These results may have direct relevance to

heterogeneous catalysis on metal surfaces.

7.2 Methods

The calculations reported in this work were performed using density functional theory (DFT)

methods as used previously on transition metal complexes by our groups (de Visser, S. P.

2010). We used DFT methods as implemented in the Jaguar 7.7 (Schrodinger, LLC, 2011)

Page 199: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

199

and Gaussian-09 (Frisch, M. J.; et al 2009) program packages. Initial geometries were

optimized (without constraints) in Jaguar using hybrid density functional theory method

(B3LYP) (Becke, A. D. 1993, Lee, C.; Yang, W.; et al 1988) coupled with an LACVP basis

set on ruthenium and platinum and 6-31G on the rest of the atoms: basis set BS1 (Hay, P. J.;

Wadt, W. R. 1985). A subsequent analytical frequency in Gaussian confirmed the structures

as local minima with real frequencies only or first order saddle points with one imaginary

frequency for the correct mode. Energies were then improved by running a single point

calculation in Jaguar using a triple- quality LACV3P+ basis set on iron and 6-311+G* on

the rest of the atoms: basis set BS2. We also tested the effect of dispersion on the reaction

energies by doing a UB3LYP-D3 single point calculation in Jaguar using the model of

Grimme and co-workers (Grimme, S.; Antony, J.; et al 2010). For a selection of chemical

systems, i.e. reactants and transition states, we did geometry optimizations at UB3LYP/BS2

and UB3LYP-D3/BS2, but found very little changes with respect to those obtained at

UB3LYP/BS1. Since, transition metal complexes often react via multi-state-reactivity (Shaik,

S.; de Visser, S. P.; et al 2002), we calculated the complete potential energy profile on the

lowest lying singlet, triplet and quintet spin states, and however, in all cases the singlet spin

state was the ground state and well separated from the other spin states.

As occasionally DFT calculations on spin state ordering and relative energies of transition

metal complexes shows fluctuations depending on the choice of the density functional

method (Quesne, M. G.; Latifi, R.; et al 2014), we decided to do a series of test calculations

with alternative DFT methods, namely B3LYP-D3 (Grimme, S.; Antony, J.; et al 2010),

BP86 (Becke, A. D. 1988, Perdew, J. P.; Burke, K.; et al 1996), MO6 (Zhao, Y.; Truhlar, D.

G. 2008), and PBE0 (Adamo, C.; Barone, V. 1999, Perdew, J. P.; Burke, K.; et al 1996).

These studies confirmed the trends and did not change the conclusions.

Page 200: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

200

The methods described here were used previously for analogous transition metal containing

systems in comparison with experimentally determined data and reproduced free energies of

activation within 4 kcal mol–1

(Kumar, D.; de Visser, S. P.; et al 2005, Vardhaman, A. K.;

Sastri, C. V.; et al 2011), but also gave reasonable agreement for spectroscopic data

(Karamzadeh, B.; Kumar, D.; et al 2010). To locate transition state structures we ran

extensive geometry scans between the various local minima, and used the maximum of those

scans as a starting point for the actual transition state searches. Moreover, these scans

confirmed that reactants and products indeed connect via the transition state.

Our chemical model is based on the crystal structure coordinates of [(Cp*Ru)2(2-H)(2-

NHPh)(3-C)PtCH3(P(CH3)3)2]+ (Takemoto, S.; Morita, H.; et al 2009) whereby we replaced

the Cp* groups with Cp and phenyl by methyl. The stereochemical effects of the methyl

groups of the P(CH3)3 groups on the reaction mechanism was investigated by replacement of

the P(CH3)3 groups with PH3. The latter structures are identified as 3’, IMe’ and IH’.

Charges reported in this work were obtained from a Natural Bond Occupation (NBO)

calculation as implemented in Gaussian program package at B3LYP/BS2 level of theory

(Weinhold, F.; Carpenter, J. E. 1988).

In order to fully understand the thermodynamics and kinetics of the mechanism of the

overall chemical reaction, we calculated the bond dissociation free energies (BDFE) of key

Ru/C–H and Pt/C–CH3 bonds in complexes 3, IMe, IH and PH. The release of either H•, H

+,

CH3• and CH3

+ from these four complexes as described by Eqs 1 – 4, whereby each

individual structure was calculated through a full geometry optimization to give the adiabatic

bond dissociation free energies. In addition, we calculated the methylenium transfer free

energy (MTFE) and the proton transfer free energy (PTFE). The latter was calculated through

the use of H2O/H3O+ couple to balance the equation. As there were large structural changes

Page 201: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

201

upon H/CH3 bond breaking on the various complexes we also calculated the diabatic bond

dissociation free energies, or vertical bond dissociation free energies, whereby we took the

optimized geometries of 3, IMe, IH and PH and ran single point calculations with either H•, H

+,

CH3• or CH3

+ removed.

3+ [3 – H]

+ + H

• + BDFEH (1)

3+ [3 – CH3]

+ + CH3

• + BDFEMe (2)

3+ + H2O [3 – H]

0 + H3O

+ + PTFE (3)

3+ [3 – CH3]

0 + CH3

+ + MTFE (4)

7.3 Results and Discussion

Before we discuss our calculated reaction mechanisms, let us briefly summarize the main

results on our reactant structure, 3. The optimized geometry of 3 is shown in Figure 6.1 and

gives bond distances to within 0.011 Å of the crystal structure coordinates reported by

Takemoto and co (Takemoto, S.; Morita, H.; et al 2009) and as such the DFT calculations

give reasonable chemical structures. The Pt–C bond of 1.960 Å represents a single bond and

is in good agreement with several reported Pt–C containing crystal structures from the

literature (de Quadras, L.; Bauer, E. B.; et al 2007, Che, C.-M.; Huang, J.-S. 2002, Das, R.

K.; Saha, B.; et al 2010, Algarra, A. G.; Grushin, V. V.; et al 2012). The two Ru–C bonds are

very close in bond length and the fact that one of the two atoms is protonated, therefore,

appears to have little effect on its interaction with carbon. Both Ru–C bonds have a typical

length corresponding to a single bond and compare to analogous structures reported before

Page 202: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

202

(Mutter, S. T.; Platts, J. A. 2011, Conner, D.; Jayaprakash, K. N.; et al 2003, Patra, S. K.;

Sadhukhan, N.; et al 2006, Patra, S. K.; Bera, J. K. 2006, MacGregor, S. A.; McKay, D.; et al

2013). As transition metal complexes containing iron or ruthenium generally contain close-

lying spin states and often react via multistate reactivity patterns (Shaik, S.; de Visser, S. P.;

et al 2002, Latifi, R.; Sainna, M. A.; et al 2013), structure 3 was optimized in the singlet,

triplet and quintet spin states. The closed-shell singlet spin state is the ground state and is

well separated from the triplet and quintet spin states by Ggas = 17.1 and 44.8 kcal mol–1

,

respectively. These values change by less than 1 kcal mol–1

when solvent corrections are

included. Clearly, catalyst 3 is a closed-shell species that will react via single-state reactivity

in contrast to metal(IV)-oxo and metal(V)-nitrido complexes from the literature (Vardhaman,

A. K.; Barman, P.; et al 2013).

Figure 7.1: Optimized geometry of 3 as calculated with B3LYP. Bond lengths are given in

angstroms and group NBO charges Q in atomic units. The right-hand-side displays the

natural bond orbitals and their ordering for those involving the central carbon atom with it

ligands.

1.9701.960

1.938

rRu-Ru = 2.531rRu-H = 1.610rPt-Me = 2.119

QC = 0.04QRu = –0.05QRuH = –0.19QPtMe = –0.32

Page 203: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

203

In order to shed light on the bonding pattern of the central carbon atom, we did a Natural

Bond Orbital (NBO) analysis of structure 3. Thus, the NBO analysis reveals three bonding

type orbitals for the central carbon atom, namely three single bonds for the interaction with

the two Ru atoms and the Pt atom. The PtC orbital represents the -bonding orbital between

Pt and C and is built up from the interaction of the 5dz2 orbital on Pt with the 2pz orbital on C.

In addition, there are two single bonds between the carbido atom and the two Ru atoms,

which result from the mixing of the 2s, 2px and 2pz atomic orbitals on C with the 4dz2 orbital

on Ru or the 4dxz on Ru(H). As such the central carbon atom is sp2 hybridized. However, the

fourth valence orbital on carbon is a virtual 2py atomic orbital. The NBO analysis, therefore,

assigns the central carbon atom as a carbene with three sp2 hybridized orbitals that are

occupied with two electrons in an orbital donating to Pt (PtC), and two bonding orbitals

where both C and Ru donate one electron each into the bond (RuC, RuC). This hybridization

scheme gives the central carbon atom its planar structure, Scheme 7.2. As such the chemical

structure of Ru2CPt represents a Fischer-carbene, where the lone pair of the carbon atom

interacts with an empty Pt orbital and the -type Pt orbitals give backbonding into the empty

2py orbital of the carbine (Vardhaman, A. K.; Barman, P.; et al 2013). The NBO charges

displayed in Figure 7.1 reveal a charge-neutral carbon atom (QC = 0.04) that is surrounded by

slightly negatively charged metal atoms. Consequently, there is very little charge built-up in

this chemical system.

Scheme 7.2: Hybridization scheme of 3.

Page 204: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

204

Subsequently, we calculated the potential energy profile for 2-CHCH3 formation starting

from 3. We tested two mechanisms as shown in Figure 7.2: (i) initial hydrogen transfer from

RuH to the carbene followed by methyl abstraction and (ii) initial methyl transfer from PtCH3

followed by hydrogen/proton transfer. Figure 7.2 starts from the center with 3 and follows the

two possible reaction mechanisms: The initial H-transfer (mechanism to the right) and the

initial CH3 transfer (mechanism to the left). The two mechanisms are described as follows: In

the first case, the metal-carbene group abstracts a hydrogen atom from the RuH group via a

transition state TSH to form the H-transfer intermediate IH. A subsequent CH3 transfer from

the PtCH3 group via a barrier TSreb,Me gives the 2-CHCH3 bound product PH. We also tested

the alternative pathway starting with an initial CH3 abstraction by the metal-carbene group

via transition state TSMe leading to the methyl-transfer intermediate IMe. This intermediate

then reacts further by H-abstraction from RuH via a transition state TSreb,H to form product

PMe.

Figure 7.2: Free energy profile of alkyl chain growth on 3 via either (i) CH3-transfer

followed by H-transfer (mechanism from the center to the left) or (ii) H-transfer followed by

CH3-transfer (mechanism from the center to the right). Free energies (in kcal mol–1

) are

TSHTSMe

rPtC = 1.987rRuH = 1.596

rHC = 1.446rRuH = 1.738

rRuC = 1.938rPtCH3 = 2.113i629.7 cm–1

rPtCH3 = 2.355rCC = 1.824i352.4 cm–1

IMe

rPtRu = 2.820rPtRu = 2.957rRuRu = 2.548dCRuRuPt = 50.4

IH

rPtRu = 3.086rPtRu = 3.370rRuRu = 2.620dCRuRuPt = 43.3

Page 205: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

205

obtained with B3LYP-D3/BS2 and contain zero-point, thermal and entropic corrections at

298K. Values in parenthesis include solvent corrections to the free energy. Optimized

geometries report bond lengths in angstroms, angles in degrees and the imaginary frequency

in the transition states in wave numbers.

The H-atom transfer reaction has the lowest free energy of activation of Ggas (Gsolv) =

17.3 (17.6) kcal mol–1

, however, its subsequent CH3-transfer to form products gives an

additional barrier with a maximum of 43.0 (42.4) kcal mol–1

. Therefore, the reaction that is

initiated with H-atom transfer has an unfeasible free energy of activation for TSreb,Me and will

not be able to proceed. This is in agreement with experimental observation that found

evidence of a structure with analogy to IH, but the methyl group bound to one of the Ru

atoms. Clearly, any IH formed during the reaction is a dead-end product and will have to react

back to reactants to be catalytically active.

The alternative pathway with an initial CH3-transfer, on the other hand, has a calculated

free energy of activation of Ggas (Gsolv) = 21.4 (20.9) kcal mol–1

, and, therefore, is

disfavored over the initial H-atom transfer process. However, this reaction leads to a very

stable intermediate 1IMe with an exergonicity of almost 10 kcal mol

–1. Its subsequent H-atom

transfer leading to products only has a free energy of activation of 19.1 (19.0) kcal mol–1

from intermediates and therefore product formation is a viable pathway through this channel

with a rate determining TSMe barrier. Consequently, DFT establishes a possible alkyl

synthesis mechanism for 3 as a stepwise mechanism with an initial CH3-transfer followed by

H-atom transfer. An analysis of the group spin densities and charges along the complete

reaction mechanism shows that all local minima and transition states are closed-shell singlet

species with no radical built-up during the reaction. Accordingly, the transfer of a H-atom or

Page 206: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

206

CH3-group does not leave a radical center behind but instead the charges are redistributed

over the chemical system to retain maximum electron pairing.

To test whether changing the density functional method would have an effect on the free

energy of activation in the chemical reactions, we decided to calculate the energies of 3, TSMe

and TSH using B3LYP-D3, B3LYP, BP86, PBE0 and MO6 on the optimized geometries of

Figures 7.1 and 7.2. Figure 7.3 displays the energy splitting of 3, TSMe and TSH as calculated

with a range of density functional theory methods. As follows from the results in Figure 7.3,

the ordering of the barrier heights stays the same between all DFT methods and the absolute

values of the barrier heights change by less than 3.5 kcal mol–1

maximally. Moreover, the

energy splitting between TSH and TSMe is calculated to be between 4.6 kcal mol–1

(for BP86)

and 7.1 kcal mol–1

(for B3LYP). Clearly, as the singlet spin ground state is well separated

from the higher spin states, the effect of amount of exchange in the DFT calculation has little

effect on the thermodynamics and kinetics of the reaction and all methods provide the same

qualitative result. An analysis of group spin densities and charges of these structures

calculated with alternative DFT methods gave no changes and confirmed the conclusions

drawn above.

The calculations presented in this work show that the reaction starts with an H/CH3 transfer

and no radical center is left behind. We made several attempts to swap molecular orbitals and

find alternative intermediates and transition states, corresponding to H+/CH3

+ transfer

intermediates or H–/CH3

– transfer intermediates but in all cases the structures converged back

to the lower-lying H-atom/CH3-transfer pathways or led to higher energy conformers.

Clearly, the cationic and anionic pathways are considerably higher in energy and will not be

relevant for the reaction mechanism, which we will rationalize with molecular orbital

analysis and thermochemical cycles below.

Page 207: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

207

Figure 7.3: Energies of initial H-atom or CH3-transfer reactions from 3 as calculated with

different DFT methods. All structures optimized at B3LYP/BS1 and single point calculations

with basis set BS2 applied. Relative energies are given in kcal mol–1

and include ZPE

corrections calculated at B3LYP/BS1.

So why is the reaction displayed in Figure 7.2 stepwise with consecutive CH3 and H-

transfer processes and what is the function of the Ru2Pt scaffold? These questions will be

answered in the following discussion. Structurally, there are major changes happening at the

Ru2Pt-carbene center during the reaction mechanism. Although 3 has an sp2 hybridized

carbon atom that lies in the plane of the Ru2Pt atoms, upon H/CH3-transfer the planarity is

distorted and the carbon atom relaxes to a pyramidal conformation due to sp3 hybridization.

The pyramidalization is considerably more pronounced for the CH3-transfer process than in

the H-atom transfer pathway as dihedral angles C–Ru–Ru–Pt of 50.4 for IMe and 43.3 for IH

are found. At the same time the Pt–Ru distances in IMe shorten by as much as 0.8Å with

Page 208: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

208

respect to 3 to well below 3Å, while in IH they stay considerably longer: rPtRu = 3.086 and

3.370Å. These large structural changes for the CH3-transfer pathway will have an effect on

the reaction energetics and kinetics as well as on the relative stability of IMe versus IH. Thus,

the Pt atom in IMe only has three ligands due to CH3-transfer to the carbon atom, which

means limited stereochemical interactions between the two P(CH3)3 groups and the rest of the

molecule so that the Pt atom can approach the two Ru atoms closely. In IH the Pt atom

remains tetracoordinated and the stereochemical interactions between all ligands and the rest

of the molecule prevent a close approach to the Ru2C center. As a consequence of the

stereochemical repulsions IH is considerably destabilized over IMe and the reaction pathway

with initial H-atom transfer is energetically disfavored.

To further establish these stereochemical interactions in IH versus IMe, we did a subsequent

set of calculations on these structures, but with the P(CH3)3 groups replaced by PH3:

structures IH’ and IMe’, Figure 7.4. The optimized geometry of 3’ (Figure 7.4) is very close to

that reported above for 3 in Figure 7.1 and only minor changes in bond lengths are obtained.

Clearly, the P(CH3)3 groups have little effect on the structure of the Ru2PtC core of our

chemical system. A similar situation occurs for IMe’, which has large structural similarities

with IMe reported in Figure 7.2. Major changes, however, are seen when IH is compared to

IH’. Thus, one of the Ru–Pt distances decreases from 3.370 to 3.116 Å and as a consequence

the pyramidality of the carbon center changes from dCRuRuPt = 43.3 in IH to a value of 51.9

in IH’. This has a major effect on the stability of IH’ and it is stabilized by 6.0 kcal mol–1

with

respect to 3’, however, the IMe’ remains the most stable conformation. This implies that the

P(CH3)3 groups incur a destabilizing effect on the reaction mechanism, but will not reverse

the preference of initial CH3-transfer over H-atom transfer.

Page 209: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

209

Figure 7.4: Optimized geometries of 3’, IMe’ and IH’ with bond lengths in angstroms.

To gain insight into the electronic changes during the reaction mechanism, a NBO analysis

was also performed on the intermediate and product complexes. The NBO calculation reveals

a configuration on the central carbon atom of 2s1.08

2p2.92

for IMe and 2s1.18

2p3.10

for IH. Both

intermediate complexes, therefore, have an electronic configuration close to sp3

hybridization, and as a consequence their geometries have changed from planar to pyramidal

in agreement with the optimized structures.

In order to fully understand the thermodynamics and kinetics of the mechanism displayed

in Figure 7.2, and, in particular, gain insight into the electron transfer mechanisms during the

reaction processes we set up Valence Bond (VB) curve crossing diagrams (Figure 7.5) for the

methyl transfer and hydrogen atom transfer steps. The VB curve crossing diagrams start at

the bottom-left with the reactant electronic configuration and wave function (3) of 3, which

is a carbene where two sp2 electrons form bonding orbitals with electrons on the two

ruthenium atoms and the other two electrons represent a lone pair orbital that donates a bond

IH’

1.9561.9651.500

1.935

rRu-Ru = 2.534rRu-H = 1.609rPt-C = 2.108

1.089

IMe’3’

rRu-Ru = 2.553rPt-Ru = 2.778rPt-Ru = 2.915rRu-H = 1.604dCRuRuPt = 48.5

rRu-Ru = 2.628rPt-Ru = 3.057rPt-Ru = 3.116rPt-C = 2.122dCRuRuPt = 51.9

Page 210: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

210

to Pt. Each of the dots in Figure 7.5 represents one of these valence electrons. In part (a) in

Figure 7.5 the reaction follows the methyl transfer from Pt to carbene and the structure on the

right-hand-side gives the electronic configuration of IMe with wave function Me. In IMe the

central carbon atom is sp3 hybridized and forms a single bond with Ru, Ru, Pt and CH3. In

VB theory the wave functions of the reactant and product complexes cross, i.e. 3 and Me in

Figure 7.5a, and lead to an avoided crossing and a transition state for the reaction from

reactants to intermediates (Shaik, S.; Hiberty, P. C. 2007). Thus, the curve crossing energy

(EX) is above the actual transition state (EMe‡) by a factor B, which is called the resonance

energy. It has been shown that the barrier height can be described by the difference in energy

by a fraction (f) of the promotion gap (G) and the resonance energy B: EMe‡ = fGMT – B. The

promotion gap signifies the excitation energy from the ground state wave function to the

product wave function in the geometry of the reactants, i.e. GMT is the energy difference

between 3 and 3* in the reactant geometry. The VB diagrams also give chemical

structures of the reactant and product wave functions of the ground and excited state species

with valence electrons identified with a dot.

Page 211: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

211

Figure 7.5: Valence bond curve crossing diagrams for methyl transfer (part a) and hydrogen

atom transfer (part b) from 3. Valence electrons are identified with a dot.

A close inspection of the VB diagrams in Figure 7.5 shows that the promotion gap for

methyl transfer (GMT) represents the breaking of the Pt–CH3 bond and the formation of a new

C–CH3 bond. In addition, the VB structures of 3 and 3* shows that the carbene carbon is

rehybridized from sp2 to sp

3 hybridization and one electron of the lone pair orbital pointing

toward Pt is promoted into the new sp3 orbital. At the same time the electron from the broken

Pt–CH3 bond moves into the carbene-Pt bond. Thus, the promotion gap also contains a

component for the excitation energy of the carbene from sp2 to sp

3 hybridization (Eex,C),

which essentially requires a promotion of an electron from the PtC orbital to the virtual 2py

molecular orbital on carbon. Energetically GMT can be described in terms of the bond

Page 212: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

212

dissociation energies (BDEs) of the respective bonds that are broken and formed in the

reaction process, Eq 5, whereby BDEPtCH3 represents the free energy to break the Pt–CH3

bond in 3 and BDECCH3 is the free energy to form the C–CH3 bond in IMe.

GMT BDEPtCH3 – BDECCH3 + Eex,C (5)

The situation is dramatically different for hydrogen atom transfer from 3 as shown in

Figure 7.5b. Thus, the barrier height for hydrogen atom transfer (EH‡) is proportional to the

promotion gap for hydrogen atom transfer (GHT), which correlates with the strength of the

Ru–H bond that is broken (BDERuH), the strength of the C–H bond that is formed (BDECH)

and again the excitation energy for the carbene to change from sp2 to sp

3 hybridization. In

addition to these three terms, the promotion gap GHT also contains a component for the

electron transfer energy from Ru to Pt (ETRuPt). The hydrogen atom transfer process,

therefore, incurs extra electron reorganization energy in the Ru2Pt scaffold, which is not

necessary for the methyl transfer step. The overall preference of hydrogen atom transfer

versus methyl transfer will depend on the relative strengths of the bonds that are broken and

formed in the process and the electron reorganization energy. To find out whether there are

major differences between the BDE values in Eq 5 and those in Eq 6, we decided to calculate

their values with DFT.

GHT BDERuH – BDECH + Eex,C + ETRuPt (6)

As the VB drawings in Figure 7.5 only give quantitative analysis of electron transfer

processes and no actual qualitative values, we decided to calculate BDEs and bond

Page 213: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

213

dissociation free energies (BDFEs) of key Ru/C–H and Pt/C–CH3 bonds in complexes 3, IMe,

IH and PH by calculating the release of a H•, H

+, CH3

• and CH3

+ from these four complexes

according to Eqs 1 – 4 above. Figure 7.6 displays the adiabatic as well as diabatic driving

forces for H/CH3 release from complexes 3, IH and IMe. Thus, in the adiabatic BDFE

calculations we did a full geometry optimization of all individual chemical structures in Eqs 1

– 4 and then took the difference in free energy for each of these reactions. However, due to

considerable rehybridization of the central carbon atom and as a consequence large

geometrical changes, the adiabatic BDFE values gave odd results. For instance, the adiabatic

bond dissociation energy for the breaking of the Ru–H bond in 3 appears negative, which

would imply spontaneous dissociation in disagreement with the experimental crystal

structure. Conversely, when we calculate the diabatic free energy change by taking the H and

[3 – H] products in the geometry of 3 then the BDFE is positive. The same is found for the

dissociation of a methyl group from 3. By contrast to structure 3, the adiabatic and diabatic

bond energies in structures IMe and IH are very similar and vary by less than 6 kcal mol–1

. In

the following, however, we will focus on the diabatic bond dissociation free energies.

Figure 7.6: Bond dissociation free energies (BDFEs in kcal mol–1

) of key bonds in structures

3, IMe and IH. Reactions calculated according to Eqs 1 – 4. Part (a) gives adiabatic BDFE

Page 214: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

214

values and part (b) diabatic BDFE values. Values in parenthesis are solvent corrected free

energies, whereas those out of parenthesis are gas-phase data.

Thermodynamically, the driving force to form IMe/IH from 3 represents the breaking of a

weak Pt–CH3 or Ru–H bond and the formation of a much stronger C–CH3/C–H bond. Hence

the reaction should be strongly exergonic, which is seen for the reaction via IMe in Figure 7.2,

but not for the one via IH. Thus, the hydrogen atom transfer from Ru in 3 to the carbene to

form IH results in the breaking of the Ru–H bond of BDFEH,diabatic = 65.7 kcal mol–1

in the

gas-phase and the formation of a C–H bond with a diabatic value of 108.6 kcal mol–1

and, if

the driving force was solely dependent on the change in bond strength the reaction would

have been exergonic by 42.9 kcal mol–1

. As shown in Figure 7.2, however, this reaction is

endergonic by 17.2 kcal mol–1

, which implies that the stereochemical and electronic effects,

i.e. the sp2 to sp

3 rehybridization (Eex,C) and the Ru PT electron transfer (ETRuPt)

accounts for a destabilization of IH by about 60 kcal mol–1

in free energy. This is explained

schematically in Figure 7.7, where we deconvolute the processes from 3 to IMe and from 3 to

IH into factors for bond breaking and formation (BDFE) and quantum mechanical effects

(EQM). Thus, the methyl transfer reaction from 3 results in the breaking of the Pt–CH3 bond,

which requires a BDFEMe,diabatic = 47.6 kcal mol–1

, and at the same time the formation of the

C–CH3 bond (BDFEMe,diabatic = 75.9 kcal mol–1

). Therefore, the change in bond dissociation

free energy for the methyl transfer is –28.3 kcal mol–1

. The reaction mechanism in Figure 7.2

found the methyl transfer process to be exergonic by 9.9 kcal mol–1

, which implies that the

quantum mechanical effect of the process accounts for 18.4 kcal mol–1

. The quantum

mechanical effect includes the rehybridization energy of the carbene group and the geometric

and stereochemical factors associated with the methyl transfer. By comparison, the hydrogen

Page 215: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

215

atom transfer shows a much larger change in bond dissociation energy due to the strong C–H

bond that is formed (BDFEH,diabatic = –42.9 kcal mol–1

). However, this reaction incurs a large

quantum mechanical effect of 60.1 kcal mol–1

and thereby makes the overall process highly

endergonic. The quantum mechanical effect includes the rehybridization of the carbon atom,

which will probably of the same order of magnitude as that found for the methyl transfer

process, i.e. about 18 kcal mol–1

. In addition, there are the above mentioned stereochemical

repulsions of the methyl groups of the P(CH3)3 moieties that raise IH by about 6 kcal mol–1

.

Finally, the electronic effects (EQM) for the hydrogen atom transfer reaction include the

electron transfer energy from Ru to Pt as described in the VB diagram in Figure 7.5.

Figure 7.7: Energy decomposition of the methyl and hydrogen atom transfer reaction from 3.

Free energies given are in kcal mol–1

.

In addition to removal of H/CH3 from our reactant complexes and intermediates, the

breaking of the bond into two ions was also investigated and is described as the proton

transfer free energy (PTFE) and methyl transfer free energy (MTFE). As follows from the

data in Figure 7.6, it requires considerably more energy to break the Ru–H and Pt–CH3 bonds

Page 216: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

216

than a homogeneous splitting into Ru+H

and Pt

+CH3

. The thermodynamic analysis

confirms the results described above in Figure 7.2 where either an H or CH3

is transferred

rather than H+ and CH3

+.

So what about the next step for the formation of 2-CHCH3 products? Scheme 7.3 displays

the electron transfer processes from IMe via TSreb,H to form products (part a) and those from

IH via TSreb,Me to form products (part b). The reaction from IMe to products results in the

breaking of the Ru–H bond and the formation of a new C–H bond. At the same time the C–Pt

bond breaks and Pt forms a new bond with Ru. By contrast, the pathway from IH leads to the

breaking of the Pt–CH3 bond and the transfer of the methyl group to carbon. However, this

process breaks the C–Pt bond and the Pt(P(CH3)3)2 group is left behind with two unpaired

electrons. Thus, an electron on Pt is promoted into a virtual orbital and the reaction proceeds

via a high energy pathway. Indeed, the DFT calculated potential energy profile in Figure 7.2

gives a high barrier for TSreb,Me in agreement with this. The VB mechanism of Scheme 7.3,

therefore, gives a rationale why the first group migration needs to come from the Pt ligand

and the electronic problems that are obtained when its ligand transfers last.

Scheme 7.3: VB description of second reaction steps. (a) hydrogen-atom abstraction. (b)

methyl-transfer.

Page 217: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

217

In summary, the DFT calculations presented here show a novel pathway for the synthesis

of linear alkanes on a trimetal-carbene center. The chemical system discussed here, however,

is inefficient as the lowest lying initial barrier is for hydrogen atom transfer, which leads to a

dead-end reaction intermediate. By contrast, the initial methyl transfer can be followed by

hydrogen atom transfer to lead to products. The question is how this catalyst may be

improved for alkane synthesis. The thermodynamic and VB Schemes above give some clear

indication to this. As shown in Eqs 5 and 6 above, the barrier heights TSH and TSMe are

dependent on the Ru–H and Pt–CH3 bonds that are broken and the C–H and C–CH3 bonds

being formed. One way to raise the hydrogen atom abstraction barrier is by replacing

hydrogen by deuterium. We tested this and find a Ggas = 17.7 kcal mol–1

for transfer of an

Ru–D group to form a C–D bond. However, this is only a minor increase with respect to the

Ggas = 17.3 kcal mol–1

found for hydrogen. An alternative solution to raise the initial

hydrogen atom abstraction barrier is by replacement of the Ru atoms by Os. Bond strengths

increase down the periodic table and it is expected that the BDFEOs–H is larger than the

BDFERu–H, which in itself is above the value of the BDFEFe–H. Indeed, the tabulated bond

strength of an Fe–H diatomic molecule is 43 kcal mol–1

, whereas a Ru–H diatomic molecule

has a reported bond strength of 56 kcal mol–1

(Lide, D. R. 1996) It may very well be that

structure 3 with both ruthenium atoms replaced by osmium is a powerful alkane synthesizing

catalyst that operates through an initial methyl transfer followed by hydrogen abstraction, but

further research will need to establish this.

The work described here may have relevance to biological and heterogeneous catalysis for

the synthesis of alkanes. In particular, in heterogeneous catalysis linear alkanes are generated

on a metal surface using often CO and H2 are reactants. It may very well be that the trimetal-

carbene structure discussed here is a common type intermediate in heterogeneous catalysis,

but further research will have to be performed to establish this.

Page 218: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

218

7.4 Conclusions

In summary, our calculations establish the key features of a synthetic homogeneous catalyst

for alkyl chain growth. A combination of DFT, NBO, VB and thermochemical studies has

been performed and established the intrinsic chemical properties of 3. Firstly, we characterize

3 as having a Ru2Pt-carbene core where the lone-pair is donated to Pt via

bonding/backbonding configuration. Secondly, we find a low energy mechanism of alkyl

formation through consecutive CH3 and H

transfer to the carbene with low free energy of

activation and exergonic driving force. We show that the reverse process is thermochemically

hampered and would lead to an excited triplet configuration.

Overall, our work identifies a novel catalyst for the synthesis of alkanes that starts from a

trimetal carbene. This unique structure is shown to be capable of intramolecular methyl and

hydrogen atom transfer to the carbene to form 2-CHCH3 products as precursor to alkanes.

Page 219: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

219

CHAPTER 8

PROJECT SIX

Page 220: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

220

Does hydrogen bonding-donation to

manganese(IV)-oxo and iron(IV)-oxo oxidants

affect the oxygen atom transfer ability? A

computational study.6

ABSTRACT

Iron(IV)-oxo intermediates are involved in oxidations catalyzed by heme and non-heme iron

enzymes, including the cytochromes P450. At the distal site of the heme in P450 Compound I

(FeIV

-oxo bound to porphyrin radical), the oxo group is involved in several hydrogen bonding

interactions with the protein, but their role in catalysis is currently unknown. In this work we

investigate the effects of hydrogen bonding on the reactivity of high-valent metal-oxo moiety

in a biomimetic model complex that has three hydrogen bond donors directed toward a

metal(IV)-oxo group. We show these interactions lower the oxidative power of the oxidant in

reactions with dehydroanthracene and cyclohexadiene dramatically as they decrease the

strength of the O–H bond (BDEOH) in the resulting metal(III)-hydroxo complex. Furthermore,

the distal hydrogen bonding effect cause stereochemical repulsions with the approaching

substrate and forces a sideways attack rather than a more favourable attack from the top. The

calculations, therefore, give important new insights into distal hydrogen bonding, and show

that in biomimetic, and, by extension, enzymatic systems, the hydrogen bond may be

important for proton relay mechanisms involved in the formation of the metal-oxo

6 Reza Latifi, Mala A. Sainna, Elena Rybak-Akimova,Sam P. de Visser. “Does hydrogen bonding-donation to manganese(IV)-oxo and

iron(IV)-oxo oxidants affect the oxygen atom transfer ability? A computational study” Chemistry A- European Journal. 2013, 19, 4058-4068.

Page 221: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

221

intermediates, but the enzyme pays the price for this by reduced hydrogen atom abstraction

ability of the intermediate. Indeed, in nonheme iron enzymes, where no proton relay takes

place, there generally is no donating hydrogen bond to the iron(IV)-oxo moiety.

8.1 Introduction

Hydrogen bonding is widespread around the active sites of oxidative metalloenzymes and has

functions ranging from protein/enzyme stability, substrate binding, protein folding and proton

relay mechanisms. In the catalytic cycles of metalloenzymes, often intramolecular hydrogen

bonding and protonation are important for the efficient generation of high-valent metal-oxo

intermediates ultimately responsible for substrate oxidation (Sono, M.; Roach, M. P. et al

1996, Groves, J. T. 2003, Ortiz de Montellano, P. R. 2004, Kadish, K. M.; Smith, K. M. et al

2010, de Visser, S. P.; Kumar, D. 2011, Dunford, H. B. 1999, Poulos, T. L. 2000, Veitch, N.

C. Smith, A. T. 2000, Kovaleva, E. G.; Neibergaul, M. B. et al 2007, Korendovych, I. V.;

Kryatov, S. V. et al 2007). For example, proton delivery to the coordinated peroxo moiety in

metal monooxygenase hydroxylase, a dinuclear non-heme iron enzyme, triggers O–O bond

cleavage in peroxo intermediate P and affords formation of a high-valent diiron-oxo

intermediate Q, which is capable of hydroxylating methane, the native substrate of MMO

(Murray, L. J.; Lippard, S. J. 2007, Siegbahn, P. E. M.; Blomberg, M. R. A. 2010). Similar

effects operate in heme systems, where the exact location of a proton-donating functionality

in myoglobin mutants can convert an oxygen carrier into a peroxidase (Raven, E. L. 2003,

Watanabe, Y.; Nakajima, H. et al 2007). Perhaps the best studied example of non-covalent

secondary sphere interactions and their role in redox chemistry of the enzyme is provided by

the cytochromes P450.

The cytochromes P450 (P450s) are vital enzymes for human health that are found in a vast

number of biosystems ranging from eukaryotes to prokaryotes (Sono, M.; Roach, M. P. et al

Page 222: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

222

1996, Groves, J. T. 2003, Ortiz de Montellano, P. R. 2004, Kadish, K. M.; Smith, K. M. et al

2010). They catalyze a large selection of oxygen atom transfer reactions including substrate

hydroxylation (aliphatic as well as aromatic), epoxidation, sulfoxidation and dehydrogenation

(Groves, J. T.; Shalyaev, K. et al 2000). As a consequence, these enzymes are dominant in

the liver, where they participate in the metabolism of drugs and xenobiotics as well as

detoxification processes (Guengerich, F. P. 2001, Munro, A. W.; Girvan, H. M. et al 2007).

In addition, they have functions in the area of biosynthesis of, e.g., hormones. The P450s

contain a heme-based active site where the metal (iron) is attached to the protein via a thiolate

linkage with a cysteinate residue. The sixth ligand site of the metal is the oxygen coordination

site and using a catalytic cycle that includes two reduction and two protonation steps and the

binding of molecular oxygen on a heme center an iron(IV)-oxo heme cation radical species is

created that is also called Compound I (CpdI) (Sono, M.; Roach, M. P. et al 1996, Groves, J.

T. 2003, Ortiz de Montellano, P. R. 2004, Kadish, K. M.; Smith, K. M. et al 2010, de Visser,

S. P.; Kumar, D. 2011, Denisov, I. G.; Makris, T. M. et al 2005, Nam, W. 2007). Due to its

high reactivity and consequently short lifetime, experimental characterization of CpdI was

hampered for a long time and only recently Rittle and Green found ways to spectroscopically

detect it and study its reactivity patterns (Rittle, J.; Green, M. T. 2010). Figure 8.1 shows the

substrate bound resting state structure of P450cam as taken from the 1DZ9 protein databank

(pdb) file (Schlichting, I.; Berendzen, J. et al 2000). The oxygen atom is in close proximity

to the substrate but also interacts with several nearby hydrogen bond donors, such as the

hydroxyl group of Thr252. Although it is believed that hydrogen bonding donors, such as the

alcohol group of Thr252 are vital for efficient proton transfer during the catalytic cycle, it is

not known if this hydrogen bond affects the subsequent substrate hydroxylation.

Page 223: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

223

Figure 8.1: Active site structure of P450 with key amino acids and substrate (camphor) and

solvent water (W) highlighted. Amino acids labelled as in the pdb file.

To gain further insight into the chemical and physical properties of iron(IV)-oxo species,

biomimetic model complexes have been designed that mimic the activity and structure of key

enzyme intermediates (Costas, M.; Mehn, M. P. et al 2004, Kryatov, S. V.; Rybak-Akimova,

E. V. et al 2005, Abu-Omar, M. M.; Loaiza, A. et al 2005, van Eldik, R. 2007, Bruijnincx, P.

C. A.; van Koten, G. et al 2008). These studies have given insight into the effect of the heme,

the nature of the axial ligand and the substrate range of CpdI and related high-valent iron-oxo

species. Very little is known, however, regarding the effects of hydrogen bonding interactions

on the distal site of CpdI and the role of donating hydrogen bonds on the reactivity of the

metal-oxo groups in general. It may be envisaged that these hydrogen bonding interactions

affect the chemical and electronic properties of the oxidant and its reactivity patterns. In line

with this, environmental effects were shown to be important in biomimetic metal-oxo species.

Cys357

Thr252

W902

camphor

Page 224: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

224

Recent studies showed dramatic effects on the electronic properties and catalytic activity of

metal-oxo oxidants upon addition of redox inactive cations, e.g. Sc3+

, Ca2+

or Zn2+

, to a

reaction mixture (Pfaff, F. F.; Kundu, S. et al 2011, Fukuzumi, S.; Morimoto, Y. et al 2010).

In particular, addition of Zn2+

to a manganese(V)-oxo corrolazine system changed the

electronic state to a manganese(IV)-oxo corrolazine cation radical and improved its catalytic

properties through valence tautomerism (Leeladee, P.; Baglia, R. A. et al 2012). Although,

biomimetic studies on distal hydrogen bonding centers are scarce there have been some

recent reports on metal-oxo models with hydrogen bond donating ligands (Tani, F.; Matsu-

ura, M. et al 2001, Bénisvy, L.; Halut, S. et al 2006, Brook, R. L.; Borovik, A. S. 2010). In

one of these studies, Borovik and co-workers focussed on metal-oxo complexes,

[MIV

(O)(H3buea)]– with M = Fe, Mn and H3buea = tris[N′-tert-butylureayl-N-ethylene]amine,

as displayed in Scheme 8.1 (MacBeth, C. E.; Gupta, R. et al 2004, Borovik, A. S. 2005,

Parsell, T. H.; Behan, R. K. et al 2006, Shook, R. L.; Borovik, A. S. 2010, Gupta, R.; Lacey,

D. C. et al 2012). Thus, the distal hydrogen bonding interactions in [MnIV

(O)(H3buea)]– were

shown to affect the basicity of the metal-oxo complex considerably and as a result the

manganese(IV)-oxo complex reacts with substrates with stronger C–H bond strengths via

hydrogen atom abstraction than the corresponding manganese(III)-oxo system (Parsell, T. H.;

Yang, M.-Y. et al 2009). Furthermore, intramolecular hydrogen bonding toward the oxo

group stabilized the iron(IV)-oxo species by about 6 kcal mol–1

but more importantly,

decreased the oxygen atom transfer ability dramatically, and made it a somewhat weaker

oxidant (Dey, A.; Hocking, R. K. et al 2006). Further studies showed these complexes to be

highly reactive and even catalytically converted O2 to water in a mechanism resembling that

of the enzyme cytochrome c oxidase (Shook, R. L.; Peterson, S. M. et al 2011). Recent

studies of Kass and co-workers (Shokri, A.; Abedin, A. et al 2012) showed that

intramolecular hydrogen bonding can dramatically influence the pKa values of molecules, and

Page 225: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

225

thereby play a key role in reactivities. Clearly, intramolecular hydrogen bonding interactions

to a metal-oxo group affect the catalytic properties of the oxidant in either positive or

negative way. However, the intricate details of this process are currently unknown, and,

therefore, warrant a computational study that gains insight into the effect of distal hydrogen

bonding processes in nonheme iron(IV)-oxo and manganese(IV)-oxo oxidants. As such, we

decided to do a density functional theory (DFT) study into the electronic properties of

[MIV

(O)(H3buea)]– with M = Fe, Mn and H3buea = tris[N′-tert-butylureayl-N-ethylene]amine,

Scheme 8.1, RFe/RMn. This is a tripodal bipyramidal compound with three donating hydrogen

bonds to the oxo group. To find the effect of the substituent on the kinetics of the H-atom

abstraction reaction we also did studies where the three tert-butyl groups were replaced by

either smaller alkyl chains or a hydrogen atom, which are identified as reactants RX with X =

H, Me, Et or i-Pr. The studies presented here highlight the effect of hydrogen bonding

interactions on the distal site of transition metal-oxo complexes and the influence this has on

the kinetics and thermodynamics of hydrogen atom abstraction reactions.

Scheme 8.1: Models studied in this work.

N

N MIV

N

NO

N

N

NX

X

H

H

HX

O

O

O

M = Mn, FeX = H, Me, Et, i-Pr, t-Bu RH: M = Mn, X = H

RMe: M = Mn, X = MeREt: M = Mn, X = EtRi-Pr: M = Mn, X = i-Pr

RMn: M = Mn, X = t-BuRFe: M = Fe, X = t-Bu

Page 226: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

226

8.2 Methods

The studies presented in this work use density functional theory (DFT) methods as

implemented in the Gaussian-03 and Gaussian-09 software packages (Frisch, M. J. et al

2003). We use the unrestricted B3LYP density functional method (Becke, A. D. 1993, Lee,

C.; Yang, W. et al 1988) combined to a basis set (B1) that is build up from LACVP on Mn

and Fe and 6–31G on the rest of the atoms (Karamzadeh, B.; Kumar, D. et al 2010). All

structures described in this work are the result of a full geometry minimization (without

constraints) using these methods and a subsequent frequency calculation characterized the

structures as local minima or first order saddle points (transition states) with either real

frequencies only or a single imaginary frequency for the correct mode. Frequencies reported

in this work were scaled with a value of 0.9257 (Hay, P. J.; Wadt, W. R. 1985). Transition

states were located by initially running a stepwise geometry scan between two local minima

by changing the reaction coordinate in a full geometry optimization with one degree of

freedom fixed. The maximum point of these geometry scans was used as a starting point for

the transition state optimizations and established that the transition state is indeed connected

to the minima on each side of the barrier. These methods have been shown to accurately

reproduce structures and free energies of activation of reaction processes of metal-oxo

reactivities (Kumar, D.; Thiel, W. et al 2011). Further improvement of the energetics was

obtained through single point calculations using an LACV3P+ basis set on Mn and Fe and 6–

311+G* on the rest of the atoms: basis st B2. All energies reported in this work were taken

from the UB3LYP/B2 calculations and were corrected for ZPE. Single point calculations

with dispersion corrected B3LYP (Schwabe, T.; Grimme, S. 2007) were performed in

Gaussian and confirmed the obtained trends. Free energies reported here use UB3LYP-D/B2

dispersion corrected energies and include entropic, thermal and solvent corrections to the

energy at 298 K. To test the effect of the environment on the barrier heights we did single

Page 227: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

227

point calculations on all structures using the polarized continuum model as implemented in

Gaussian mimicking an acetonitrile solution, however, these studies only gave minor changes

in relative energies as compared to the gas-phase results and no changes in the spin state

ordering. Free energies contain entropic and thermal corrections to the energy at 298 K and

also use energies calculated with basis set B2. The described methods were applied

previously to oxygen atom transfer processes of metal-oxo complexes and calculated free

energies of activation were found to be within 3 kcal mol–1

of experimental data (Kumar, D.;

de Visser, S. P. et al 2005, de Visser, S. P.; Oh, K. et al 2007, Vardhaman, A. K.; Sastri, C.

V. et al 2011).

8.3 Results

In order to understand the effect of distal hydrogen bonding to metal(IV)-oxo intermediates,

we investigated the intrinsic chemical properties as well as the reactivity patterns of

[MIV

(O)(H3buea)]– with M = Mn or Fe, i.e. RMn and RFe. Before we discuss the reactivity

patterns, we will start, however, with an in-depth study into the low-lying electronic states of

these complexes. Figure 8.2 shows the high-lying occupied and virtual orbitals of RMn as

taken from the optimized quartet spin geometry. The molecular orbitals shown in Figure 8.2

are dominated by the metal 3d atomic orbitals and their interactions with neighboring atomic

orbitals. The molecular valence orbitals include a set of * orbitals (*xy, *xz, *yz) and a

pair of * orbitals (*x2–y2, *z2), where the z-axis is aligned with the metal-oxo bond.

Page 228: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

228

Figure 8.2: High-lying occupied and low-lying virtual orbitals of 4RMn. Orbital energies are

reported in au.

The *xz and *yz orbitals represent antibonding interactions between the metal 3dxz/3dyz and

2px/2py orbitals on oxygen. The *xy orbital is built up from antibonding interactions of the

metal 3dxy orbital with 2p orbitals of ligands in the xy-plane of symmetry. In heme enzymes

this orbital is non-bonding, but in pentacoordinated complexes like structures R here, there

are interactions with nitrogen atoms in the xy-plane (Hirao, H.; Que Jr, L. et al 2008). As a

consequence, it is substantially higher in energy than the corresponding orbital in heme based

iron(IV)-oxo species. Indeed, in 4RMn it is even higher in energy than the *xz and *yz

orbitals by more than 10 kcal mol–1

, which is unusual as compared to other heme and

nonheme metal(IV)-oxo calculations, where the *xy orbital tends to be the lowest in energy

(Kumar, D.; Hirao, H. et al 2005, de Visser, S. P. 2006, Ye, S.; Neese, F. 2011, Shaik, S.;

Kumar, D. et al 2005). Thus, in octahedral symmetry the 3dxy atomic orbital cannot interact

with orbitals in the xy-plane and hence is non-bonding (xy). By contrast, with only three

ligands in the xy-plane in trigonal bipyramidal structures there are interactions between the

3dxy atomic orbital on the metal with 2p orbitals on the ligands and hence the molecular *xy

2*z

xz* yz*

xy*

–0.083

–0.108–0.101

0.011

0.069

xy*

xz*yz*

22*yx

22*yx

2*z

0.073

0.0570.062

0.080

0.127

xy*

xz*yz*

22*yx

2*z

-orbitals -orbitals

Page 229: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

229

orbital is raised in energy with respect to heme enzymes. The * orbitals of

[MIV

(O)(H3buea)]– are high in energy and virtual, and look similar in shape and appearance

as those calculated before for nonheme metal-oxo complexes (Kumar, D.; Hirao, H. et al

2005, de Visser, S. P. 2006, Ye, S.; Neese, F. 2011, Shaik, S.; Kumar, D. et al 2005).

Interestingly, the orbitals displayed in Figure 8.2 give very little electron density

contributions on the bridging N–H bonds, which seems to implicate that their electronic

effect on the molecular orbitals is small. The electronic ground state of 4RMn has orbital

occupation *xz1 *yz

1 *xy

1 and formally corresponds to a manganese(IV)-oxo complex.

This state is well separated from the lowest lying doublet and sextet spin states 2RMn and

6RMn by 17.6 and 14.2 kcal mol

–1, respectively (E+ZPE in solvent; basis set BS2). The spin

state ordering and relative energies change little when the effect of either solvent, counterions

or a change in DFT method is applied. These spin state orderings are in agreement with

experimental EPR studies that characterized [MnIV

(O)(H3buea)]– as a high spin ground state

(Shook, R. L.; Borovik, A. S. 2010).

Replacing the manganese with iron gives only minor changes to the molecular orbitals and

electronic configuration and retains a high-spin ground state with orbital occupation *xz1

*yz1 *xy

1 *x2–y2

1. Experimental electron paramagnetic resonance studies indeed identified

this complex as a high-spin species in agreement with what we find here (Shook, R. L.;

Borovik, A. S. 2010). This result contrasts biomimetic iron(IV)-oxo complexes that are

usually have a triplet spin ground state (Rohde, J.-U.; In, J.-H. et al 2003, Martinho, M.;

Banse, F. et al 2005, Sastri, C. V.; Seo, M. S. et al 2005, De Oliveira, F. T.; Chanda, A. et al

2007, Jackson, T. A.; Rohde, J.-U. et al 2008). On the other hand, most enzymatic nonheme

iron(IV)-oxo species have a high-spin (quintet) ground state instead (Price, J. C.; Barr, E. W.

et al 2003, Hoffart, L. M.; Barr, E. W. et al 2006, Galonić, D. P.; Barr, E. W. et al 2007). This

apparent disparity in triplet-quintet spin state ordering and relative energies of nonheme

Page 230: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

230

iron(IV)-oxo complexes was proposed to be correlated with the *xy/*x2–y2 orbital energy

difference and a stabilization of the *xy orbital led to a triplet spin ground state (Hirao, H.;

Que Jr, L. et al 2008). The data shown in Figure 8.2 show that the *xy orbital, in contrast to

other nonheme iron(IV)-oxo complexes in the literature, is higher in energy than the *xz and

*yz molecular orbitals. Since, the *xy and *x2–y2 orbitals are in the same plane of

symmetry, their energy separation affects the spin state ordering. In pentacoordinated metal

complexes both orbitals have similar amount of antibonding interactions and consequently

they are of similar energy, thereby stabilizing the quintet spin state dramatically. Thus, in

trigonal bipyramidal conformation the 3dxy atomic orbital on the metal in the metal(IV)-oxo

complex will be able to form bonding and antibonding interactions with atoms in the xy-

plane. These interactions are not dramatically different for the 3dx2–y2 orbital and hence the

*xy and *x2–y2 orbitals are close in energy for structures R. Metals with four ligands in the

xy-plane, by contrast, have 3dxy and 3dx2–y2 orbitals that are either aligned with the metal-

ligand bonds or are in between the metal-ligand bonds. Hence, the energy gap between *xy

and *x2–y2 is wider in hexacoordinated metal complexes as compared to pentacoordinated

complexes. As a result of that spectroscopic studies on a trigonal bipyramidal iron(IV)-oxo

species give these species as either intermediate (triplet) or high spin state (England, J.; Guo,

Y. et al 2011, Wong, S. D.; Bell III, C. B. et al 2011), dependent on the ligand system.

Optimized geometries of 4RMn and

5RFe are given in Figure 8.3 and are in good agreement

with previous calculations on related complexes (Gupta, R.; Lacey, D. C. et al 2012). The

metal-oxo bonds are long, i.e. 1.706 Å for the Mn–O bond in 4RMn and 1.683 Å for the Fe–O

bond in 5RFe. This is typical for high-spin metal-oxo complexes and, for instance, previous

studies on Mn(IV)-oxo porphyrins also found distances around 1.7 Å for quintet spin state

structures (Balcells, D.; Raynaud, C. et al 2008, Latifi, R.; Tahsini, L. et al 2011, de Visser,

S. P.; Ogliaro, F. et al 2001). The calculated Mn–O distance is dependent on the overall spin

Page 231: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

231

state and oxidation state of the metal, and for instance Mn(V)-oxo complexes give

considerably shorter Mn–O distances of around 1.54 Å (Prokop, K. A.; Neu, H. M. et al

2011). Our optimized geometry of 5RFe is similar to related models calculated before by other

groups and also matches typical high-spin iron(IV)-oxo distances that also found bond

lengths around 1.68 Å (Dey, A.; Hocking, R. K. et al 2006, Rohde, J.-U.; In, J.-H. et al 2003,

Martinho, M.; Banse, F. et al 2005, Sastri, C. V.; Seo, M. S. et al 2005, De Oliveira, F. T.;

Chanda, A. et al 2007, Jackson, T. A.; Rohde, J.-U. et al 2008, de Visser, S. P.; Shaik, S. et al

2003, Siegbahn, P. E. M.; Borowski, T. 2006). This implies that donating hydrogen bonding

interactions toward the metal(IV)-oxo group, have little effect on the metal(IV)-oxo bond

length and bond strength. Green showed that the metal-oxo distance correlates with the

metal-oxo vibration according to Badger’s rule (Green, M. T. 2006), hence little

spectroscopic differences are expected due to internal hydrogen bonding interactions. Our

calculations give a Mn–O stretch vibration of 714 cm–1

, which compares favourably with the

reported FT-IR value from Ref (Parsell, T. H.; Behan, R. K. et al 2006). Furthermore,

replacing MnIV

-16

O by MnIV

-18

O gives a decrease of the manganese-oxo stretch vibration by

17 cm–1

as compared to the experimentally reported value of –18 cm–1

. Therefore, the

calculations correctly reproduce experimental spectroscopic features.

Figure 8.3: Optimized geometries with bond lengths in angstroms of 4RMn and

5RFe; group

spin densities () taken from UB3LYP/BS2 calculations. Also given are relative energies

rM(IV)O = 1.706 / 1.683

rM(IV)Nax = 2.207 / 2.152

M = 2.86 / 3.25

O = 0.24 / 0.43

4RMn / 5RFe

4RMn (2RMn) [6RMn]: 0.0 (22.1) [14.2]5RFe (3RFe) [1RFe]: 0.0 (17.0) [26.6]

E + ZPE + Esolv

Page 232: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

232

(including ZPE and solvent corrections in kcal mol–1

) of all low lying spin states for RMn and

RFe.

Subsequently, we investigated the hydrogen atom abstraction mechanism of substrates by

2,4,6RMn and

1,3,5RFe using dehydroanthracene (DHA) and 1,4-cyclohexadiene (CHD) as

model substrates. The reactions are stepwise with two subsequent hydrogen atom abstraction

processes leading to dehydrogenated products, i.e. anthracene and benzene, respectively. The

initial hydrogen atom abstraction is rate determining and passes a transition state (TSH) to

form a metal(III)-hydroxo complex (I) and is followed by a dehydrogenation transition state

(TSdh) leading to water and dehydrogenated product (P). Figure 8.4 displays the mechanism

for the initial hydrogen atom abstraction step by RMn using DHA as a substrate (left-hand-

side), while the optimized geometries of the H-abstraction transition states for both DHA

(TSH,DHA) and CHD (TSH,CHD) are given on the right-hand-side of Figure 8.4. The rest of the

mechanisms found for CHD is similar and only the energies vary slightly. The reaction takes

place via single-state-reactivity on dominant quartet (Mn) and quintet (Fe) spin state surfaces.

During the complete mechanism these spin states remain the ground state and other spin

states are well higher in energy, therefore, we will focus on the lowest lying spin state results

only. We also calculated the reaction mechanisms starting from 2RMn,

6RMn,

1RFe and

3RFe.

However, the barrier heights and the local minima for these alternative spin state surfaces

were considerably higher in energy than those reported for the structures found in the quartet

spin state for the manganese complex and the quintet spin state for the iron complex. As a

consequence, only the quartet spin state of RMn and the quintet spin state of RFe are

catalytically relevant, i.e. the reaction proceeds via single-state reactivity. This is unusual for

metal-oxo reactivities since often multiple spin states are involved in reaction mechanisms

and the intricate surface crossings between these spin states determines the actual mechanism

(Shaik, S.; de Visser, S. P. et al 2002, Manner, V. W.; Lindsay, A. D. et al 2012). However,

Page 233: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

233

extensive previous studies on nonheme iron(IV)-oxo oxidants showed that generally the

barriers on the quintet spin state surface are the lowest in energy (Kumar, D.; Hirao, H. et al

2005, de Visser, S. P. 2006, Ye, S.; Neese, F. 2011, Shaik, S.; Kumar, D. et al 2005, Aluri,

S.; de Visser, S. P. 2007, de Visser, S. P.; Tahsini, L. et al 2009, de Visser, S. P.; Latifi, R. et

al 2011). Thus, on the quintet spin state an electron is transferred from the substrate into the

*z2 molecular orbital, which has been termed a 5 process. By contrast, the lowest lying

triplet spin mechanism contains an electron transfer into the *xz orbital, labelled as a 3

process (Hirao, H.; Kumar, D. et al 2006, de Visser, S. P. 2006). Generally, mechanisms via a

5 pathway are lower in energy than those passing a 3 state. Since the structures discussed in

this work already start off from a quintet spin state, no further spin state crossing is needed

and the reaction proceeds on the same spin state surface via the favourable 5 mechanism.

The first H-atom abstraction from substrate by the oxidant leads to either a Mn(III)-hydroxo

or Fe(III)-hydroxo complex, whereby the electron transfer from the substrate enters the metal

*z2 orbital to give an electronic state for 4IMn of *xz

↑ *yz

↑ *xy

↑ *z2

↑ Sub

↓ and for

5IFe of

*xz↑ *yz

↑ *xy

↑ *z2

↑ *x2–y2

↑ Sub

↓, whereby the Sub orbital represents the radical on the

substrate rest group. Thus, the metal 3d-block is largely exchange coupled with either four

(Mn) or five (Fe) unpaired electrons, whereas the radical on the substrate moiety (Sub)

contains a down-spin electron. We also found an alternative transition state (5TSH,Fe’) with

configuration *xy2 *xz

↑ *yz

↑ *z2

↑ Sub

↑, however, in the gas-phase this barrier was 5.2 kcal

mol–1

higher in energy than the one reported in Figure 8.4. Therefore, electron transfer from

the substrate into the virtual *x2–y2 orbital is energetically favourable over electron transfer

in the low-lying *xy molecular orbital as it gives a fully exchange coupled metal 3d-block.

As follows from Figure 8.4, the H-atom abstraction from DHA by 4RMn costs 14.8 kcal mol

–1

in the gas-phase, whereas for the mechanism starting from 5RFe it is 16.6 kcal mol

–1. The H-

abstraction from CHD has a much higher barrier for 5RFe than for

4RMn, but the effect of

Page 234: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

234

solvent reverses the trends and makes the iron complex a slightly better oxidant by 2.2 kcal

mol–1

. These relative barrier heights implicate that RFe is a slightly better oxidant than RMn.

The results contrast the reactivity patterns of manganese-porphyrin versus iron-porphyrin,

where generally more favourable reactions are observed with manganese based oxidants

(Groves, J. T.; Stern, M. K. 1988, Yoshizawa, K.; Shiota, Y. et al 1998, Kurahashi, T.;

Kikuchi, A. et al 2008). The reason for this difference is that manganese-porphyrins,

manganese-corroles and manganese-corrolazines tend to have a low-spin ground state and the

corresponding Mn(V)-oxo species are closed-shell singlets (Nehru, K.; Kim, S. J. et al 2007,

Song, W. J.; Seo, M. S. et al 2007, Prokop, K. A.; de Visser, S. P. et al 2010, Kumar, A.;

Goldberg, I. et al 2010). By contrast, iron(IV)-oxo porphyrin cation radical species have

degenerate doublet and quartet spin states. As has been shown in many computational

studies, high-spin states react with substrates with lower H-abstraction barriers than

intermediate or low-spin states (Shaik, S.; Chen, H. et al 2011).

Figure 8.4: (a) Potential energy profile of hydrogen atom abstraction from DHA by 4,6,2

RMn

as calculated using DFT methods with energies in kcal mol–1

relative to the quartet spin

reactant complex. Energies are taken from the UB3LYP/BS2 calculations in the gas-phase,

while solvent corrected values are in parenthesis. Free energies with solvent, entropic and

dispersion corrections are given in square brackets. (b) Optimized geometries of the transition

2IMn

4TSH,Mn (5TSH,Fe)

4RMn

4IMn

4TSH,Mn

0.0 (0.0) [0.0]

14.8 (15.2)

[3.2]

rMO = 1.806 (1.815)

rHO = 2.948

(3.385)

rOH = 1.249 (1.322)

rHC = 1.346 (1.270)

i1822.6 (i1595.7) cm–1

4TSH,CHD,Mn [5TSH,CHD,Fe]

rMO = 1.803 [1.803]

rOH = 1.316 [1.179]

rHC = 1.295 [1.425]

i1667.4 [i1224.3] cm–1

E+ZPE = 15.8 (14.9) [22.7 (12.7)]

24.7 (23.6)

34.6 (34.6) [21.8]

6RMn

2RMn

15.4 (15.1) [13.5]

16.5 (17.6) [19.0]

6TSH,Mn

2TSH,Mn

–10.1 (–6.4) [–13.9]

–10.2 (–6.4) [–14.3]

10.0 (12.4) [1.7]

6IMn

(a) (b)

Page 235: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

235

states for hydrogen atom abstraction from DHA and CHD with bond lengths in angstroms

and the imaginary frequency in the transition state in cm–1

.

Optimized geometries are typical for nonheme metal-oxo complexes and are similar to

previous studies in the field (Hirao, H.; Que Jr, L. et al 2008, Kumar, D.; Hirao, H. et al 2005,

de Visser, S. P. 2006, Ye, S.; Neese, F. 2011, Shaik, S.; Kumar, D. et al 2005, Hirao, H.;

Kumar, D. et al 2006, de Visser, S. P. 2006, Johansson, A. J.; Blomberg, M. R. A. et al 2007,

Bernasconi, L.; Baerends, E. J. 2008, Comba, P.; Maurer, M. et al 2009, Prat, I.; Mathieson,

J. S. et al 2011). With DHA as a substrate the transition states are product-like for

4TSH,Mn,DHA with longer C–H than O–H bonds, but early for

5TSH,Fe,DHA: rCH = 1.346 (1.270)

and rOH = 1.249 (1.322) Å for 4TSH,Mn,DHA, respectively. Much larger differences are found

between the two H-atom abstraction transition states with CHD as a substrate, whereby the

Mn-structure has a central H-atom (rCH = 1.295 and rOH = 1.316 Å), but a more product-like

geometry is found for 5TSH,Fe,CHD (rCH = 1.425 and rOH = 1.179 Å). All these transition states

are characterized with large imaginary frequencies from i1224.3 – i1822.6 cm–1

, which

indicate that the reactions will encounter a large kinetic isotope effect for the replacement of

hydrogen atoms by deuterium as found before on related metal-oxo reactivities (Kamachi, T.;

Yoshizawa, K. 2003, Kumar, D.; de Visser, S. P. et al 2003, Kumar, D.; de Visser, S. P. et al

2004, D. Kumar, de Visser, S. P. et al 2004).

To test the effect of the substituents on the donating distal hydrogen bond, we did a further

set of calculations on the [MnIV

(O)(H3buea)]– complex where the t-butyl groups of H3buea

were replaced by i-propyl (Ri-Pr), ethyl (REt), methyl (RMe) or hydrogen (RH). We initially did

a full geometry optimization of the reactant complexes of RX, X = i-Pr, Et, Me and H in the

quartet spin state and the optimized geometries are given in Figure 8.5 alongside those for

4RMn that contains X = t-Bu. Geometrically, no dramatic changes are observed when the

substituent group is gradually reduced from t-butyl to a hydrogen atom. The Mn–O distance

Page 236: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

236

varies by a maximum of 0.012 Å, whereas the bond length of the metal with the axial ligand

Mn–Nax varies by 0.026 Å. Similar variations are found for the donating hydrogen bond

distances. A look at the group spin densities of these complexes shows minor differences in

electronic configuration and unpaired spin density.

Figure 8.5 also gives the vibrational frequencies of the Mn–O distances in all complexes,

MnO. As can be seen there are strong variations in the Mn–O stretch vibration from a value of

701.7 cm–1

for 4Ri-Pr to 731.4 cm

–1 for

4RH.

Figure 8.5: Optimized geometries of 4RX structures with different substituents X with bond

lengths in angstroms and the vibrational frequency in wave numbers. Note that 4RMn has X =

t-Bu.

OH

H

H

N

N

N

rMnO = 1.706 / 1.718 / 1.717 / 1.717 / 1.709rMnNax = 2.207 / 2.205 / 2.201 / 2.202 / 2.181

rOH = 1.780 / 1.825 / 1.820 / 1.822 / 1.804

rOH = 1.718 / 1.731 / 1.745 / 1.739 / 1.738rOH = 1.769

/ 1.780/ 1.780/ 1.790/ 1.788

X

X

X

4Rt-Bu / 4Ri-Pr / 4REt / 4RMe / 4RH

MnO = 714.4 / 701.7 / 706.4 / 712.2 / 731.4 cm–1

Page 237: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

237

Figure 8.6: (a) Barrier heights (E‡+ZPE in solvent) for H-atom abstraction from DHA by

various complexes manganese based complexes RX (X = t-Bu, i-Pr, Et, Me or H). (b) Barrier

heights as a function of the Mn–O frequency (MnO) in the reactant complex.

Subsequently, we investigated the H-atom abstraction of DHA by each of the oxidants

displayed in Figure 8.5. Similar to what we reported above, the reactions are stepwise with an

initial hydrogen atom abstraction to form a radical intermediate. DFT calculated barrier

heights (with solvent corrections included) on the quartet spin state are given in Figure 8.6 for

comparison. The lowest H-abstraction barrier is found with i-propyl substituents (E‡+ZPE =

13.4 kcal mol–1

). This barrier gradually increases to 16.1 kcal mol–1

upon replacement of the

i-propyl groups with shorter alkyl chains or a hydrogen atom. Consequently, there is a small

but significant substituent effect on the H-atom abstraction processes by these metal(IV)-oxo

complexes. Enlarging the substituent group further to t-butyl reverses the trend and raises the

barrier heights again. This is probably due to stereochemical interactions of the atoms of the

substituents.

In order to understand the trend in barrier heights, we plot in the lower panel of Figure 8.6

the barrier height of the hydrogen atom abstraction reaction relative to the MnO frequency in

the reactants. An almost linear correlation is found, which implicates that the strength of the

Mn–O bond affects the hydrogen atom abstraction. Interestingly, the complex with the longer

y = 0.09x - 48.47

R² = 0.91

0

5

10

15

20

25

690 700 710 720 730 740 750

E‡+ZPE

[kcal mol–1]

MnO [cm–1]

0

5

10

15

20

E‡+ZPE

[kcal mol–1]15.2

13.414.1

15.016.1

t-Bu i-Pr Et Me H

Page 238: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

238

Mn–O bond length in the reactants, and consequently the system with the weakest Mn–O

bond strength, gives the highest reaction barrier for this series of hydrogen atom abstractions.

Thus, even though the Mn–O bond is broken in the substrate hydroxylation/dehydrogenation

reaction, it is not its binding energy that determines the rate constant of hydrogen atom

abstraction. Instead, as will be explained in the Discussion section, it is the pKa value of the

oxo group and the electron affinity of the oxidant.

8.4 Discussion

This work describes the electronic and catalytic properties of metal(IV)-oxo complexes

with distal hydrogen bonding interactions. In the following we will try to rationalize the

effect of distal hydrogen bonding interactions on the mechanism and kinetics of H-atom

abstraction reactions by metal(IV)-oxo oxidants. Experimental studies (Mayer, J. M. 1998,

Mayer, J. M. 2004, Mader, E. A.; Manner, V. W. et al 2009, Kaizer, J.; Klinker, E. J. et al

2004, Yoon, J.; Wilson, S. A. et al 2009, Lansky, D. E.; Goldberg, D. P. 2006, Bell, S. R.;

Groves, J. T. 2009) on many occasions have shown that the rate constant of a H-atom

abstraction reaction correlates linearly with the strength of the C–H bond of the substrate that

is broken. Generally, a plot of the natural logarithm of the rate constant of hydrogen atom

abstraction of a selection of substrates by a specific oxidant gives a linear correlation with the

strength of these C–H bonds, i.e. bond dissociation energy (BDECH). Density functional

theory studies in combination with valence bond descriptions showed this correlation to

originate from the relative electronic configurations of the reactants and products (Shaik, S.;

Kumar, D. et al 2008, Shaik, S.; Lai, W. et al 2010, Latifi, R.; Bagherzadeh, M. et al 2009,

Kumar, D.; Karamzadeh, B. et al 2010, Kumar, D.; Sastry, G. N. et al 2011, Kumar, D.;

Sastry, G. N. et al 2012, de Visser, S. P.; Kumar, D. et al 2004). Further studies highlighted

the importance of the strength of the O–H bond that is formed in the processes and essentially

Page 239: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

239

assigned this O–H bond dissociation energy (BDEOH) of the metal(III)-hydroxo product as a

mimic to assign the catalytic power of the oxidant (de Visser, S. P. 2010). Recent

electrochemical studies of Fujii and co-workers (Takahashi, A.; Kurahashi, T. et al 2011)

established redox potentials of iron(IV)-oxo porphyrin cation radical models with varying

axial ligands. They found that an anionic axial ligand gave different redox potentials than a

neutral axial ligand, but changing an anionic axial ligand by another anionic axial ligand gave

little changes in redox potential. Subsequently, they investigated reactivity patterns and did

not find correlations between the rate constants of olefin epoxidation with electron properties

of the oxidant including the redox potential, however, they did find correlations with the

redox potential of the iron(III)porphyrin product complex (Takahashi, A.; Yamaki, D. et al

2012). By contrast, Nam and co-worker (Sastri, C. V.; Lee, L. et al 2007) investigated the

axial ligand effect of nonheme iron(IV)-oxo complexes with a TMC or TMCS ligand system

(TMC = 1,4,8,11-tetramethyl-1,4,8,11-tetraazacyclotetradecane; TMCS = 1-mercaptoethyl-

4,8,11-trimethyl-1,4,8,11-tetraazacyclotetradecane) and found the reactivity to be dependent

on the blending of quintet and triplet spin state surfaces. Thus the systems of Nam are in a

triplet spin ground state and during the H-atom transfer a spin state crossing to a more

favourable quintet spin state surface occurs. It was shown that the amount of blending and

spin state transfer affects the rate constants of the reaction. To find out whether the bond

dissociation energy of the O–H bond (BDEOH) in the iron(III)-hydroxo complex is related to

the rate constants and hydrogen atom abstraction barriers, we calculated reaction 1 for the

complexes described in this work.

[MIII

(OH)(L)] → [MIV

(O)(L)] + H• + BDEOH (1)

The calculated BDEOH values for 4RMn (

5RFe) are 81.2 (84.1) kcal mol

–1 in the gas-phase

and 83.9 (83.6) kcal mol–1

in solvent, and if the barrier height of H-atom abstraction is

Page 240: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

240

proportional to these values, the rate determining barrier should be very similar for these

oxidants. Indeed, the barrier heights for H-atom abstraction from DHA and CHD in Figure

8.3 above show that they are very similar for the iron and manganese complexes in support of

the relative BDEOH values. Our calculated BDEOH values also compare favourably with

experiment as Borovik and co-workers isolated and kinetically investigated the

[MnIV

(O)(H3buea)]– and [Mn

III(OH)(H3buea)]

– complexes and determined a BDEOH value for

this complex of 89 kcal mol–1

.[42]

The calculated BDEOH values for 4RMn and

5RFe are

somewhat lower than those found for [FeIV

(O)(Por+•

)X] with Por = Porphyrin and X an axial

ligand, such as chloride or thiolate, for which values of around 88 kcal mol–1

were obtained

(de Visser, S. P. 2010). Furthermore, much higher BDEOH values were obtained for nonheme

iron(IV)-oxo complexes of enzymatic as well as biomimetic model complexes with values

well over BDEOH = 90 kcal mol–1

(de Visser, S. P. 2010). These systems react with aliphatic

substrates with substantially lower H-atom abstraction barriers than those reported here for

4RMn and

5RFe in agreement with the relatively low BDEOH values.

To understand the effect of hydrogen bonding on the height of the H-abstraction barriers

we investigated the corresponding BDEOH values of hydrogen bonding and non-hydrogen

bonding metal(IV)-oxo complexes. Thus, we took the optimized geometry of 5RFe and the

corresponding iron(III)-hydroxo complex and rotated the peptide fragments of the H3buea

ligand by about 90 in such a way that no hydrogen bonding interaction to the oxo/hydroxo

group remains and recalculated the BDEOH value. The geometries were then again optimized

but with these fixed dihedral angles and the new non-hydrogen bonding energies were

calculated. This way, without hydrogen bonding interactions the [FeIV

(O)(H3buea)] system

has a BDEOH of 94.4 kcal mol–1

. Consequently, hydrogen bonding toward the oxo group in

iron(IV)-oxo complexes lowers the BDEOH value of the oxidant with 10.1 kcal mol–1

. Thus, if

BDEOH correlates linearly with hydrogen atom abstraction barrier as reasoned before (de

Page 241: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

241

Visser, S. P. 2010) this would imply a raise in barrier upon addition of hydrogen bonding

interactions at the distal position of the metal(IV)-oxo group by several kcal mol–1

. Indeed, a

comparison of a hydrogen atom abstraction barrier of RFe with the iron(IV)-oxo species of

taurine/-ketoglutarate dioxygenase gives substantially lower hydrogen abstraction barriers

for the enzyme (Latifi, R.; Bagherzadeh, M. et al 2009) possibly due to the lesser number of

hydrogen bonding interactions. In enzymatic systems, such as the P450s, CpdI is generated

from binding of O2 on a heme followed by two protonation steps and a reduction. The distal

hydrogen bonding interactions, therefore, are probably essential for proton transfer

mechanisms in order to facilitate production of an iron(IV)-oxo heme cation radical active

species. However, as shown here, the hydrogen bonding interactions that facilitate these

proton relay processes in the enzyme, actually lower the BDEOH values of the oxidant and

consequently lower the oxidative power of the enzyme. Nonheme iron enzymes, by contrast,

utilize -ketoglutarate as a co-substrate to generate an iron(IV)-oxo active species rather than

protonation and electron transfer mechanisms. To find evidence for our hypothesis that distal

hydrogen bonding interactions will lower the catalytic activity of metal(IV)-oxo oxidants, we

investigated a selection of crystal structures of typical nonheme iron dioxygenases and

searched for distal hydrogen bonding interactions, see Figure 8.7. We selected taurine/-

ketoglutarate dioxygenase (TauD), AlkB repair enzymes and cysteine dioxygenase (CDO)

and took an available pdb file for each of these from the protein databank (O’Brien, J. R.;

Schuller, D. J. et al 2003, Ye, S.; Wu, X. et al 2007, Yu, B.; Hunt, J. F. 2009). Thus, TauD is

a nonheme iron dioxygenase with the metal bound to a 2-His/1-Asp facial triad of His99,

Asp101 and His255 and with the help of -ketoglutarate (KG) as a co-substrate and O2 it

generates an iron(IV)-oxo active species, which abstracts a hydrogen atom from substrate

(taurine) to give hydroxylated products. In TauD, KG binds as a bidentate ligand and the

last ligand position (vacant in Figure 8.7) is reserved for molecular oxygen. As can be seen

Page 242: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

242

from Figure 8.7 this last ligand position has no hydrogen bonding donors nearby. The only

polar residues in the active site are involved in hydrogen bonding interactions with the

substrate, such as Arg270 and Asn95. In the AlkB repair enzyme the situation is very similar

and the metal is coordinated via a 2-His/1-Asp facial motif with His131, Asp133 and His187.

Also in AlkB the KG group binds as a bidentate ligand and the sixth ligand site of the metal

is occupied by a water molecule. No hydrogen bonding donors are located nearby this sixth

ligand site, although there is an arginine residue in the vicinity that forms a salt bridge with

KG. The third structure displayed in Figure 8.7 represents cysteine dioxygenase (CDO),

which is a nonheme iron enzyme involved in the detoxification of cysteine in the body. It

catalyzes the dioxygenation of the thiolate group of cysteine to cysteine sulfinic acid

products. The pdb file has an iron active site with the metal linked through interactions with

three histidine residues (His86, His88, His140) to the protein and substrate cysteinate binds as a

bidentate ligand via the amine and thiolate groups to the metal. The carboxylate group of

cysteine is involved in hydrogen bonding interactions with several residues including His155

and Tyr157. These residues are out of range of a possible iron(IV)-oxo intermediate and,

therefore, also in CDO there are no visible hydrogen bonding donors to the iron(IV)-oxo

active species. Clearly, the absence of hydrogen bonding donors to the active species of

nonheme iron enzymes implicates that hydrogen bonding has a negative effect on catalysis

and will raise barrier heights and, therefore, in metalloenzymes hydrogen bonding donors to a

metal-oxo active species are lacking unless an essential proton transfer step is required in the

catalytic cycle.

Page 243: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

243

Figure 8.7: Extracts of the active site environments of nonheme iron dioxygenases

representing from left-to-right: TauD (1OS7 pdb), AlkB (3I2O pdb) and CDO (2IC1 pdb).

Amino acids are labelled as in the pdb file.

Figure 8.8: (a) Orientation of substrate attack on the metal(IV)-oxo group with angles in

degrees and group spin densities in au. (b) Electron transfer processes and LUMO orbital that

is filled with one electron in the H-abstraction process.

His99 His255

KG

taurine Arg270

Asn95

TauD (1OS7)

Asp133

KGHis131

His187

DNA chain Arg210

AlkB (3I2O) CDO (2IC1)

His86

His140

His88 Cys93

Tyr157His155Cys

Asp101

2*z

Fe-O-H: 120.0 [117.8] Fe-O-H: 130.4 [121.8]

4TSMn,DHA [4TSMn,CHD] 5TSFe,DHA [5TSFe,CHD]

*z2

FeIV

O

e

H

C

(a) (b)

Mn = 3.63 [3.72]

OH = –0.07 [–0.16]

L = –0.18 [–0.15]

Sub = –0.38 [–0.40]

Fe = 3.96 [3.74]

OH = 0.03 [0.15]

L = 0.39 [0.27]

Sub = –0.38 [–0.16]

Page 244: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

244

Table 8.1. Calculated barrier heights for hydrogen atom abstraction from

dehydroanthracene and 1,4-cyclohexadiene by metal-oxo complexes (Energies are in

kcal mol–1

).

Complex E‡ + ZPE E

‡ + ZPE + Esolv G

‡ + Esolv Ref

Dehydroanthracene Data: 4[Mn

IV(O)(H3buea)] or

4RMn 14.8 15.2 21.1 This work

4RMn,H 15.6 16.1 19.1 This work

4RMn,Me 14.5 15.0 18.5 This work

4RMn,Et 14.4 14.1 17.7 This work

4RMn,i-Pr 14.4 13.4 16.1 This work

5RFe 16.6 16.5 23.6 This work

1[Mn

V(O)(corrolazine)] 18.7 18.7 31.2 [32c]

1[Mn

V(O)(corrolazine)F]

– 11.3 13.7 24.4 [32c]

1[Mn

V(O)(corrolazine)CN]

– 9.0 13.0 21.8 [32c]

2[Mn

V(O)(corrolazine

+•)]

+ 0.7 8.4 19.3 [25]

2[Fe

IV(O)(corrole

+•)] 4.1 14.7 25.7 [45]

Cyclohexadiene Data: 4RMn 15.8 14.9 21.4 This work

5RFe 22.7 12.7 18.9 This work

2[Fe

IV(O)(corrole

+•)] 5.1 11.5 22.1 [45]

2[Fe

IV(O)(Por

+•)Cl] 8.3 11.0 9.7 [46]

4[Fe

IV(O)(Por

+•)Cl] 9.1 11.4 11.0 [46]

5[Fe

IV(O)(TMC)(CF3COO)]

+ 3.8 7.8 17.4 [44]

5[Fe

IV(O)(TMCS)]

+ 8.2 11.3 21.6 [44]

4[Ni

II(O2)(TMC)]

+ 18.4 14.2 22.4 [47]

Another factor that affects the barrier heights here is the orientation of the substrate. In

particular, the relatively high H-abstraction barriers of 4RMn and

5RFe complexes discussed

above is due to the non-ideal geometric conformation of the transition states. Figure 8.8

highlights key structural features and group spin densities of the transition states to explain

this. Thus, the substrate attacks the metal(IV)-oxo group sideways whether we choose DHA

or CHD as a substrate with Fe–O–H angles of around 120. These angles are very small for

reaction mechanisms that take place via a 5 pathway.

Page 245: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

245

In principle, the hydrogen atom abstraction is accomplished with a one electron transfer

from substrate to oxidant. Starting from 4RMn with *xz

1 *yz

1 *xy

1 configuration, there are

five possible pathways, namely whereby each of the metal 3d-type orbitals is filled with one

extra electron. In the so-called -pathways the H-atom abstraction leads to double occupation

of either the *xz, *yz or *xy orbital with a second electron, whereas in the alternative -

pathways either the *x2–y2 or *z2 orbital is filled with one electron. These electron transfer

processes have been shown to determine the optimized geometry of the transition states (de

Visser, S. P. 2006). Thus, the lowest lying hydrogen atom abstraction barriers, 4TSMn,DHA,

4TSMn,CHD,

5TSFe,DHA and

5TSFe,CHD, include an electron transfer from the substrate into the

*z2 orbital that becomes singly occupied in the radical intermediate, hence these reactions

follows a 4/

5 pathway. As the electron transfer leads to donation of an electron into *z2

orbital, this geometrically should lead to an upright transition state structure with an almost

linear Fe–O–H angle of close to 180 as obtained for analogous complexes without the

hydrogen bonding features (de Visser, S. P. 2006). On the other hand, the -pathway

mechanisms take place with an electron transfer into a * orbital and consequently the

substrate will attacks the oxidant under an angle of around 120 to get favourable orbital

overlap between the donor and acceptor of the electron. In nonheme iron(IV)-oxo oxidants

the triplet spin barriers are usually of 3-pathway type and give structures with Fe–O–H

angles of around 120. Our triplet spin transition states indeed show electron transfer into the

metal -system and a geometry with Fe–O–H angle around 120. Starting from 5RFe, we also

located the 5 transition state (

5TSH,Fe’) with configuration *xy

2 *xz

↑ *yz

↑ *z2

↑ Sub

↑, and

found it to be substantially higher in energy than the transition state for the 5 pathway. It is

clear, therefore, that the -pathways for electron transfer are less favourable in energy than

the -pathways and the occupation of *z2 with an electron. Even a change of transition state

Page 246: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

246

geometry from an “ideal” conformation with almost linear Fe–O–H–C dihedral angle to a

more bend structure does not destabilize the 5-pathway enough to bring it above the

3/5

pathways.

In Table 8.1 we compare the H-atom abstraction barriers of the hydrogen bonded iron and

manganese complexes discussed in this work with DHA and CHD with similar complexes

calculated before using density functional theory (DFT) (Prokop, K. A.; de Visser, S. P. et al

2010, Hirao, H.; Que Jr, L. et al 2008, Latifi, R.; Valentine, J. S. et al 2012, Kumar, D.;

Tahsini, L. et al 2009, Latifi, R.; Tahsini, L. et al 2011). We have grouped the calculated

reactivity trends of metal-oxo and metal-superoxo complexes with DHA and CHD substrates

in Table 8.1 as a comparison with the reactivity studies reported here. We give calculated

gas-phase barrier heights (E‡+ZPE) as well as solvent corrected barrier heights

(E‡+ZPE+Esolv) and free energies of activation in solvent (G

‡+Esolv). Due to the fact that

the overall charge is not equal in all these complexes some very large variations are observed

between the gas-phase and solvent corrected results. In the discussion we, therefore, will

focus on the free energies of activation in solvent.

The H-atom abstraction of DHA by 4RMn has a free energy of activation that is somewhat

lower in energy than that found for analogous manganese(V)-oxo corrolazine complexes

(Prokop, K. A.; de Visser, S. P. et al 2010), and considerably lower in energy than

[FeIV

(O)(corrole+•

)] and Ni(II)-superoxo complexes. The free energies of activation of CHD

H-abstraction by 4RMn and

5RFe show comparable reactivity with CHD with

[FeIV

(O)(TMC)(CF3COO)]. Of course, the systems decribed in Table 8.1 follow different

electron transfer processes, and, hence correlations are difficult to assign here. However, the

work described above clearly highlights that 4RMn and

5RFe are in the high-spin configuration

due to the fact that these systems are trigonal bipyramical. Because of that the relative

Page 247: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

247

energies of the *xy and *x2–y2 molecular orbitals are small and the systems are stabilized in

high-spin states.

High-spin states, such as 4RMn and

5RFe react with substrates via electron abstraction into a

vacant *z2 orbital that is aligned along the metal-oxo axis. Substrate approach to the oxidant,

therefore, will encounter lesser stereochemical interactions with the ligand along this 5

pathway than in alternative spin states, where the substrate approach is more sideways (de

Visser, S. P. 2006). A comparison of the barrier height of H-abstraction from CHD by 4RFe,

5[Fe

IV(O)(TMC)(CF3COO)]

+ and

5[Fe

IV(O)(TMCS)]

+ shows that the free energy of activation

of 4RFe lies exactly in between that found for the other two complexes. The calculations here

implicate that 4RFe is a comparable oxidant to nonheme iron(IV)-oxo complexes with a TMC

ligand system.

Page 248: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

248

8.5 Conclusion

In this work we present a series of DFT calculations on nonheme metal(IV)-oxo complexes,

[MIV

(O)(H3buea)]– with M = Fe/Mn. We show that hydrogen bonding interactions to the oxo

group effect the catalytic properties of the oxidant dramatically. Firstly, they restrict the

approach of the substrate and as a consequence the hydrogen atom abstraction on the low-

lying 5-pathway takes place at an almost perpendicular angle to the metal-oxo bond rather

than aligned with it as would be expected from electron transfer mechanisms. Secondly,

hydrogen bonding lowers the BDEOH value and consequently the oxidative power of the

oxidant. Indeed, relatively high hydrogen atom abstraction barriers from dehydroanthracene

and cyclohexadiene are found as compared to alternative iron and manganese based oxidants.

The work has relevance to P450 chemistry and shows that distal hydrogen bonding

interactions may restrict the catalytic properties of the active species, but on the other hand

may stabilize the proton transfer mechanisms that generate Compound I.

Page 249: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

249

9.1 Concluding remarks

The presented thesis is strickly focused on primarily: the electronic structure calculations

aimed at studying the reactivity and trends associated with Iron porphyrin complexes.

Although, two out of the five result chapters were not entirely porphyrin but the

characteristics studied from them is directly relatable to Metal Porphyrin complexes. A quick

recarpitulation of the results obtained from each chapter is professed as viz;

The first result chapter detailed an earlier experimental result presented by sankar and co-

workers on the effect of counter anions on spin state ordering on diiron porphyrin complexes;

although their spectroscopic studies shows a considerable distressing of the porphyrin rings

depending on the approaching counter anions, our computational studies shows that there are

variety of external perturbation that could affect the spin state ordering of the complex apart

from the approaching counter anions.

The second result chapter presented within this thesis shows that All olefins undergo

oxygen atom transfer, whereas compounds with low ionization energy also give a certain

degree of hydride transfer and charge transfer reactions; the computational finding was

strongly supported by the experimentally determined reaction rate which correlate linearly

with the ionization potential of the substrate and indicated that the electron transfer from

substrate to oxidant is rate determining.

A chapter was also presented on the effect of axial ligands, namely; chloride verses

acetonitrile. It was shown that neutral axial ligand leads to displacement of the metal from the

plane through the porphyrin ring and results in different orbital interactions between metal

and porphyrin ring as compared to systems with an anionic ligand. The effect of equatorial

ligand on Porpyrin versus TPFPP was also investigated; it is shown that the substituents on

Page 250: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

250

the porphyrin ring can guide substrate binding through electrostatic interactions with halide

atoms, which lowers the barrier heights.

A chapter indicating the regioselectivity of aliphatic hydroxylation against desaturation was

also presented. The key steps in the mechanisms were identified. The detailed computational

studies suggests that both regioselectivity is initiated by a hydrogen atom abstraction step

tailed by OH isomerisatin to form either the two pro-hydroxylation radical intermediate or the

pro-desaturation intermediate.

The last two result chapters differ from the earlier chapters discussed in the sense that the

compex studied in the earlier chapters are related as they are either reactivity or trends

associated to metal porphyrin complexes.

However in the next chapter our findings recognize the key features of a synthetic

homogeneous catalyst for alkyl chain growth on Ru2Pt-carbene. Our result shows that a low

energy mechanism of alkyl formation through consecutive CH3 and H

transfer to the

carbene with low free energy of activation and exergonic driving force. We show that the

reverse process is thermochemically hindered and would lead to an excited triplet

configuration. Whereas in the last chapter, a thorough density functional theory analysis was

done on a biomimetic nonheme metal(IV)-oxo complexes, and it was found that hydrogen

bonding interactions to the oxo group effect the catalytic properties of the oxidant intensely.

Even though the studied complex is a non heme, the work has significance to P450 chemistry

and demonstrated that distal hydrogen bonding interactions may restrict the catalytic

properties of the active species, but on the other hand may stabilize the proton transfer

mechanisms that generate Compound I.

Page 251: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

251

References

Abatjoglou, A. G., Billig, E., Bryant, D.R. Organometallics, 2010, 3, 923

Abdulsattar, M. A. 2012. "SiGe superlattice nanocrystal infrared and Raman spectra: A

density functional theory study". J. Appl. Phys. 111 (4): 044306.

Abu-Omar, M. M., Loaiza, A., Hontzeas, N. Chem. Rev. 2005, 105, 2227

Adamo, C., Barone, V. J. Chem. Phys. 1999, 110, 6158

Ahlrichs, R., Bär, M., Häser, M., Horn, H., Kölmel, C., Chem. Phys. Lett. 1989, 162, 165

Alcaide, B., Almendros, P., Luna, A. Chem. Rev. 2009, 109, 3817

Alcalde, M., Farinas, E. T., Arnold, F. H. J. Biomol. Screen. 2004, 9, 141

Alcock, N.W., Bergamini, P., Kemp, T. J., Pringle, P. G. J. Chem. Soc. Chem. Commun.

1987, 235

Algarra, A. G., Grushin, V. V., MacGregor, S. A. Organometallics 2012, 31, 1467

Aluri, S., de Visser, S. P. J. Am. Chem. Soc. 2007, 129, 14846

Amatore, C., Jutand, A. Acc. Chem. Res. 2000, 33, 314

Anderson, G. K., Cross, R. J. Acc. Chem. Res. 2010, 17, 67

Angelelli, F., Chiavarino, B., Crestoni, M. E., Fornarini, S. J. Am. Soc. Mass Spectrom.

2005, 16, 589

Angelici, R. J. Acc. Chem. Res. 1988, 21, 387

Atkins, W. M., Sligar, S. G. J. Am. Chem. Soc. 1987, 109, 3754

Page 252: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

252

Atkins, W., Friedman, R. S. Molecular Quantum Mechanics, 3rd

Ed., Oxford University

Press, 1997

Auclair, K., Hu, Z., Little D. M., Ortiz de Montellano, P. R., Groves J. T. J. Am. Chem.

Soc. 2002, 124, 6020

Auclair, K., Moënne-Loccoz, P., Ortiz de Montellano, P. R. J. Am. Chem. Soc. 2001, 123,

4877

Bäckvall, J. E., Åkermark B., Ljunggren, S. O. J. Am.Chem. Soc. 1979, 101, 2411

Bakowies, D., Thiel, W. J. Phys. Chem. 1996, 100, 10580

Balasubramanian, R., Rosenzweig, A. C. Acc. Chem. Res. 2007, 40, 573.

Balcells, D., Raynaud, C., Crabtree, R. H., Eisenstein, O. Chem. Comm. 2008, 744

Barluenga, J., Fernández-Rodríguez, M. A., Aguilar, E. J. Organometal. Chem. 2005,

690, 539

Barnett, S. M., Goldberg, K. I., Mayer, J. M. Nature Chem. 2012, 4, 498

Bartmess, J. E., Georgiadis, R. M. Vacuum 1983, 33, 149

Bartoli, J. F., Brigaud, O., Battioni, P., Mansuy, D. J. Chem. Soc., Chem. Commun. 1991,

440

Bassan, A., Blomberg, M. R. A., Siegbahn, P. E. M., Que Jr., L. Angew. Chem. Int. Ed.

2005, 44, 2939

Bathelt, C. M., Zurek, J., Mulholland, A. J., Harvey, J. N. J. Am. Chem. Soc. 2005, 127,

12900

Becke, A. D. J. Chem. Phys. 1993, 98, 5648

Page 253: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

253

Becke, A. D. Phys. Rev. A 1988, 38, 3098

Belcher, J., McLean, K. J., Matthews, S., Woodward, L. S., Fisher, K., Rigby, S. E. J.,

Nelson, D. R., Potts, D., Baynham, M. T., Parker, D. A., Leys, D., Munro, A. W. J. Biol.

Chem. 2014, 289, 6535

Belevich, I., Verkhovsky, M. I., Wikström, M. Nature 2006, 440, 829

Bell, S. R., Groves, J. T. J. Am. Chem. Soc. 2009, 131, 9640

Ben-Daniel, R., de Visser, S. P., Shaik, S., Neumann, R. J. Am. Chem. Soc. 2003, 125,

12116

Bénisvy, L., Halut, S., Donnadieu, B., Tuchagues, J.-P., Chottard, J.-C., Li, Y. Inorg.

Chem. 2006, 45, 2403

Bernal, M. J., Torres, O., Martín, M., Sola, E. J. Am. Chem. Soc. 2013, 135, 19008

Bernasconi, C. F. Chem. Soc. Rev. 1997, 26, 299

Bernasconi, E. J., Baerends, Eur. J. Inorg. Chem. 2008, 1672

Bhowmik, S., Dey, S., Sahoo, D., Rath, S. P. Chem. Eur. J. 2013, 19, 13732

Bhowmik, S., Ghosh, S. K., Layek, S., Verma, H. C., Rath, S. P. Chem. Eur. J. 2012, 18,

13025

Bhowmik, S., Ghosh, S. K., Rath, S. P. Chem. Commun. 2011, 47, 4790

Bhowmik, S., Sil, D., Patra, R., Rath, S. P. J. Chem. Sci. 2011, 123, 827

Bianchini, C., Meli, A. Acc. Chem. Res. 1983, 31, 109

Billig, E., Jamerson, J. D., Pruett, R. L. J. Organomet. Chem. 1980, 192, C49

Page 254: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

254

Borovik, S. Acc. Chem. Res. 2005, 38, 54

Borren, E. S., Hill, A. F., Shang, R., Sharma, M., Willis, A. C. J. Am. Chem. Soc. 2013,

135, 4942

Bouaoud, S-E., Braunstein, P., Grandjean, D., Matt, D. Inorg. Chem. 1986, 25, 3765.

Brook, R. L., Borovik, A. S. Inorg. Chem. 2010, 49, 3646

Brookhart, M., Green, M. L. H. J. Organomet. Chem. 1983, 250, 395

Brooks, B. R., Bruccoleri, R. E., Olafson, B. D., States, D. J., Swaminathan, S., Karplus,

M. J. Comput. Chem. 1983, 4, 187

Brown, J. M., Cooley, N. A. Chem. Rev. 1988, 88, 1031

Bruijnincx, P. C. A., Buurmans, I. L. C., Gosiewska, S., Moelands, M. A. H., Lutz, M.,

Spek, A. L., van Koten, G., Klein Gebbink, R. J. M. Chem. Eur. J. 2008, 14, 1228

Bruijnincx, P. C. A., Buurmans, I. L. C., Gosiewska, S., Moelands, M. A. H., Lutz, M.,

Spek, A. L., van Koten, G., Klein Gebbink, R. J. M. Chem. Eur. J. 2008, 14, 1228

Bruijnincx, P. C. A., van Koten, G., Klein Gebbink, R. J. M. Chem. Soc. Rev. 2008, 37,

2716

Bryant, J. R., Mayer, J. M. J. Am. Chem. Soc. 2003, 125, 10351

Buchowicz, W., Herbaczyńska, B., Jerzykiewicz, L. B., Lis, T., Pasynkiewicz, S.,

Pietrzykowski, A. Inorg. Chem. 2012, 51, 8292

Busca, G., Chapter 1 - Heterogeneous Catalysts, in Heterogeneous Catalytic Materials,

G. Busca, Editor 2014, Elsevier: Amsterdam. p. 1

Calhorda, M. J., Brown J. M., Cooley, N. A. Organometallics, 1991, 10, 1431

Page 255: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

255

Caselli, A., Solari, E., Scopelliti, R., Floriani, C. J. Am. Chem. Soc. 2000, 122, 538

Chabinyc, M. L., Craig, S. L., Regan, C. K., Brauman J. I. Science 1983, 279, 1882

Chadwick, R. W., Chuang, L. T., Williams, K. Pestic. Bio-chem. Phys. 1975, 5, 575

Chan, A. S. C., Caroll, W. E., Willis, D. E. J. Mol. Catal. 1983, 19, 377

Che, C.-M., Huang, J.-S. Coord. Chem. Rev. 2002, 231, 151

ChemShell, a Computational Chemistry Shell, see www.chemshell.org, and "QUASI: A

general purpose implementation of the QM/MM approach and its application to problems

in catalysis" Sherwood, P., de Vries, A. H., Guest, M. F., Schreckenbach, G., Catlow, C.

R. A., French, S. A., Sokol, A. A., Bromley, S. T., Thiel, W., Turner, A. J., Billeter, S.,

Terstegen, F., Thiel, S., Kendrick, J., Rogers, S. C., Casci, J., Watson, M., King, F.,

Karlsen, E., Sjøvoll, M., Fahmi, A., Schäfer, A., Ch. Lennartz, J. Mol. Struct.

(Theochem.) 2003, 632, 1

Chen, J. B., Angelici, R. J. Coord. Chem. Rev. 2000, 206, 63

Chen, J., Daniels, L. M., Angelici, R. J. J. Am. Chem. Soc. 1990, 112, 199

Chiavarino, B., Cipollini, R., Crestoni, M. E., Fornarini, S., Lanucara, F., Lapi A., J. Am.

Chem. Soc. 2008, 130, 3208

Chinn, M. S., Heinekey, D. M. J. Am. Chem. Soc. 1990, 112, 5166

Chufán, E. E., Puiu, S. C., Karlin, K. D. Acc. Chem. Res. 2007, 40, 563

Collman, J. P., Boulatov, R., Sunderland, C. J., Fu, L. Chem. Rev. 2004, 104, 561

Collman, J. P., Brauman, J. I., Hampton, P. D., Tanaka, H., Bohle, D. S., Hembre, R. T. J.

Am. Chem. Soc. 1990, 112, 7980

Page 256: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

256

Collman, J. P., Kodadek, T., Brauman, J. I. J. Am. Chem. Soc. 1986, 108, 2588

Collman, J. P., Zeng, L., Wang, H. J. H., Lei, A., Brauman, J. I. Eur. J. Org. Chem. 2006,

2707

Comba, P., Maurer, M., Vadivelu, P. Inorg. Chem. 2009, 48, 10389

Comba, P., Rajaraman, G. Inorg. Chem. 2008, 47, 78

Company, A., Feng, Y., Güell, M., Ribas, X., Luis, J. M., Que Jr, L., Costas, M. Chem.

Eur. J. 2009, 15, 3359

Concepcion, J. J., Jurss, J. W., Brennaman, M. K., Hoertz, P. G., Patrocinio, A. O. T., Iha,

N. Y. M., Templeton, J. L., Meyer, T. J. Acc. Chem. Res. 2009, 42, 1954

Cong, Z., Kurahashi, T., Fujii, H. Angew. Chem. Int. Ed. 2011, 50, 9935

Conner, D., Jayaprakash, K. N., Wells, M. B., Manzer, S., Gunnoe, T. B., Boyle, P. D.

Inorg. Chem. 2003, 42, 4759

Conroy-Lewis, F. M., Mole, L., Redhouse, A. D., Litster, S. A., Spencer, J. L. J. Chem.

Soc. Chem. Commun. 1991, 1601

Cooper, H. L. R., Groves, J. T., Arch. Biochem. Biophys. 2011, 507, 111

Corey R.B., Pauling, L. 1953: “Molecular Models of Amino Acids, Peptides, and

Proteins”. Review of Scientific Instruments, Volume 24, Issue 8, 621

Costas, M., Mehn, M. P., Jensen, M. P., Que Jr, L. Chem. Rev. 2004, 104, 939

Crestoni, M.E., Fornarini S., Lanucara F., Warren J.J., Mayer J.M., J. Am. Chem. Soc.

2010, 132, 4336

Crestoni, M.E., Fornarini, S. Inorg. Chem. 2005, 44, 5379

Page 257: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

257

Crestoni, M.E., Fornarini, S. Inorg. Chem. 2007, 46, 9018

Crestoni, M.E., Fornarini, S., Lanucara, F. Chem. Eur. J. 2009, 15, 7863

Crevier, T. J., Lovell, S., Mayer, J. M., Rheingold, A. L., Guzei, I. A. J. Am. Chem. Soc.

1983, 120, 6607

Crevier, T. J., Mayer, J. M. J. Am. Chem. Soc. 1983, 120, 5593

Czarnecki, K., Nimri, S., Gross, Z., Proniewicz, L. M., Kincaid, J. R. J. Am. Chem. Soc.

1996, 118, 2929

Dansette, P. M., Bertho, G., Mansuy, D. Biochem. Biophys. Res. Commun. 2005, 338, 450

Das, R. K., Saha, B., Rahaman, S. M. W., Bera, J. K. Chem. Eur. J. 2010, 16, 14459

Dassama, L. M. K., Silakov, A., Krest, C. M., Calixto, J. C., Krebs, C., Bollinger Jr., J.

M., Green, M. T. J. Am. Chem. Soc. 2013, 135, 16758

Davydov, D. R., Bon Hoa, G. H., Peterson, J. A. Biochemistry 1999, 38, 751

Dawson, J. H., Holm, R. H., Trudell, J. R., Barth, G., Linder, R. E., Bunnenberg, E.,

Djerassi, C., Tang, S. C. J. Am. Chem. Soc. 1976, 98, 3707

De Oliveira, T., Chanda, A., Banerjee, D., Shan, X., Mondal, S., Que Jr, L., Bominaar, E.

L., Münck, E., Collins, T. J. Science 2007, 315, 835

de Quadras, L., Bauer, E. B., Stahl, J., Zhuravlev, F., Hampel, F., Gladysz, J. A. New. J.

Chem. 2007, 31, 1594

de Visser, S. P. Angew. Chem. Int. Ed. 2006, 45, 1790

de Visser, S. P. Chem. Eur. J. 2006, 12, 8168

de Visser, S. P. Inorg. Chem. 2006, 45, 9551.

Page 258: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

258

de Visser, S. P. J. Am. Chem. Soc. 2006, 128, 15809

de Visser, S. P. J. Am. Chem. Soc. 2006, 128, 9813

de Visser, S. P. J. Am. Chem. Soc. 2010, 132, 1087

de Visser, S. P. J. Biol. Inorg. Chem. 2006, 11, 168

de Visser, S. P. J. Phys. Chem. A 2005, 109, 11050

de Visser, S. P., Kaneti, J., Neumann, R., Shaik, S. J. Org. Chem. 2003, 68, 2903

de Visser, S. P., Kumar, D. (Eds.) Iron-containing enzymes: Versatile catalysts of

hydroxylation reactions in nature. RSC Publishing, Cambridge (UK), 2011

de Visser, S. P., Kumar, D., Cohen, S., Shacham, R., Shaik, S. J. Am. Chem. Soc. 2004,

126, 8362

de Visser, S. P., Latifi, R., Tahsini, L., Nam, W. Chem. Asian J. 2011, 6, 493

de Visser, S. P., Ogliaro F., Harris, N., Shaik S. J. Am. Chem. Soc. 2001, 123, 3037

de Visser, S. P., Ogliaro F., Sharma, P. K., Shaik S. Angew. Chem. Int. Ed. 2002, 41,

1947

de Visser, S. P., Ogliaro, F., Gross, Z., Shaik, S. Chem. Eur. J. 2001, 7, 4954

de Visser, S. P., Ogliaro, F., Harris, N., Shaik, S. J. Am. Chem. Soc. 2001, 123, 3037

de Visser, S. P., Ogliaro, F., Shaik, S. Angew. Chem. Int. Ed. 2002, 41, 1947

de Visser, S. P., Ogliaro, F., Shaik, S. Angew. Chem. Int. Ed. 2001, 40, 2871

Page 259: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

259

de Visser, S. P., Ogliaro, F., Sharma, P. K., Shaik, S. Angew. Chem. Int. Ed. 2002, 41,

194

de Visser, S. P., Ogliaro, F., Sharma, P. K., Shaik, S. J. Am. Chem. Soc. 2002, 124, 11809

de Visser, S. P., Oh, K., Han, A.-R., Nam, W. Inorg. Chem. 2007, 46, 4632

de Visser, S. P., Quesne, M. G., Martin. B., Comba, P., Ryde, U. Chem. Commun. 2014,

50, 262

de Visser, S. P., Shaik, S., Sharma, P. K., Kumar, D., Thiel, W. J. Am. Chem. Soc. 2003,

125, 15779

de Visser, S. P., Tahsini, L., Nam, W. Chem. Eur. J. 2009, 15, 5577

de Visser, S. P., Tan, L.-S. J. Am. Chem. Soc. 2008, 130, 12961

de Visser, S.P. Editor 2012, Academic Press. p. 1

de Visser, S.P. Introduction to quantum behaviour – a primer. In Quantum Tunnelling in

Enzyme-Catalysed Reactions, R. K. Alleman and N. S. Scrutton, Eds., RSC Publishing,

Cambridge, 2009, Chapter 2, 18

de Visser, S.P., Coord. Chem. Rev. 2009. 253, 5–6, 754

de Visser, S.P., et al. J. Am. Chem. Soc. 2001. 123, 13, 3037

Dekker, G. P. C. M., Buijs, A., Elsevier, C. J., Vrieze, K., van Leeuwen, P. W. N. M.,

Smeets,W. J. J., Spek, A. L., Wang, Y. F., Stam, C. H. Organometallics, 1992, 11, 1937

Denisov, I. G., Makris, T. M., Sligar, S. G., Schlichting, I. Chem. Rev. 2005, 105, 2253

Denmark, S. E., Fu, J. Chem. Rev. 2003, 103, 2763.

Dey, R. K. Hocking, P. L. Larsen, A. S. Borovik, B. Hedman, K. O. Hodgson, E. I.

Solomon, J. Am. Chem. Soc. 2006, 128, 9825

Page 260: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

260

Diver, S. T. Coord. Chem. Rev. 2007, 251, 671

Dolinsky, T. J., Czodrowski, P., Li, H., Nielsen, J. E., Jensen, J. H., Klebe, G., Baker, N.

A. Nucl. Acids Res. 2007, 35, 522

Dong, L., Duckett, S. B., Ohman K. F., Jones, W. D. J. Am. Chem. Soc. 1992, 114, 151

Dowers, T. S., Rock, D. A., Jones, J. P. J. Am. Chem. Soc. 2004, 126, 8868

Duan, L., Bozoglian, F., Mandal, S., Stewart, B., Privalov, T., Llobet, A., Sun, L. Nature

Chem. 2011, 4, 418

Dunford, H. B., Heme Peroxidases, Wiley-VCH, New York, 1999

E. MacBeth, R. Gupta, K. R. Mitchell-Koch, V. G. Young Jr., G. H. Lushington, W. H.

Thompson, M. P. Hendrich, A. S. Borovik, J. Am. Chem. Soc. 2004, 126, 2556

E. O'Reilly, V. Koehler, S. Flitsch and N. Turner, Chem. Commun. 2011, 47, 2490,

England, J., Guo, Y., Van Keuvelen, K.M., Cranswick, M. A., Rohde, G. T., Bominaar,

E. L., Münck, E., Que Jr, L. J. Am. Chem. Soc. 2011, 133, 11880

Enthaler, S. ACS Catalysis, 2013. 3, 2, 150

Evans, D. R., Mathur, R. S., Heerwegh, K., Reed, C. A., Xie, Z. Angew. Chem., Int. Ed.

Engl. 1997, 36, 1335

Ferguson-Miller, S., Babcock, G. T. 1996, 96, 2889

Fernández, I., Frenking, G. Chem. Commun. 2006, 5030.

Fioroni, M., T. Dworeck, and F. Rodríguez-Ropero, Theoretical Considerations and

Computational Tools, in ß-barrel Channel Proteins as Tools in Nanotechnology 2014,

Springer Netherlands. p. 69

Page 261: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

261

Fitton, P., Rick, E. A. J. Organomet. Chem. 1971, 28, 287

Fleming, I. Frontier Orbitals and Organic Chemical Reactions, Wiley, New York 1976,

161.

Flood, T.C., Jensen, J.E., Statler, J.A. J. Am. Chem. Soc. 1981, 103, 4410

Forkert, P.-G., Lee, R. P. Toxicol. Appl. Pharmacol. 1997, 146, 245

Fox, B. G., Lyle, K. S., Rogge, C. E. Acc. Chem. Res. 2004, 37, 421

Franke, M. Wolak and R. van Eldik, Chem. Eur. J. 2009, 15, 10182

Frisch, J. et al. Gaussian-03, Revision C.02, Wallingford, CT, 2003

Frisch, J. et al. Gaussian-09, Revision A.02, Wallingford, CT, 2009

Frisch, M. J. Gaussian 09, revision C.01, Gaussian, Inc.: Wallingford, CT, 2010

Frisch, M. J. Gaussian 03, revision D.01, Gaussian, Inc., Wallingford, CT, 2004

Fürstner, A. Angew. Chem. Int. Ed. 2000, 39, 3012

Fukuzumi, S., Morimoto, Y., Kotani, H., Naumov, P., Lee, Y.-M., Nam, W. Nature

Chem. 2010, 2, 756

Fukuzumi, S., Ohkubo, K. Coord. Chem. Rev. 2010, 254, 372

Galonić, P., Barr, E. W., Walsh, C. T., Bollinger Jr, J. M., Krebs, C. Nat. Chem. Biol.

2007, 3, 113

García Alonso, F. J., Llamazares, A., Riera, V., Vivanco, M., Diaz M. R., García Granda,

S. J. Chem. Soc. Chem. Commun. 1991, 1058

Garrou, P. E. Chem. Rev. 1985, 85, 171

Page 262: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

262

Gelb, M. H., Heimbrook, D. C., Mälkönen, P., Sligar, S. G. Biochemistry 1982, 21, 370

Ghosh, A., Acc. Chem. Res. 1983, 31, 189

Ghosh, A., Taylor, P. R. Curr. Opin. Chem. Biol. 2003, 7, 113,

Ghosh, S. K., Patra, R., Rath, S. P. Inorg. Chem. 2008, 47, 10196

Ghosh, S. K., Patra, R., Rath, S. P. Inorg. Chem. 2010, 49, 3449

Ghosh, S. K., Patra, R., Rath, S. P. Inorg. Chim. Acta. 2010, 363, 2791

Goel, A. B. Inorg. Chim. Acta, 2010, 84, L25.

Goertz, W., Kamer, P. C. J., van Leeuwen, P. W. N. M., Vogt, D. Chem, Eur. J. 2001,

7,1614

Gonzalez-Diaz, H., et al. Current Topics in Medicinal Chemistry, 2013. 13,14, 1713

Goodson, F. E., Wallow, T. I., Novak, B. M. J. Am. Chem. Soc. 1997, 119, 12441

Green, M. T. J. Am. Chem. Soc. 1999, 121, 7939

Green, M. T. J. Am. Chem. Soc. 2006, 128, 1902

Green, M. T., Dawson, J. H., Gray, H. B. Science 2004, 304, 1653

Grimme, S. J. Comput. Chem. 2006, 27, 1787

Grimme, S., Antony, J., Ehrlich, S., Krieg, H. J. Chem. Phys. 2010, 132, 154104

Grogan, G., Curr. Opin. Chem. Biol. 2011, 15, 241

Gross, Z. J. Biol. Inorg. Chem. 1996, 1, 368

Gross, Z., Nimri, S. Inorg. Chem. 1994, 33, 1731

Groves, J. T. Gross, Z., Stern, M. K. Inorg. Chem. 1994, 33, 5065

Page 263: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

263

Groves, J. T. Models and Mechanisms of Cytochrome P450 Action. In Cytochrome P450:

Structure, Mechanism and Biochemistry. Ortiz de Montellano P.R., (Ed.), 3rd Ed.,

Kluwer Academic/Plenum Publishers, New York, 2005, Chapter 1, pp 1

Groves, J. T. Proc. Natl. Acad. Sci. USA 2003, 100, 3569

Groves, J. T., Ahn, K.-H. Inorg. Chem. 1987, 26, 3831

Groves, J. T., Ahn, K.-H., Quinn, R. J. Am. Chem. Soc. 1988, 110, 4217

Groves, J. T., Avaria-Neisser, G. E., Fish, K. M., Imachi, M., Kuczkowski, R. L. J. Am.

Chem. Soc. 1986, 108, 3837

Groves, J. T., Bonchio, M., Carofiglio, T., Shalyaev, K. J. Am. Chem. Soc. 1996, 118,

8961

Groves, J. T., Gross, Z., Stern, M. K. Inorg. Chem. 1994, 33, 5065

Groves, J. T., McClusky, G. A. J. Am. Chem. Soc. 1976, 98, 859.

Groves, J. T., Myers, R. S. J. Am. Chem. Soc. 1983, 105, 5791

Groves, J. T., Quinn, R. J. Am. Chem. Soc. 1985, 107, 5790

Groves, J. T., Roman, J. C. J. Am. Chem. Soc. 1995, 117, 5594

Groves, J. T., Shalyaev, K., Lee, J. In The Porphyrin Handbook, Kadish, K. M., Smith, K.

M., Guilard, R. Eds., Academic Press: New York, 2000, Vol. 4, Chapter 27, pp. 17

Groves, J. T., Subramanian, D. V. J. Am. Chem. Soc. 2010, 106, 2177

Groves, J. T., Watanabe, Y. J. Am. Chem. Soc. 1986, 108, 507

Groves, T., Stern, M. K., J. Am. Chem. Soc. 1988, 110, 8628

Page 264: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

264

Guan, X., Fisher, M. B., Lang, D. H., Zheng, Y. M., Koop, D. R., Rettie, A. E. Chem.

Biol. Interact. 1983, 110, 103

Guengerich, F. P. Arch. Biochem. Biophys. 2003, 409, 59

Guengerich, F. P. Chem. Res. Toxicol. 2001, 14, 611

Guengerich, F. P. Chem. Res. Toxicol. 2008, 21, 70

Guengerich, F. P., Kim, D. H. Chem. Res. Toxicol. 1991, 4, 413

Guengerich, P. Chem. Res. Toxicol. 2001, 14, 611

Gütlich, P., Hauser, A., Spiering, H. Angew. Chem. Int. Ed. Engl. 1994, 33, 2024

Gunes, A., Bilir, E., Zengil, H., Babaoglu, M. O., Bozkurt, A., Yasar, U. Pharmacol.

Toxicol. 2007, 100, 383

Gupta, A., Gupta, G. P. J. Phys. Condens. Matter 1994, 6, 11071

Gupta, R., Lacey, D. C., Bominaar, E. L., Borovik, A. S., Hendrich, M. P. J. Am. Chem.

Soc. 2012, 134, 9775

Hagen, W. R., van den Berg., van Dongen W. A. M., Reijerse W. M. A. M., E. J., van

Kan, P. J. M. J. Chem. Soc. Farad. Trans. 1983, 94, 2969

Handy, N. C., Cohen, A. J. Mol. Phys. 2001, 99, 403

Harding, D. J., Kerpal, C., Meijer, G., Fielicke, A. J. Phys. Chem. Lett. 2013, 4, 892

Harrison, J.G., 1987. “Electron affinities in the self-interaction-correlated local spin

density approximation”. J. Chem. Phys, 86, 2849.

Hay, P. J., Wadt, W. R. J. Chem. Phys. 1985, 82, 270

Hehre, W. J., Ditchfield, K., Pople, J. A. J. Chem. Phys. 1972, 56, 2257

Page 265: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

265

Hellmann, H. J. Chem. Phys., 1935, 3, 61

Hessenauer-Ilicheva, N., Franke, A., Meyer, D., Woggon, W.-D. van Eldik, R. J. Am.

Chem. Soc. 2007, 129, 12473

Hessenauer-Ilicheva, N., Franke, A., Meyer, D., Woggon, W.-D., van Eldik, R. Chem.

Eur. J. 2009, 15, 2941

Hessenauer-Ilicheva, N., Franke, A., Meyer, D., Woggon, W.-D., van Eldik, R. J. Am.

Chem. Soc. 2007, 129, 12473

Hessenauer-Ilicheva, N., Franke, A., Wolak, M., Higuchi, T., van Eldik, R. Chem. Eur. J.

2009, 15, 12447

Heyes, D. J., Sakuma, M., de Visser, S. P., Scrutton, N. S. J. Biol. Chem. 2009, 284, 3762

Hickey, C. E., Maitlis, P. M. J. Chem. Soc. Chem. Commun. 2010, 1609

Hirao, H., Kumar, D., Que Jr, L., Shaik, S. J. Am. Chem. Soc. 2006, 128, 8590

Hirao, H., Kumar, D., Thiel, W., Shaik, S. J. Am. Chem. Soc. 2005, 127, 13007

Hirao, H., Que Jr, L., Nam, W., Shaik, S. Chem. Eur. J. 2008, 14, 1740

Hoe, W.-M., Cohen, A., Handy, N. C. Chem. Phys. Lett. 2001, 341, 319

Hoffart, L. M., Barr, E. W., Guyer, R. B., Bollinger Jr, J. M., Krebs C. Proc. Natl. Acad.

Sci. USA 2006, 103, 14738

Hohenberg, P., Kohn, W., Phys. Rev. B, 1964, 136, 864

Holm, R., Kennepohl, P., Solomon, E. I., 1996, Chemistry Reviews, 96, 2239

Hückel, Z. Physik, 1931, 70, 204.

Page 266: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

266

Hull, J. F., Balcells, D., Sauer, E. L. O., Raynaud, C., Brudvig, G. W., Crabtree, R. H.,

Eisenstein, O. J. Am. Chem. Soc. 2010, 132, 7605

Hull, J. F., Sauer, E. L. O., Incarvito, C. D., Faller, J. W., Brudvig, G. W., Crabtree, R. H.

Inorg. Chem. 2009, 48, 488

Hurst, J. K. Coord. Chem. Rev. 2005, 249, 313

Huynh, M. H. V., White, P. S., Carter, C. A., Meyer, T. J. Angew. Chem., Int. Ed. 2001,

40, 3037

Ikariya, T., Murata, K., Noyori, R. Org. Biomol. Chem. 2006, 4, 393

Inuki, T., Kojima, T. J. Org. Chem. 1967, 32, 869

Ishihara, K., Kurihara, H., Matsumoto, M., Yamamoto, H. J. Am. Chem. Soc. 1983, 120,

6920

Ishihara, K., Ohara, S., Yamamoto, H. Science 2000, 290, 1140

Isobe, H., Yamanaka, S., Okumura, M., Yamaguchi, K., Shimada, J. J. Phys. Chem. B

2011, 115, 10730

Itoh, S. Curr. Opin. Chem. Biol. 2006, 10, 115

Jackson, T. A., Rohde, J.-U., Seo, M. S., Sastri, C. V., DeHont, R., Stubna, A., Ohta, T.,

Kitagawa, T., Münck, E., Nam, W., Que Jr, L. J. Am. Chem. Soc. 2008, 130, 12394

Jaguar 7.6, Schrödinger LLC, New York NY, 2007

Jaguar 7.7, version 7.9, Schrodinger, LLC, New York, NY, 2011

Jaguar 7.8, Schrödinger LLC, New York NY, 2010

Page 267: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

267

Jaguar, version 7.9, Schrödinger, LLC, New York, NY, 2011

James, B. R., Mahajan, D. Can. J. Chem. 1979, 57, 180

Jensen, F. Introduction to Computational Chemistry, 2nd

Ed., John Wiley & Sons, Ltd.,

Chichester, 2007

Jensen, H. Molecular Modeling Basics. CRC Press, Boca Raton, 2010

Ji, L., Franke, A., Brindell, M., Oszajca, M., Zahl, A., van Eldik, R. Chem. Eur. J. 2014,

20, 2347.

Ji, L., Schüürmann, G. Angew. Chem. Int. Ed. 2013, 52, 744

Ji, L., Zhang, J., Liu, W., de Visser, S. P. Chem. Asian J. 2014, 9, 1175

Johansson, J., Blomberg, M. R. A., Siegbahn, P. E. M. J. Phys. Chem. C 2007, 111,

12397

Johnston, J. B., Kells, P. M., Podust, L. M., Ortiz de Montellano, P. R. Proc. Natl. Acad.

Sci. USA. 2009, 106, 20687

Kadish, K. M., Smith, K. M., Guilard, R. Eds. Handbook of Porphyrin Science, World

Scientific Publishing Co.: New Jersey, 2010

Kahn, O. Molecular Magnetism, VCH Publishers: Weinheim, 1993, 2

Kaizer, J., Klinker, E. J., Oh, N.Y., Rohde, J.-U., Song, W. J., Stubna, A., Kim, J., Münck

E., Nam, W., Que Jr, L. J. Am. Chem. Soc. 2004, 126, 472

Kamachi, T., Shiota, Y., Ohta, T., Yoshizawa, K. Bull. Chem. Soc. Jpn. 2003, 76, 721

Kamachi, T., Yoshizawa, K. J. Am. Chem. Soc. 2003, 125, 4652

Page 268: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

268

Kaminsky, L. S., Fasco, M. J., Guengerich, F. P. J. Biol. Chem. 1980, 255, 85

Kaneda, K., Sano, K., Teranishi, S. Chem. Lett. 1979, 82

Karamzadeh, B., Kumar, D., Sastry, G. N., de Visser, S. P. J. Phys. Chem. A 2010, 114,

13234

Kataoka, Y., Shibahara, A., Yamagata, T., Tani, K. Organometallics 2001, 20, 2431

Kelly, S. L., Lamb, D. C., Kelly, D. E. FEBS. Lett. 1997, 412, 233

Kenneth S. K. 1987. Introductory Nuclear Physics. Wiley. ISBN 978-0-471-80553-3.

Kikukawa, K., Takagi, M., Matsuda, T. Bull. Chem. Soc. Japan, 1979, 52, 1493.

Kitchen, J. A., Brooker, S. Coord. Chem. Rev. 2008, 252, 2072

Kitchen, J. A., White, N. G., Jameson, G. N. L., Tallon, J. L., Brooker, S. Inorg. Chem.

2011, 50, 4586

Kohn, W., Sham, L. J. Phys. Rev. A, 1965, 140, 1133

Kong, K.-C., Cheng, C.-H. J. Am. Chem. Soc. 1991, 113, 6313

Korendovych, I. V., Kryatov, S. V., Rybak-Akimova, E. V. Acc. Chem. Res. 2007, 40,

510

Kovaleva, G., Neibergaul, M.B., Chakrabarty, S., Lipscomb, J. D., Acc. Chem. Res. 2007,

40, 475

Kratsch, J., Roesky, P.W. Angewandte Chemie International Edition, 2014. 53, 2, 376

Page 269: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

269

Krebs, C., Dassama, L. M. K., Matthews, M. L., Jiang, W., Price, J. C., Korboukh, V., Li,

N., Bollinger Jr, J. M. Coord. Chem. Rev. 2013, 257, 234

Kryatov, S. V., Rybak-Akimova, E. V., Schindler, S. Chem. Rev. 2005, 105, 2175

Kumar, A., Goldberg, I., Botoshansky, M., Buchman, Y., Gross, Z. J. Am. Chem. Soc.

2010, 132, 15233

Kumar, D, Latifi, R., Kumar, S., Rybak-Akimova, E. V., Sainna, M. A., de Visser, S. P.

Inorg. Chem. 2013, 52, 7968

Kumar, D., de Visser, S. P., Shaik S. Chem. Eur. J. 2005, 11, 2825

Kumar, D., de Visser, S. P., Shaik S. J. Am. Chem. Soc. 2003, 125, 13024

Kumar, D., de Visser, S. P., Shaik, S., J. Am. Chem. Soc. 2004, 126, 5072

Kumar, D., de Visser, S. P., Sharma, P. K., Cohen, S., Shaik, S. J. Am. Chem. Soc. 2004,

126, 1907

Kumar, D., de Visser, S. P., Sharma, P. K., Derat, E., Shaik, S. J. Biol. Inorg. Chem.

2005, 10, 181

Kumar, D., Derat, E., Khenkin, A. M., Neumann, R., Shaik, S. J. Am. Chem. Soc. 2005,

127, 17712

Kumar, D., Hirao, H., Que Jr, L., Shaik, S. J. Am. Chem. Soc. 2005, 127, 8026

Kumar, D., Karamzadeh, B., Sastry, G. N., de Visser, S. P. J. Am. Chem. Soc. 2010, 132,

7656

Kumar, D., Sastry, G. N, de Visser, S. P., J. Phys. Chem. B 2012, 116, 718

Kumar, D., Sastry, G. N., de Visser, S. P. Chem. Eur. J. 2011, 17, 6196

Page 270: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

270

Kumar, D., Tahsini, L., de Visser, S.P., Kang, H. Y., Kim, S. J., Nam, W. J. Phys. Chem.

A 2009, 113, 11713

Kumar, D., Thiel, W., de Visser, S. P. J. Am. Chem. Soc. 2011, 133, 3869

Kumar, S., Faponle, A. S., Barman, P., Vardhaman, A. K., Sastri, C. V., Kumar. D., de

Visser, S. P. J. Am. Chem. Soc. 2014, in press, DOI 10.1021/ja508403w.

Kurahashi, T., Kikuchi, A., Tosha, T., Shiro, Y., Kitagawa, T., Fujii, H. Inorg. Chem.

2008, 47, 1674

Lafite, P., André, F., Zeldin, D. C., Dansette, P. M., Mansuy, D. Biochemistry 2007, 46,

10237

Lansky, D. E., Goldberg, D. P., Inorg. Chem. 2006, 45, 5119

Lanucara, F., Chiavarino, B., Crestoni, M.E., Scuderi, D., Sinha, R.K., Maitre P.,

Fornarini S., Inorg. Chem. 2011, 50, 4445

Lanucara, F., Crestoni, M. E. Chem. Eur. J. 2011, 17, 12092

Latifi, R., Sainna, M. A., Rybak-Akimova, E. V., de Visser, S. P. Chem. Eur. J. 2013,

19, 4058

Latifi, R., Bagherzadeh, L., de Visser, S. P. Chem. Eur. J. 2009, 15, 6651

Latifi, R., J. S. Valentine, W. Nam, S. P. de Visser, Chem. Commun. 2012, 48, 3491

Latifi, R., Tahsini L., B. Karamzadeh, N. Safari, W. Nam, S. P. de Visser, Arch. Biochem.

Biophys. 2011, 507, 4

Latifi, R., Tahsini L., Kumar, D., Sastry, G. N., Nam, W., de Visser, S. P. Chem.

Commun. 2011, 47, 10674

Page 271: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

271

Lee, C., Yang, W., Parr, R. G. Phys. Rev. B 1988, 37, 785

Lee, R. P., Parkinson, A., Forkert, P.-G. Drug Metabol. Disp. 1983, 26, 60

Leeladee, P., Baglia, R. A., Prokop, K. A., Latifi, R., de Visser, S. P., Goldberg, D. P. J.

Am. Chem. Soc. 2012, 134, 10397

Leeladee, P., Goldberg, D. P. Inorg. Chem. 2010, 49, 3083

Leithauser, M. T., Liem, A., Stewart, B. C., Miller, E. C., Miller, J. A. Carcinogenesis

1990, 11, 463

Lewis, J.C. ACS Catal 2013, 3, 12, 2954

Li, J., Lord, R. L., Noll, B. C., Baik, M.-H., Schulz, C. E., Scheidt, W. R. Angew. Chem.

Int. Ed. 2008, 47, 10144

Lias, S. G., Bartmess, J. E., Liebman, J. F., Holmes, J. L., Levin, R. D., Mallard, W. G.,

In Ion Energetics Data in NIST Chemistry WebBook, Linstrom, P. J., Mallard, W. G.

(Eds.), NIST Standard Reference Database Number 69, National Institute of Standards

and Technology: Gaithersburg, MD, 20899 (http://webbook.nist.gov).

Lide, D. R. (Ed) Handbook of Chemistry and Physics, CRC Press, Boca Raton, New

York, 1996

Lin, Y. A., Davis, B. G. Beilstein J. Org. Chem. 2010, 6, 1219

Linde, C., Åkermark, B., Norrby, P.-O., Svensson, M. J. Am. Chem. Soc. 1999, 121, 5083

Ling, Y., Zhang, Y. J. Am. Chem. Soc. 2009, 131, 6386

Page 272: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

272

Loch, J. M., Potter, J., Bachmann, K. A. Pharmacol. 1995, 50, 146

Lonsdale, R., Harvey, J. N., Mulholland, A. J. J. Phys. Chem. Lett. 2010, 1, 3232

Lonsdale, R., Oláh, J., Mulholland, A. J., Harvey, J. N. J. Am. Chem. Soc. 2011, 133,

15464

Love, J. A., Sanford, M. S., Day, M. W., Grubbs, R. H. J. Am. Chem. Soc. 2003, 125,

10103

Lowe, John P. Quantum Chemistry, Academic Press: New York, 1978, pp 315

Lu Y, Yeung N, Sieracki N, Marshall N., M. Nature 2009, 460,855

Lundquist, E. G., Folting, K., Huffman, J.C., Caulton, K.G. Organometallics 1990, 7,

2254.

Luo, X.-L., Crabtree, R. H. J. Am. Chem. Soc. 1989, 111, 2527

MacGregor, S. A., McKay, D., Panetier, J. A., Whittlesey, M. K. Dalton Trans. 2013, 42,

7386

Mader, A., Manner, V. W., Markle, T. F., Wu, A., Franz, J. A., Mayer, J. M. J. Am.

Chem. Soc. 2009, 131, 4335

Maestri, A. G., Cherry, K. S., Toboni, J. J., Brown, S. N. J. Am. Chem. Soc. 2001, 123,

7459

Major, A., Horváth, I. T., Pino, P. J. Mol. Catal. 1988, 45, 275

Man, W.-L., Lam, W. W. Y., Lau, T.-C. Acc. Chem. Res. 2014, 47, 427

Page 273: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

273

Manner, V. W., Lindsay, A. D., Mader, E. A., Harvey, J. N., Mayer, J. M., Chem. Sci.

2012, 3, 230

Marcone, J. E., Moloy, K. G. J. Am. Chem. Soc. 1983, 120, 8527

Marshall G., Hendrickson C.L., Jackson G.S., Mass Spec. Rev. 1983, 17, 1G. Marshall, C.

L. Hendrickson and G. S. Jackson, Mass Spec. Rev. 1983, 1

Martinho, M., Banse, F., Bartoli, J.-F., Mattioli, T. A., Battioni, P., Horner, O., Bourcier,

S., Girerd, J.-J. Inorg. Chem. 2005, 44, 9592

Mas-Balleste, R., Que Jr, L. J. Am. Chem. Soc. 2007, 129, 15964

Maseras, F., Lledós, A. Computational Modeling of Homogeneous Catalysis, Kluwer

Academic Publishers: Dordrecht, The Netherlands, 2002.

Matsuo, T., Mayer, J. M. Inorg. Chem. 2005, 44, 2150

Mawby, R.J., Basolo, F., Pearson, R.G. J. Am. Chem. Soc. 1964, 86, 5043

Mayer, J. M. Acc. Chem. Res. 1983, 31, 441

Mayer, M. Annu. Rev. Phys. Chem. 2004, 55, 363

McGown, A. J., Kerber, W. D., Fujii, H., Goldberg, D. P. J. Am. Chem. Soc. 2009, 131,

8040

McLean, K. J., Munro, A. W. Drug Metabol. Rev. 2008, 40, 427

Meot-Ner, M., In Gas Phase Ion Chemistry. M. T. Bowers (Ed.), Academic Press, New

York, 1979, Vol. 1

Meunier, B., de Visser, S. P., Shaik, S. Chem. Rev. 2004, 104, 3947

Meyer, T. J., Huynh,M. H. V. Inorg. Chem. 2003, 42, 8140

Page 274: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

274

Miller, R. L., Wolczanski, P. T., Rheingold, A. L. J. Am. Chem. Soc. 1993, 115, 10422

Millikan, R.A., Physical Review. 1911, 32,349

Mirbach, M. F. J. Metalorg. Chem. 2010, 265, 205.

Moënne-Loccoz, P., Krebs, C., Herlihy, K., Edmondson, D. E., Theil, E. C., Huynh, B.

H., Loehr, T. M. Biochemistry 1999, 38, 5290

Mole, L., Spencer, J. L., Carr, N., Orpen, A. G. Organometallics, 1991, 10, 49

Molecular Quantum Mechanics Parts I and II: An Introduction to QUANTUM

CHEMISTRY (Volume 1), Atkins, P.W. Oxford University Press, 1977

Møller, C., Plesset, M. S. Phys. Rev., 1934, 46, 0618

Monsaert, S., Lozano Vila, A., Drozdzak, R., van der Voort, P., Verpoort, F. Chem. Soc.

Rev. 2009, 38, 3360

Munro, A. W., Girvan, H. M., McLean, K. J. Nat. Prod. Rep. 2007, 24, 585

Murphy, C. Biotechnology Letters, 2014, 1

Murray, L. J., Lippard, S. J. Acc. Chem. Res. 2007, 40, 466

Mutter, S. T., Platts, J. A. J. Phys. Chem. A 2011, 115, 11293

Nagata, K., Liberato, D. J., Gillette, J. R., Sasame, H. A. Drug Metab. Dispos. 1986, 14,

559

Nakamura, K., Ikezaki, A., Ohgo, Y., Ikeue, T., Neya, S., Nakamura, M. Inorg. Chem.

2008, 47, 10299

Page 275: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

275

Nakamura, M., Ohgo, Y., Ikezaki, A., Electronic and Magnetic Structures of Iron

Porphyrin Complexes, In Handbook of Porphyrin Science, Kadish, K.M., Smith, K.M.,

Guilard, R. (eds.), World Scientific, 2010, 1

Nam, W. Acc. Chem. Res. 2007, 40, 522

Nam, W., et al. Am. Chem. Soc. 2000. 122, 28, 6641

Nam, W., Ryu, Y.O., Song, W. J. J. Biol. Inorg. Chem. 2004, 9, 654

Neese, ORCA – an ab initio, DFT and semiempirical SCF-MO package, Bonn, Germany,

Version 2.7-00, 2009

Nehru, K., Kim, S. J., Kim, I. Y., Seo, M. S., Kim, Y., Kim, S.-J., Kim, J., and Nam, W.

Chem. Commun. 2007, 4623

Nelson, D. R. Human Genomics 2009, 4, 59

Nelson, D.R., Methods Mol. Biol. 2006, 320, 1

Neu, H. M., Quesne, M.G., Yang, T., Prokop-Prigge, K.,A., Lancaster K.,M., Donohoe J.,

DeBeer, S., de Visser, S. P., Goldberg, D. P. Chem. Eur. J. 2014, 20, 14584

Nibbering, N. M. M. Mass Spec. Rev. 2006, 25, 962

Nihei, M., Shiga, T., Maeda, Y., Oshio, H. Coord. Chem. Rev. 2007, 251, 2606

Noack, K.,Calderazzo, F. J. Organometal. Chem. 1967, 10, 101

Nordlund, P., Reichard, P. Annu. Rev. Biochem. 2006, 75, 681

Nyman, G., Int. J. of Quan. Chem., 2014. 114, 1183

O’Brien, J. R., Schuller, D. J., Yang, V. S., Dillard, B. D., Lanzilotta W. N. Biochemistry

2003, 42, 5547

Page 276: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

276

O’Connor, C. J. Prog. Inorg. Chem. 1982, 29, 203.

Ogliaro, F., de Visser, S. P., Cohen, S., Kaneti, J., Shaik, S. ChemBioChem. 2001, 11, 848

Ogliaro, F., Harris, N., Cohen, S., Filatov, M., de Visser, S. P., Shaik, S. J. Am. Chem.

Soc. 2000, 122, 8977

Ohno, T., Suzuki, N., Dokoh, T., Urano, Y., Kikuchi, K., Hirobe, M., Higuchi, T.,

Nagano, T. J. Inorg. Biochem. 2000, 82, 123,

Omura, T., Sato, R. J. Biol. Chem. 1962, 237, 1375

O'Reilly, E., Koehler, V., Flitsch, S., Turner, N. Chem. Commun. 2011, 47, 2490

Ortiz de Montellano, P. R. Chem. Rev. 2010, 110, 932

Ortiz de Montellano, P. R. Ed. Cytochrome P450: Structure, Mechanism and

Biochemistry. 3rd ed., Kluwer Academic/Plenum Publishers, New York, 2004

Ortiz de Montellano, P. R., Fruetel, J. A., Collins, J. R., Camper, D. L., Loew, G. H. J.

Am. Chem. Soc. 1991, 113, 3195

Ostović, D., Bruice, T.C., J. Am. Chem. Soc. 1989, 111, 6511

Pan, Z., Zhang, R., Newcomb, M. J. of Inorg. Biochem, 2006,100, 524

Parsell, T. H., Behan, R. K., Green, M. T., Hendrich, M. P., Borovik, A. S. J. Am. Chem.

Soc. 2006, 128, 8728

Parsell, T. H., Yang, M.-Y., Borovik, A. S. J. Am. Chem. Soc. 2009, 131, 2762

Patra, S. K., Bera, J. K. Organometallics 2006, 25, 6054

Patra, S. K., Sadhukhan, N., Bera, J. K. Inorg. Chem. 2006, 45, 4007

Perdew, J. P. Phys. Rev. B 1986, 33, 8822

Page 277: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

277

Perdew, J. P., Burke, K., Ernzerhof, M. Phys. Rev. Lett. 1996, 77, 3865

Perdew, J. P., Wang, Y. Phys. Rev. B 1992, 45, 13244

Peters, J. C., Odom, A. L., Cummins, C. C. Chem. Commun. 1997, 1995

Petrik, I. D., Liu, J., Lu, Y. Curr Opin Chem Biol 2014, 19, 0, 67

Pfaff F., Kundu S., Risch M., Pandian S., Heims F., Pryjomska-Ray I., Haack P.,

Metzinger R., Bill E., Dau H., Comba P., Ray K., Angew. Chem. Int. Ed. 2011, 50, 1711

Posner, G. H., O’Neill, P. M. Acc. Chem. Res. 2004, 37, 397

Poulos, T. L. Peroxidases and Cytochrome P450. In The Porphyrin Handbook, Kadish,

K. M., Smith, K. M., Guilard, R. Eds., Academic Press, San Diego, 2000: Vol. 4, Chapter

32, 189

Poulos, T. L., Finzel, B. C., Gunsalus, I. C., Wagner, G. C., Kraut, J. J. Biol. Chem. 1985,

260, 6122

Poulos, T. L., Finzel, B. C., Howard, A. J. Biochemistry 1986, 25, 5314.

Prat, I., Mathieson, J. S., Güell, M., Ribas, X., Luis, J. M., Cronin, L., Costas, M. Nature

Chem. 2011, 3, 788

Price, J. C., Barr, E. W., Tirupati, B., Bollinger Jr, J. M., Krebs, C. Biochemistry 2003,

42, 7497

Prokop, K. A., de Visser, S. P., Goldberg, D. P. Angew. Chem. Int. Ed. 2010, 49, 5091

Prokop, K. A., Neu, H. M., de Visser, S. P., Goldberg, D. P. J. Am. Chem. Soc. 2011, 133,

15874

Page 278: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

278

Quesne, M. G., Latifi, R., Gonzalez-Ovalle, L. E., Kumar, D., de Visser, S. P. Chem. Eur.

J. 2014, 20, 435

Quiñinero, D., Musaev, D. G., Morokuma, K. Inorg. Chem. 2003, 42, 8449

Quintanar, L., Stoj, C., Taylor, A. B., Hart, P. J., Kosman, D. J., Solomon, E. I. Acc.

Chem. Res. 2007, 40, 445

Radoń, M., Broclawic, E., Pierloot, K. J. Chem. Theory Comput. 2011, 7, 898

Rath, S. P., Olmstead, M. M., Balch, A. L. Inorg. Chem. 2006, 45, 6083

Raven, L. Nat. Prod. Rep. 2003, 20, 367

Reed, C. A., Mashiko, T., Bentley, S. P., Kastner, M. E., Scheidt, W. R., Spartalian, K.,

Lang, G. J. Am. Chem. Soc. 1979, 101, 2948

Reetz, M. T., Kyung, S. H., Bolm, C., Zierke, T. Chem. & Ind. (London) 1986, 824

Reetz, M.T., Angewandte Chemie International Edition, 2014. 53, 33, 8562

Reiher, M., Salomon O., Hess, B. A. Theor. Chem. Acc., 2001, 107, 48

Rettie, A. E., Boberg, M., Rettenmeier, A. W., Baillie, T. A. J. Biol. Chem. 1988, 263,

13733

Rettie, A. E., Rettenmeier, A. W., Howald, W. N., Baillie, T. A. Science 1987, 235, 890

Reynald, R. L., Sansen, S., Stout, C. D., Johnson, E. F. J. Biol. Chem. 2012, 287, 44581

Riley, K. E., Op’t Holt, B. T., Merz, K. M. J. J. Chem. Theory Comput., 2007, 3, 407

Rittle, J., Green, M. T. Science 2010, 330, 933

Roberts, K. M., Jones, J. P. Chem. Eur. J. 2010, 16, 8096

Page 279: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

279

Rohde, J-U., In J-H., Lim, M. H., Brennessel, W. W., Bukowski, M. R., Stubna, A.,

Münck, E., Nam, W., Que Jr, L. Science 2003, 299, 1037

Roithová, J., Schröder, D. Chem. Rev. 2010, 110, 1170

Romain, S., Vigara, L., Llobet, A. Acc. Chem. Res. 2009, 42, 1944

Rude, M. A., Baron, T. S., Brubaker, S., Alibhai, M., Del Cardayre, S. B., Schirmer, A.

Appl. Env. Microbiol. 2011, 77, 1718

Ruettinger, R. T., Fulco, A. J. J. Biol. Chem. 1981, 256, 5728

Ruettinger, R. T., Wen, L. P., Fulco, A. J. J. Biol. Chem. 1989, 264, 10987

Sadeque, A. J. M., Fisher, M. B., Korzekwa, K. R., Gonzalez, F. J., Rettie, A. E. J.

Pharmacol. Exp. Therap. 1997, 283, 698

SAINT+, 6.02 ed., Bruker AXS, Madison, WI, 1999

Sakakura, T. J. Organometal. Chem. 2010, 267, 171

Sala, X., Romero, I., Rodríguez, M., Escriche, L., Llobet, A. Angew. Chem. Int. Ed. 2009,

48, 2842

Saltzman, H., Sharefkin, J.G. Organic Syntheses. Wiley. New York. 1973, Collect. Vol. 5,

658

Sastri, V., Lee, L., Oh, K., Lee, Y. J., Lee, J., Jackson, T. A., Ray, K., Hirao, H., Shin,

W., Halfen, J. A., Kim, J., Que Jr, L., Shaik, S., Nam, W. Proc. Natl. Acad. Sci. USA

2007, 104, 19181

Sastri, V., Seo, M. S., Park, M. J., Kim, K. M., Nam, W. Chem. Commun. 2005, 1405

Schaefer, A., Huber, C., Ahlrichs, R. J. Chem. Phys. 1994, 100, 5829

Page 280: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

280

Scheidt, W. R., Cheng, B., Safo, M. K., Cukiernik, F., Marchon, J.-C., Debrunner, P. G.

J. Am. Chem. Soc. 1992, 114, 4420

Schindelin, H., Kisker, C., Schlessman, J.L., Howard, J.B., Rees, D.C. Nature, 1997,

387, 370

Schlichting, I., Berendzen, J., Chu, K., Stock, A. M., Maves, S. A., Benson, D. E., Sweet,

R. M., Ringe, D., Petsko, G. A., Sligar, S. G. Science 2000, 287, 1615

Schrock, R. R. Acc. Chem. Res. 1979, 12, 98

Schrödinger, E. Phys. Rev., 1926, 28, 1049

Schünemann, V., Gerdan, M., Trautwein, A. X., Haoudi, N., Mandon, D., Fischer, J.,

Weiss, R., Tabard, A., Guilard, R. Angew. Chem., Int. Ed. Engl. 1999, 38, 3181

Schwabe, T., Grimme, S. Acc. Chem. Res. 2008, 41, 569,

Schwabe, T., Grimme, S. Phys. Chem. Chem. Phys. 2007, 9, 3397

Shaik S., Chen H., Janardanan, D. Nature Chem. 2011, 3, 19

Shaik S., de Visser, S. P., Ogliaro, F., Schwarz H., Schröder, D. Curr. Opin Chem. Biol.

2002, 6, 556

Shaik S., Kumar D., de Visser, S. P. J. Am. Chem. Soc. 2008, 130, 10128

Shaik, S. S. J. Am. Chem. Soc. 1981, 103, 3692

Shaik, S., de Visser, S.P., Kumar, D. J. Bio. Inorg. Chem. 2004. 9, 6, 661

Shaik, S., de Visser, S. P., Kumar, D. J. Am. Chem. Soc. 2004, 126, 11746

Shaik, S., de Visser, S. P., Ogliaro, F., Schwarz, H., Schröder, D. Curr. Opin. Chem. Biol.

2002, 6, 556

Page 281: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

281

Shaik, S., et al. Chem. Rev, 2005. 105,6, 2279

Shaik, S., Filatov, M., Schröder, D., Schwarz, H. Chem. Eur. J. 1983, 4, 193

Shaik, S., Hiberty, P. C. A Chemist’s Guide to Valence Bond Theory, Wiley-Interscience,

New York, NY, 2007

Shaik, S., Kumar, D., de Visser, S. P. J. Am. Chem. Soc. 2008, 130, 10128

Shaik, S., Kumar, D., de Visser, S. P., Altun, A., Thiel W. Chem. Rev. 2005, 105, 2279

Shaik, S., Lai, W., Chen, H., Wang, Y. Acc. Chem. Res. 2010, 43, 1154

Shanklin, J., Somerville, C. Proc. Natl. Acad. Sci. USA 1991, 88, 2510

Sharma, P. K., de Visser, S. P., Ogliaro, F., Shaik, S. J. Am. Chem. Soc. 2003, 125, 2291

Shelby, M.L., Mara M.W., Chen, L.X. Coord. Chem. Rev, 2014. 277–278, 291

Sheldrick, G. M. SADABS 2.0, 2000

Sheldrick, G. M. SHELXL-2014: Program for Crystal Structure Refinement, University

of Göttingen: Göttingen, Germany, 2014

Sherwood, P., et al. J. Mol. Struct. 2003, 632, 1

Shi, S., Klotz, U. Clinical Pharmacokinetics, 2012. 51, 2,77

Shinkyo, R., Xu, L., Tallman, K. A., Cheng, Q., Porter, N. A., Guengerich, T. P. J. Biol.

Chem. 2011, 286, 33021

Shokri, A., Abedin, A., Fattahi, A., Kass, S. R. J. Am. Chem. Soc. 2012, 134, 10646

Shook, R. L., Peterson, S. M., Greaves, J., Moore, C., Rheingold, A. L., A. Borovik, S. J.

Am. Chem. Soc. 2011, 133, 5810

Page 282: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

282

Siegbahn, P. E. M., Blomberg, M. R. A. Chem. Rev. 2010, 110, 7040

Siegbahn, P. E. M., Blomberg, M. R. A. J. Phys. Chem. A 2008, 112, 12772

Siegbahn, P. E. M., Borowski, T. Acc. Chem. Res. 2006, 39, 729

Sil, D., Khan, F. S. T., Rath, S. P. Inorg. Chem. 2014, 53, 11925

Sisak, A., Ungváry, F., Kiss, G. J. Mol. Catal. 1983, 18, 223

Slater, C., The self-consistent field for molecular and solids. In Quantum Theory of

Molecular and Solids, McGraw-Hill, New York, 1974, Vol. 4.

Smith, W., Forester, T. R. J. Mol. Graphics 1996, 14, 136

Song, J., Seo, M. S., DeBeer George, S., Ohta, T., Song, R., Kang, M-J., Tosha, T.,

Kitagawa, T., Solomon, E. I., Nam, W. J. Am. Chem. Soc. 2007, 129, 1268

Song, W. J., Ryu, Y. O., Song, R., Nam, W. J. Biol. Inorg. Chem. 2005, 10, 294

Sono, M., Roach, M. P., Coulter, E. D., Dawson, J. H. Chem. Rev. 1996, 96, 2841

Stenkamp, R. E. Chem. Rev. 1994, 94, 715

Stephenson, N. A., Bell, A. T. Inorg. Chem. 2006, 45, 2758

Su, C.-J., Su, P.-C., Chi, Y., Peng, S.-M., Lee, G.-H. J. Am. Chem. Soc. 1996, 118, 3289

Su, T., Chesnavich, W. J. J. Chem. Phys. 1982, 76, 5183

Swart, M., Groenhof, A. R., Ehlers, A. W., Lammertsma, K. J. Phys. Chem. A 2004, 108,

5479

Takahashi, A., Kurahashi, T., Fujii, H. Inorg. Chem. 2011, 50, 6922

Page 283: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

283

Takahashi, A., Yamaki, D., Ikemura, K., Kurahashi, T., Ogura, T., Hada, M., Fujii, H.

Inorg. Chem. 2012, 51, 7296

Takemoto, S., Morita, H., Karitani, K., Fujiwara, H., Matsuzaka, H. J. Am. Chem. Soc.

2009, 131, 18026

Tani, M. Matsu-ura, S. Nakayama, M. Ichimura, N. Nakamura, Y. Naruta, J. Am. Chem.

Soc. 2001, 123, 1133

Tao, J. M., Perdew, J. P., Staroverov, V. N., Scuseria, G. E. Phys. Rev. Lett. 2003, 91,

146401

Thomann, H., Bernardo, M., Goldfarb, D., Kroneck, P. M. H., Ullrich, V. J. Am. Chem.

Soc. 1995, 117, 8243

Thomas, J.M. Phy. Chem. Chem. Phys, 2014. 16, 17, 7647

Thorn D. L., Hoffmann, R. J. Am. Chem. Soc. 1978,100, 2079

Tomasi, J., Mennucci, B., Cammi, R. Chem. Rev., 2005, 105, 2999

Tomter, A. B., Zoppellaro, G., Andersen, N. H., Hersleth, H.-P., Hammerstad, M., Røhr,

Å. K., Sandvik, G. K., Strand, K. R., Nilsson, G. E., Bell III, C. B., Barra, A.-L., Blasco,

E., Le Pape, L., Solomon, E. I., Andersson, K. K. Coord. Chem. Rev. 2013, 257, 3

Tsipis, A.C. Coord. Chem. Rev, 2014. 272, 0, 1

TURBOMOLE V6.2 2010, a development of University of Karlsruhe and

Forschungszentrum Karlsruhe GmbH, 1989-2007, TURBOMOLE GmbH, since 2007,

available from http://www.turbomole.com.

Ueki, M., Matsumoto, Y., Jodry, J. J., Mikami, K. Synlett 2001, 1889

Page 284: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

284

Ungváry, F. J. Metalorg. Chem. 1972, 36, 363

Usharani, D., Lai, W., Li, C., Chen, H., Danovich, D., Shaik, S. Chem. Soc. Rev. 2014,

43, 4968

Valdez, C. E., Smith, Q. A., Nechay, M. R., Alexandrova, A. N. Acc Chem Res 2014, 47,

3110

van Eldik, R. Coord. Chem. Rev. 2007, 251, 1649

van Leeuwen, P. W. N. M. Appl. Catal. A: General 2001, 212, 61

van Leeuwen, P. W. N. M., Morokuma, K., van Lenthe, J. H. Eds., Theoretical Aspects of

Homogeneous Catalysis, Kluwer Academic Publishers: Dordrecht, The Netherlands,

1995

van Leeuwen, P. W. N. M., Roobeek, C. F., Orpen, A. G. Organometallics, 1990, 9, 2179

van Leeuwen, P. W. N. M., Roobeek, C. F., van der Heijden, H. J. J. Am. Chem. Soc.

1994, 116, 12117

Van Leeuwen, P. W. N. M., Roobeek, C. F., Wife, R. L., Frijns, J. H. G. J. Chem. Soc.

Chem. Commun. 1986, 31

Vardhaman, A. K., Barman, P., Kumar, S., Sastri, C. V., Kumar, D., de Visser, S. P.

Angew. Chem. Int. Ed. 2013, 52, 12288

Vardhaman, A. K., Sastri, C. V., Kumar, D., de Visser, S. P. Chem. Commun. 2011, 47,

11044

Vardhaman, K., Barman, P., Kumar, S., Sastri, C. V., Kumar, D., de Visser, S. P. Angew.

Chem. Int. Ed. 2013, 52, 12288

Page 285: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

285

Vardhaman, K., Sastri, C. V., Kumar, D., de Visser, S. P. Chem. Commun. 2011, 47,

11044

Vaska, L. Acc. Chem. Res. 1968, 1,335

Vaska, L., Diluzio, J. W. J. Am. Chem. Soc. 1962, 84, 679

Veitch, N. C., Smith, A. T. Adv. Inorg. Chem. 2000, 51, 107

Versluis, L., Ziegler, T., Fan, L. Inorg. Chem. 1990, 29, 4530

Vosko, H., Wilk L., Nusair M., Can. J. Phys., 1980, 58, 1200

Vougioukalakis, G. C., Grubbs, R. H. Chem. Rev. 2010, 110, 1746

Wachters, A. J. H. J. Chem. Phys. 1970, 52, 1033

Wang, L.-P., Wu, Q., van Voorhis, T. Inorg. Chem. 2010, 49, 4543

Watanabe, Y., Iyanagi, T., Oae, S. Tetrahedron Lett. 1980, 21, 3685

Watanabe, Y., Iyanagi, T., Oae, S. Tetrahedron Lett. 1982, 23, 533

Watanabe, Y., Nakajima, H., Ueno, T. Acc. Chem. Res. 2007, 40, 554

Watanabe, Y., Numata, T., Iyanagi, T., Oae, S. Bull. Chem. Soc. Jpn. 1981, 54, 1163

Weinberg, D. R., Gagliardi, C. J., Hull, J. F., Murphy, C. F., Kent, C. A., Westlake, B. C.,

Paul, A., Ess, D. H., McCafferty, D. G., Meyer, T. J. Chem. Rev. 2012, 112, 4016

Weinhold, F., Carpenter, J. E. The Natural Bond Orbital Lewis Structure Concept for

Molecules, Radicals, and Radical Ions. In The Structure of Small Molecules and Ions,

Naaman, R., Vager, Z., Eds., Plenum Press, New York, 1988, pp 227

Page 286: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

286

Weiss, R., Fischer, J., Bulach, V., Schünemann, V., Gerdan, M., Trautwein, A. X.,

Shelnutt, J. A., Gros, C. P., Tabard, A., Guilard, R. Inorg. Chim. Acta. 2002, 337, 223

Weiss, R., Gold, A., Terner, J. Chem. Rev. 2006, 106, 2550

Wen, X., Wang, J.-S., Kivistö, K. T., Neuvonen, P. J., Backman, J. T. Br. J. Clin.

Pharmacol. 2001, 52, 547

Weymuth, T., Couzijn, E. P. A., Chen, P., Reiher, M. J. Chem. Theory Comput. 2014, 10,

3092

White, R. E., Groves, J. T. McClusky, G. A. Acta Biol. Med. Germ. 1979, 38, 475

Widenhoefer, R. A., Buchwald, S. L. J. Am. Chem. Soc. 1983, 120, 6504

Wilkins, P. C., Wilkins, R. G. Coord. Chem. Rev. 1987, 79, 195

Wong, S. D., Bell III, C. B., Liu, L. V., Kwak, Y., England, J., Alp, E. E., Zhao, J., Que

Jr, L., Solomon, E. I. Angew. Chem. Int. Ed. 2011, 50, 3215

Wu, A., Mayer, J. M. J. Am. Chem. Soc. 2008, 130, 14745

Yamada, M., Hagiwara, H., Torigoe, H., Matsumoto, N., Kojima, M., Dahan, F.,

Tuchagues, J.-P., Re, N., Iijima, S. Chem. Eur. J. 2006, 12, 4536

Ye, S., Neese, F. Proc. Natl. Acad. Sci. USA 2011, 108. 1228

Ye, S., Wu, X., Wei, L., Tang, D., Sun, P., Bartlam, M., Rao, Z. J. Biol. Chem. 2007, 282,

3391

Yeong, Y. J., Kang, Y., Han, A. R., Lee, Y. M., Kotani, H., Fukuzumi, S., Nam, W.

Angew. Chem., Int. Ed. 2008, 47, 7321

Page 287: Electronic and catalytic properties of iron porphyrin ...Electronic and catalytic properties of iron porphyrin complexes: Trends and reaction mechanisms. A thesis submitted to the

287

Yoon, J., Wilson, S. A., Jang, Y. K., Seo, M. S., Nehru, K., Hedman, B., Hodgson, K. O.,

Bill, E., Solomon, E. I., Nam, W. Angew. Chem. Int. Ed. 2009, 48, 1257

Yoshioka, S., Takahashi, S., Ishimori, K., Morishima, I. J. Inorg. Biochem. 2000, 81, 141

Yoshizawa, K., Shiota, Y., Yamabe, T. J. Am. Chem. Soc. 1983, 120, 564

Yu, F., Cangelosi, V. M., Zastrow, M. L., Tegoni, M., Plegaria, J. S., Tebo, A. G.,

Mocny, C. S., Ruckthong, L., Qayyum, H., Pecoraro, V. L. Chem. Rev. 2014, 114,3495

Yu, J., Hunt, F. Proc. Natl. Acad. Sci. USA 2009, 106, 14315

Yuan, X., Wang, Q., Horner, J. H., Sheng, X., Newcomb, M. Biochemistry 2009, 48,

9140

Zhao, Y., Truhlar, D. G. Acc. Chem. Res. 2008, 41, 157

Zhao, Y., Truhlar, D. G. Theor. Chem. Acc. 2008, 120, 215