Wednesday 25 December 2013

Why sugar is sweet ?


The human tongue responds to a range of different substances, registering them as various tastes. Evolution programmed our gustatory sense to find nutritious things tasty, and un-nutritious things un-tasty — for the most part. Human beings are flexible about what they eat relative to many animals, hence our omnivore status, but there are many types of organic material that we are incapable of digesting, and hence perceive as unpleasant. Sugar is highly digestible and offers a very condensed source of calories, so to us it tastes good — and has a distinct flavor that we label sweet. All mammals, except cats, can taste and enjoy this substance and are more inclined to eat poor-tasting food if it contains some.
Scientists now know that the tongue is covered with tiny clusters of chemical sensors called taste buds. It is the geometric shape of incoming molecules that determines how they taste. Some foods have multiple molecules that all contribute to their overall taste sensation. There are several types of sugar that exist, but by far the most frequently consumed by humans is a molecule called sucrose. There are also other molecules, like saccharine, that taste sweet even though they aren't sugar, although the taste sensation is slightly different.


Sunday 8 September 2013

Notes on MOEL diagrams of octahedral,tetrahedral and square planer complexes involving sigma and pi bonding

MOEL diagram for a square planer complex

MOEL diagrams for octahedral,tetragonal and square planer complexes

Reactions of RMgX and RLi with esters



Reactions of RLi and RMgX with Esters
reaction of RLi or RMgX with esters 
Reaction usually in Et2O followed by H3O+ work-up
Reaction type:  Nucleophilic Acyl Substitution then Nucleophilic Addition
Summary:
  • Carboxylic esters, R'CO2R'', react with 2 equivalents of organolithium or Grignard reagents to give tertiary alcohols.
  • The tertiary alcohol contains 2 identical alkyl groups (see R)
  • The reaction proceeds via a ketone intermediate which then reacts with the second equivalent of the organometallic reagent.
  • Since the ketone is more reactive than the ester, the reaction cannot be used as a preparation of ketones.
REACTION OF RLi or RMgX WITH AN ESTER

Step 1:
The nucleophilic C in the organometallic reagent adds to the electrophilic C in the polar carbonyl group of the ester. Electrons from the C=O move to the electronegative O creating an intermediate metal alkoxide complex.
addition of Grignard reagent to an ester
Step 2:
The tetrahedral intermediate collapses and displaces the alcohol portion of the ester as a leaving group, this produces a ketone as an intermediate.
Step 3:
The nucleophilic C in the organometallic reagent adds to the electrophilic C in the polar carbonyl group of the ketone. Electrons from the C=O move to the electronegative O creating an intermediate metal alkoxide complex.
Step 4:
This is the  work-up step, a simple acid/base reaction. Protonation of the alkoxide oxygen creates the alcohol product from the intermediate complex.


Oxymercuration-demercuration of alkenes



Oxymercuration-Demercuration of Alkenes
alkene alkoxymercuration-demercuration to give an alcohol
Reaction type:  Electrophilic Addition
Summary
  • Overall transformation C=C to H-C-C-OH
  • This is an alternative method for hydrating alkenes to give alcohols
  • Typical reagents are mercury acetate, Hg(OAc)2 in aqueous THF
  • Unfortunately, mercury compounds are generally quite toxic
  • Regioselectivity predicted by Markovnikov's rule (most highly substituted alcohol)
  • The reaction is not stereoselective
  • Reaction proceeds via the formation of a cyclic mercurinium ion (compare with bromination of alkenes)
a mercurinium ion
  • The mercurinium ion is opened by the attack of water to complete the oxymercuration.
  • When the water attacks, it does so at the more highly substituted carbon.
  • Demercuration is effected by a reduction using sodium borohydride, NaBH4
  • If the reaction is carried out in the presence of an alcohol rather than water, then ethers are obtained via an alkoxymercuration :
alkoxymercuration-demercuration to give an ether
  • The only difference here is a change in the nucleophile from H2O to ROH

MECHANISM FOR REACTION OF ALKENES WITH Hg(OAc)2 / H2O
Step 1:
The C=C Ï€ electrons act as the nucleophile with the electrophilic Hg and loss of an acetate ion as a leaving group, forming the mercurinium ion.
oxymercuration / demercuration of C=C
Step 2:
Water functions as a nucleophile and attacks one of the carbons substituted with mercury resulting in cleavage of the C-Hg bond.
Step 3:
The acetate ion functions as a base deprotonating the oxonium ion to give the alcohol. This completes the oxymercurationpart of the reaction.
Step 4:(mechanism not shown)
The hydride reduces the Hg off, creating aC-H bond while breaking the C-Hg bond. This is the demercuration part of the process.

Simmons Smith reaction,formation of cycloalkanes


Synthesis of Cyclopropanes using RZnX (The Simmons-Smith reaction)
Cyclopropanation using the Simmons-Smith reaction
Reaction type:  1.  Oxidation-Reduction,  2.   Addition
Summary
  • This is the most important reaction involving an organozinc reagent.
  • Also known as the Simmons-Smith reaction
  • The iodomethyl zinc iodide is usually prepared using Zn activated with Cu.
  • The iodomethyl zinc iodide reacts with an alkene to give a cyclopropane.
  • The reaction is stereospecific with respect to to the alkene (mechanism is concerted).

  • Substituents that are trans in the alkene are trans in the cyclopropane etc.
Simmons-Smith reaction is stereospecific
Related reactions
 
MECHANISM OF THE SIMMONS-SMITH REACTION
Step 1:
A concerted reaction : both new C-C are formed simultaneously. Best viewed as the nucleophilic C=C causing loss of the iodide leaving group and the electrons from the nucleophilic C-Zn bond being used to form the other C-C bond.
cycloaddition mechanism of the Simmons-Smith
 

Reactions of RMgX and RLi with aldehydes and ketones



Reactions of RLi and RMgX with Aldehydes and Ketones

reaction of RLi or RMgX with aldehydes and ketones 
Reactions usually in Et2O or THF followed by H3O+ work-ups
Reaction type:  Nucleophilic Addition
Summary
  • Organolithium or Grignard reagents react with the carbonyl group, C=O, in aldehydes or ketones to give alcohols.
  • The substituents on the carbonyl dictate the nature of the product alcohol.
  • Addition to methanal (formaldehyde) gives primary alcohols.
  • Addition to other aldehydes gives secondary alcohols.
  • Addition to ketones gives tertiary alcohols.
  • The acidic work-up converts an intermediate metal alkoxide salt into the desired alcohol via a simple acid base reaction.
NUCLEOPHILIC ADDITION OF RLi or RMgX TO AN ALDEHYDE
Step 1:
The nucleophilic C in the organometallic reagent adds to the electrophilic C in the polar carbonyl group, electrons from theC=O move to the electronegative O creating an intermediate metal alkoxide complex.
addition of Grignard reagent to an aldehyde
Step 2:
This is the  work-up step, a simple acid/base reaction. Protonation of the alkoxide oxygen creates the alcohol product from the intermediate complex.

Organocatalysis in synthesis


Organocatalysis

Organocatalysis uses small organic molecules predominantly composed of C, H, O, N, S and P to accelerate chemical reactions. The advantages of organocatalysts include their lack of sensitivity to moisture and oxygen, their ready availability, low cost, and low toxicity, which confers a huge direct benefit in the production of pharmaceutical intermediates when compared with (transition) metal catalysts.
In the example of the Knoevenagel Condensation, it is believed that piperidine forms a reactive iminium ion intermediate with the carbonyl compound:
Another organocatalyst is DMAP, which acts as an acyl transfer agent:
Thiazolium salts are versatile umpolung reagents (acyl anion equivalents), for example finding application in the Stetter Reaction:
All of these organocatalysts are able to form temporary covalent bonds. Other catalysts can form H-bonds, or engage in pi-stacking and ion pair interactions (phase transfer catalysts). Catalysts may be specially designed for a specific task - for example, facilitating enantioselective conversions.
An early example of an enantioselective Stetter Reaction is shown below: :
D. Enders, K. Breuer, J. Runsink, Helv. Chim. Acta1996, 79, 1899-1902.

model explaining the facial selectivity
Enantioselective Michael Addition using phase transfer catalysis:
T. Ooi, D. Ohara, K. Fukumoto, K. Maruoka, Org. Lett.200573195-3197.

The first enantioselective organocatalytic reactions had already been described at the beginning of the 20th century, and some astonishing, selective reactions such as the proline-catalyzed synthesis of optically active steroid partial structures by Hajos, Parrish, Eder, Sauer and Wiechert had been reported in 1971 (Z. G. Hajos, D. R. Parrish, J. Org. Chem. 197439, 1615; U. Eder, G. Sauer, R. Wiechert, Angew. Chem. Int. Ed. 197110, 496, DOI). However, the transition metal-based catalysts developed more recently have drawn the lion’s share of attention.

Hajos-Parrish-Eder-Sauer-Wiechert reaction (example)

The first publications from the groups of MacMillan, List, Denmark, and Jacobson paved the way in the year 1990. These reports introduced highly enantioselective transformations that rivaled the metal-catalyzed reactions in both yields and selectivity. Once this foundation was laid, mounting interest in organocatalysis was reflected in a rapid increase in publications on this topic from a growing number of research groups.
Proline-derived compounds have proven themselves to be real workhorse organocatalysts. They have been used in a variety of carbonyl compound transformations, where the catalysis is believed to involve the iminium form. These catalysts are cheap and readily accessible:
A. J. A. Cobb, D. M. Shaw, D. A. Longbottom, J. B. Gold, S. V. Ley, Org. Biomol. Chem.20053, 84-96.

Y. Hayashi, T. Sumiya, J. Takahashi, H. Gotoh, T. Urushima, M. Shoji, Angew. Chem. Int. Ed.200645, 958-961.

Kumaragurubaran, K. Juhl, W. Zhuang, A. Gogevig, K. A. Jorgensen, J. Am. Chem. Soc.20021246254-6255.

A general picture of recent developments: V. D. B. Bonifacio, Proline Derivatives in Organic SynthesisOrg. Chem. Highlights2007, March 25.

Books on Organocatalysis
Albrecht Berkessel, Harald Gröger
Hardcover, 440 Pages
First Edition, 2005
ISBN: 3-527-30517-3 - Wiley-VCH


Recent Literature
An efficient one-pot procedure allows the preparation of substituted quinolines from activated acetylenes and o-tosylamidocarbonyl compounds under base-catalyzed, mild conditions. The generation of a β-phosphonium enoate α-vinyl anion in situ is followed by Michael addition of the deprotonated tosylamides and subsequent rapid aldol cyclization. Detosylation of the dihydroquinoline intermediates occurred readily in the presence of aqueous HCl.
S. Khong, O. Kwon, J. Org. Chem.2012778257-8267.

A practical and highly enantioselective Michael addition of malonates to enones to yield 1,5-ketoesters with good yields and excellent enantioselectivities is catalyzed by a simple and readily available bifunctional primary amine-thiourea derived from 1,2-diaminocyclohexane. The addition of weak acids and elevated temperature improved the efficiency of the reaction. This approach is applicable in multigram scale synthesis.
K. DudziÅ„ski, A. M. Pakulska, P. Kwiatkowski, Org. Lett.2012144222-4225.

N-hydroxyphthalimide (NHPI) catalyzes a metal-free, aerobic oxidative cleavage of olefins. This methodology avoids the use of toxic metals or overstoichiometric amounts of traditional oxidants, showing good economical and environmental advantages. Based on the experimental observations, a plausible mechanism is proposed.
R. Lin, F. Chen, N. Jiao, Org. Lett.2012144158-4161.

A one-pot conversion of aldehydes to esters interfaces N-heterocyclic carbene-based organocatalysis with electro-organic synthesis to achieve direct oxidation of catalytically generated electroactive intermediates. A broad range of aldehyde and alcohol substrates has been converted. Furthermore, the anodic oxidation reactions are very clean, producing only H2 gas as a result of cathodic reduction.
E. E. Finney, K. A. Ogawa, A. J. Boydston, J. Am. Chem. Soc.2012134, 12374-12377.

Aryl iodides are efficient catalysts in an organocatalytic syn diacetoxylation of alkenes. A broad range of substrates, including electron-rich as well as electron-deficient alkenes,  furnish the desired products in very good yields with high diastereoselectivity.
W. Zhong, S. Liu, J. Yang, X. Meng, Z. Li, Org. Lett.2012143336-3339.

Confined chiral Brønsted acids catalyze asymmetric oxidations of a broad range of sulfides to sulfoxides with hydrogen peroxide. The wide generality and high enantioselectivity of the developed method is comparable even to the best metal-based systems.
S. Liao, I. ÄŒorić, Q. Wang, B. List, J. Am. Chem. Soc.2012134, 10765-10768.

A direct reductive amination of ketones using the Hantzsch ester in the presence of S-benzyl isothiouronium chloride as a recoverable organocatalyst converts a wide range of ketones as well as aryl amines to the expected products in good yields.
Q. P. B. Nguyen, T. H. Kim, Synthesis2012, 1977-1982.

An organocatalytic Dakin oxidation of electron-rich arylaldehydes to phenols can be performed under mild, basic conditions using flavin catalysts. Catechols are readily prepared and the oxidation of 2-hydroxyacetophenone was achieved.
S. C. M. S. Hoassain, F. W. Foss, Jr, Org. Lett.201214, 2806-2809.

The in situ generation of α-amino aldehydes followed by reaction with dimethyloxosulfonium methylide under Corey-Chaykovsky reaction conditions gives 4-hydroxypyrazolidine derivatives in high yields with excellent enantio- and diastereoselectivities. This organocatalytic sequential method enables an efficient synthesis of anti-1,2-aminoalcohols.
B. S. Kumar, V. Venkataramasubramanian, A. Sudalai, Org. Lett.201214, 2468-2471.

Ozonolysis in the presence of pyridine directly generates ketones or aldehydes through a process that neither consumes pyridine nor generates any detectable peroxides. The reaction is hypothesized to involve nucleophile-promoted fragmentation of carbonyl oxides via formation of zwitterionic peroxyacetals.
R. Willand-Charnley, T. J. Fisher, B. M. Johnson, P. H. Dussault, Org. Lett.201214, 2242-2245.

A bifunctional organocatalyst efficiently catalyzed not only enantioselective conjugate addition of aromatic ketones to nitroolefins in good yields with excellent enantioselectivities but also enantioselective conjugate addition of acetone to nitroolefins in excellent yields with high enantioselectivities.
Z.-W. Sun, F.-Z. Peng, Z.-Q. Li, L.-W. Zhou, S.-X. Zhang, X. Li, Z.-H. Shao, J. Org. Chem.2012774103-4110.

Organocatalytic stereospecific dibromination of various functionalized alkenes was achieved using a simple thiourea catalyst and 1,3-dibromo 5,5-dimethylhydantoin as a stable, inexpensive halogen source at room temperature. The procedure was extended to alkynes and aromatic rings and to dichlorination reactions by using the 1,3-dichloro hydantoin derivative.
G. Hernández-Torres, B. Tan, C. F. Barbas III, Org. Lett.201214, 1858-1861.

A nitroxyl-radical-catalyzed oxidation using diisopropyl azodicarboxylate (DIAD) allows the conversion of various primary and secondary alcohols to their corresponding aldehydes and ketones without overoxidation to carboxylic acids. 1,2-Diols are oxidized to hydroxyl ketones or diketones depending on the amount of DIAD used.
M. Hayashi, M. Shibuay, Y. Iwabuchi, J. Org. Chem.2012773005-3009.

An enantioselective synthesis of γ-nitroesters by a one-pot asymmetric Michael addition/oxidative esterification of α,β-unsaturated aldehydes is based on an enantioselective organocatalytic nitroalkane addition followed by an N-bromosuccinimide-based oxidation. The γ-nitroesters are obtained in good yields and enantioselectivities, and the method provides an attractive entry to optically active γ-aminoesters, 2-piperidones, and 2-pyrrolidones.
K. L. Jensen, P. H. Poulsen, B. S. Donslund, F. Morana, K. A. Jørgensen, Org. Lett.201214, 1516-1519.

A simple chiral primary amine catalyses a highly efficient reaction for the synthesis of both Wieland-Miescher ketone and Hajos-Parrish ketone as well as their analogues in high enantioselectivity and excellent yields. This procedure represents one of the most efficient methods for the synthesis of these versatile chiral building blocks even in gram scale with 1 mol% catalyst loading.
P. Zhou, L. Zhang, S. Luo, J.-P. Cheng, J. Org. Chem.201277, 2526-2530.

The of silica-coated magnetic nanoparticles allowed the construction of magnetically recoverable organic hydride compounds. Magnetic nanoparticle-supported BNAH (1-benzyl-1,4-dihydronicotinamide) showed efficient activity in the catalytic reduction of α,β-epoxy ketones. After reaction, the catalyst can be separated by simple magnetic separation and can be reused.
H.-J. Xu, X. Wan, Y.-Y. Shen, S. Xu, Y.-S. Feng, Org. Lett.201214, 1210-1213.

Activation of diphenylsilane in the presence of a catalytic amount of an N-heterocyclic carbene (NHC) enables hydrosilylation of carbonyl derivatives under mild conditions. Presumably, a hypervalent silicon intermediate featuring strong Lewis acid character allows dual activation of both the carbonyl moiety and the hydride at the silicon center. Some interesting selectivities have been encountered.
Q. Zhao, D. P. Curran, M. Malacria, L. Fensterbank, J.-P. Goddard, E. Lacôte, Synlett2012, 433-437.

A bifunctional squaramide catalyzes a sulfa-Michael/aldol cascade reaction between 1,4-dithiane-2,5-diol and chalcones with a low catalyst loading to yield trisubstituted tetrahydrothiophenes with three contiguous stereogenic centers in a highly stereocontrolled manner.
J.-B. Ling, Y. Su, H.-L. Zhu, G.-Y. Wang, P.-F. Xu, Org. Lett.201214, 1090-1093.

A phosphinite derivative that can be easily prepared in two steps from commercially available aminoindanol is an effective catalyst for enantioselective acylation of diols. For the asymmetric desymmetrization of meso-1,2-diols, the corresponding monoester was obtained in high enantioselectivity.
H. Aida, K. Mori, Y. Yamaguchi, S. Mizuta, T. Moriyama, I. Yamamoto, T. Fujimoto, Org. Lett.201214, 812-815.

Commercially available and very inexpensive benzoic acids catalyze an efficient and simple isomerization of readily prepared allylic alcohols to yield cyclic products, unusual enyne, and dienols. The catalysts can be tuned for reactivity and substrate sensitivity.
J. A. McCubbin, S. Voth, O. V. Krokhin, J. Org. Chem.2011768537-8542.

Cinchona-alkaloid-thiourea-based bifunctional organocatalysts enable a straightforward asymmetric cycloetherification of ε-hydroxy-α,β-unsaturated ketones for the synthesis of tetrahydrofuran rings. This catalytic process represents a highly practical cycloetherification method that provides excellent enantioselectivities, even with low catalyst loadings at ambient temperature.
K. Asano, S. Matsubara, J. Am. Chem. Soc.2011133, 16711-16713.



Tuesday 27 August 2013

Reaction kinetics and their applications to the study of organic chemistry

Stability,chelation and chelate effect in coordination compounds


Stability, Chelation and the Chelate Effect

A metal ion in solution does not exist in isolation, but in combination with ligands (such as solvent molecules or simple ions) or chelating groups, giving rise to complex ions or coordination compounds.
These complexes contain a central atom or ion, often a transition metal, and a cluster of ions or neutral molecules surrounding it. Many complexes are relatively unreactive species remaining unchanged throughout a sequence of chemical or physical operations and can often be isolated as stable solids or liquid compounds.
Other complexes have a much more transient existence and may exist only in solution or be highly reactive and easily converted to other species.
All metals form complexes, although the extent of formation and nature of these depend very largely on the electronic structure of the metal.
The concept of a metal complex originated in the work of Alfred Werner, who in 1913 was awarded the first Nobel Prize in Inorganic chemistry. A description of his life and the influence his work played in the development of coordination chemistry is given by G.B. Kauffman in "Inorganic Coordination Compounds", Heyden & Son Ltd, 1981.
Complexes may be non-ionic (neutral) or cationic or anionic, depending on the charges carried by the central metal ion and the coordinated groups. The total number of points of attachment to the central element is termed the coordination number and this can vary from 2 to greater than 12, but is usually 6.
The term ligand (ligare [Latin], to bind) was first used by Alfred Stock in 1916 in relation to silicon chemistry. The first use of the term in a British journal was by H. Irving and R.J.P. Williams in Nature, 1948, 162, 746 in their paper describing what is now called the Irving-Williams series.
For a fascinating review of the origin and dissemination of the term 'ligand' in chemistry see: W.H. Brock, K.A Jensen, C.K. Jorgensen and G.B. Kauffman, Polyhedron, 2, 1983, 1-7.
Ligands can be further characterised as monodentate, bidentate, tridentate etc. where the concept of teeth (dent) is introduced, hence the idea of bite angle etc.
The term chelate was first applied in 1920 by Sir Gilbert T. Morgan and H.D.K. Drew [J. Chem. Soc., 1920, 117, 1456], who stated: 
"The adjective chelate, derived from the great claw or chela (from the Greek χηλη) of the lobster or other crustaceans, is suggested for the caliperlike groups which function as two associating units and fasten to the central atom so as to produce heterocyclic rings." 

lobster claw 
Metal complexation is of widespread interest. It is studied not only by inorganic chemists, but by physical and organic chemists and by biochemists, pharmacologists, molecular biologists and environmentalists. 

Thermodynamic Stability

In the laboratory course, it will have been pointed out that the "stability of a complex in solution" refers to the degree of association between the two species involved in the state of equilibrium. Qualitatively, the greater the association, the greater the stability of the compound. The magnitude of the (stability or formation) equilibrium constant for the association, quantitatively expresses the stability. Thus, if we have a reaction of the type: 
M   +   4L    →   ML4
then the larger the stability constant, the higher the proportion of ML4 that exists in the solution. Free metal ions rarely exist in solution so that M, will usually be surrounded by solvent molecules which will compete with the ligand molecules, L, and be successively replaced by them. For simplicity, we generally ignore these solvent molecules and write four stability constants as follows: 
l.       M + L → ML            K1 = [ML] / [M] [L] 
2.      ML + L → ML2        K2 = [ML2] / [ML] [L] 
3.      ML2 + L → ML3      K3 = [ML3] / [ML2] [L] 
4.      ML3 + L → ML4      K4 = [ML4] / [ML3] [L] 
where K1, K2 etc. are referred to as "stepwise stability constants". 
Alternatively, we can write the "Overall Stability Constant" thus: 
M   +   4L    →   ML4      β4 = [ML4]/ [M] [L]4
The stepwise and overall stability constants are therefore related as follows: 
β4 =K1.K2.K3.K4 or more generally, 
βn =K1.K2.K3.K4--------------K n
If we take as an example, the steps involved in the formation of the cuprammonium ion, we have the following: 
Cu2+ + NH3 ↔ Cu(NH3)2+ K1 = [Cu(NH3)2+]/[Cu2+] [NH3]
CuNH32+ + NH3 ↔ Cu(NH3)22+ K2 = [Cu(NH3)22+]/[Cu(NH3)2+] [NH3]
etc. where K1, K2 are the stepwise stability constants. 
Also: 
          β4  = [Cu(NH3)42+]/[Cu2+] [NH3]4
The addition of the four ammine groups to copper shows a pattern found for most formation constants, in that the successive stability constants decrease. In this case, the four constants are: 
     logK1 =4.0, logK2 =3.2, logK3 =2.7,  logK4 =2.0 or logβ4 =11.9
A number of texts refer to the instability constant or the dissociation constant of coordination complexes. This value corresponds to the reciprocalof the formation constant, since the reactions referred to are those where fully formed complexes break down to the aqua ion and free ligands. 
This should be compared with the equation for the formation constant given earlier. 
It is usual to represent the metal-binding process by a series of stepwise equilibria which lead to stability constants that may vary numerically from hundreds to enormous values such as 1035 and more. 
That is 100,000,000,000,000,000,000,000,000,000,000,000.0 
For this reason, they are commonly reported as logarithms. 
so log10 (β) = log10 (1035) = 35. 
It is additionally useful to use logarithms, since log(K) is directly proportional to the free energy of the reaction.

ΔG° = -RTln(β) 
ΔG° = -2.303 RTlog10(β) 
ΔG° = ΔH° - TΔS° 
For a problem relating to metal complex formation and calculations of free metal ions concentrations, try your hand at CALCULATION # ONE. Other problems can be found in the Tutorial paper for this course. 

The Chelate Effect

The chelate effect can be seen by comparing the reaction of a chelating ligand and a metal ion with the corresponding reaction involving comparable monodentate ligands. For example, comparison of the binding of 2,2'-bipyridine with pyridine or 1,2-diaminoethane (ethylenediamine=en) with ammonia.
It has been known for many years that a comparison of this type always shows that the complex resulting from coordination with the chelating ligand is much more thermodynamically stable. This can be seen by looking at the values for adding two monodentates compared with adding one bidentate, or adding four monodentates compared to two bidentates, or adding six monodentates compared to three bidentates. 

Some tables of thermodynamic data

Reaction of ammonia and 1,2-diaminoethane with Cd2+.
# of ligandsΔG° (kJmol-1)ΔH° (kJmol-1)ΔS° (JK-1mol-1)log β
2 NH3(1 en)-28.24 (-33.30)-29.79 (-29.41)-5.19 (+13.05)4.95 (5.84)
4 NH3(2 en)-42.51 (-60.67)-53.14 (-56.48)-35.50 (+13.75)7.44 (10.62)

Reaction of pyridine and 2,2'-bipyridine with Ni2+.
# of ligandslog βΔG° (kJmol-1)
2 py (1 bipy)3.5 (6.9)-20 (-39)
4 py (2 bipy)5.6 (13.6)-32 (-78)
6 py (3 bipy)9.8 (19.3)-56 (-110)

Reaction of ammonia and 1,2-diaminoethane with Ni2+.
# of ligandslog βΔG° (kJmol-1)
1 NH32.8-16
2 NH3 (1 en)5.0 (7.51)-28.5 (-42.8)
3 NH36.6-37.7
4 NH3 (2 en)7.87 (13.86)-44.9 (-79.1)
5 NH38.6-49.1
6 NH3 (3 en)8.61 (18.28)-49.2 (-104.4)

A number of points should be highlighted from this data. 
In the first table, it can be seen that the ΔH° values for the formation steps are almost identical, that is, heat is evolved to about the same extent whether forming a complex involving monodentate ligands or bidentate ligands. 
What is seen to vary significantly is the ΔS° term which changes from negative (unfavourable) to positive (favourable). Note as well that there is a dramatic increase in the size of the ΔS° term for adding two compared to adding four monodentate ligands. (-5 to -35 JK-1mol-1). 
What does this imply, if we consider ΔS° to give a measure of disorder?
In the case of complex formation of Ni2+ with ammonia or 1,2-diaminoethane, by rewriting the equilibria, the following equations are produced.
Ni reactions

Using the equilibrium constant for the reaction (3 above) where the three bidentates replace the six monodentates, we find that at a temperature of 25C:
ΔG° = -2.303 RT log10 (K) 
= -2.303 R T (18.28 - 8.61) 
= -54 kJ mol-1 
Based on measurements made over a range of temperatures, it is possible to break down the ΔG° term into the enthalpy and entropy components. ΔG° = ΔH° - TΔS° 
The result is that:
ΔH° = -29 kJ mol-1 
- TΔS° = -25 kJ mol-1 
and at 25C (298K) 
ΔS° = +88 J K-1 mol-1 

Note that for many years, these numbers have been incorrectly recorded in textbooks. 
For example, the third edition of "Basic Inorganic Chemistry" by F.A. Cotton, G. Wilkinson and P.L. Gaus, John Wiley & Sons, Inc, 1995, on page 186 gives the values as:
ΔG° = -67 kJ mol-1 
ΔH° = -12 kJ mol-1 
- TΔS° = -55 kJ mol-1 
The conclusion they drew from these incorrect numbers was that the chelate effect was essentially an entropy effect, since the TΔS° contribution was nearly 5 times bigger than ΔH°.

In fact, the breakdown of the ΔG° into ΔH° and TΔS° shows that the two terms are nearly equal (-29 cf. -25 kJ mol-1) with the ΔH° term a little bigger! The entropy term found is still much larger than for reactions involving a non-chelating ligand substitution at a metal ion. 
How can we explain this enhanced contribution from entropy? One explanation is to count the number of species on the left and right hand side of the equation above. 
It will be seen that on the left-hand-side there are 4 species, whereas on the right-hand-side there are 7 species, that is a net gain of 3 species occurs as the reaction proceeds. This can account for the increase in entropy since it represents an increase in the disorder of the system. 
An alternative view comes from trying to understand how the reactions might proceed. To form a complex with 6 monodentates requires 6 separate favourable collisions between the metal ion and the ligand molecules. To form the tris-bidentate metal complex requires an initial collision for the first ligand to attach by one arm but remember that the other arm is always going to be nearby and only requires a rotation of the other end to enable the ligand to form the chelate ring. 
If you consider dissociation steps, then when a monodentate group is displaced, it is lost into the bulk of the solution. On the other hand, if one end of a bidentate group is displaced the other arm is still attached and it is only a matter of the arm rotating around and it can be reattached again. 
Both sets of conditions favour the formation of the complex with bidentate groups rather than monodentate groups.

Advanced organic chemistry notes

Factors governing polarization and polarisability in Fazans Rule



FACTORS GOVERNING POLARIZATION AND POLARISABILITY (FAJAN’S RULE)

Cation Size: Smaller is the cation more is the value of charge density (Φ) and hence more its polarising power. As a result more covalent character will develop. Let us take the example of the chlorides of the alkaline earth metals. As we go down from Be to Ba the cation size increases and the value of Î¦ decreases which indicates that BaClis less covalent i.e. more ionic. This is well reflected in their melting points. Melting points of BeCl2 = 405°C and BaCl2 = 960°C.
Cationic Charge: More is the charge on the cation, the higher is the value of Î¦ and higher is the polarising power. This can be well illustrated by the example already given, NaBr and AlBr3. Here the charge on Na is +1 while that on Al in +3, hence polarising power of Al is higher which in turn means a higher degree of covalency resulting in a lowering of melting point of AlBr3 as compared to NaBr.  
Noble Gas vs Pseudo Noble Gas Cation:A Pseudo noble gas cation consists of a noble gas core surrounded by electron cloud due to filled d-subshell. Since d-electrons provide inadequate shielding from the nuclei charge due to relatively less penetration of orbitals into the inner electron core, the effective nuclear charge (ENC) is relatively larger than that of a noble gas cation of the same period. NaCl has got a melting point of 800°C while CuCl has got melting point of 425°C. The configuration of Cu+ = [Ar] 3d10 while that of Na+ = [Ne]. Due to presence of d electrons ENC is more and therefore Cl is more polarised in CuCl leading to a higher degree of covalency and lower melting point.
Anion Size:Larger is the anion, more is the polarisability and hence more covalent character is expected. An e.g. of this is CaF2 and CaI2, the former has melting point of 1400°C and latter has 575°C. The larger size of I ion compared to F causes more polarization of the molecule leading to a lowering of covalency and increasing in melting point.
Anionic Charge:Larger is the anionic charge, the more is the polarisability. A well illustrated example is the much higher degree of covalency in magnesium nitride (3Mg++ 2N3–) compared to magnesium fluoride (Mg++ 2F). This is due to higher charge of nitride compare to fluoride. These five factors are collectively known as Fajan’s Rule.
Example:
The melting point of KCl is higher than that of AgCl though the crystal radii of Ag+and K+ ions are almost same.
Solution:

Now whenever any comparison is asked about the melting point of the compounds which are fully ionic from the electron transfer concept it means that the compound having lower melting point has got lesser amount of ionic character than the other one. To analyse such a question first find out the difference between the 2 given compounds. Here in both the compounds the anion is the same. So the deciding factor would be the cation. Now if the cation is different, then the answer should be from the variation of the cation. Now in the above example, the difference of the cation is their electronic configuration. K+ = [Ar]; Ag+ = [Kr] 4d10. This is now a comparison between a noble gas core and pseudo noble gas core, the analysis of which we have already done. So try to finish off this answer.
Example: AlF3 is ionic while AlCl3 is covalent.
Solution: Since F is smaller in size, its polarisability is less and therefore it is having more ionic character. Whereas Cl being larger in size is having more polarisability and hence more covalent character.
Example:
Which compound from each of the following pairs is more covalent and why? 
(a) CuO or CuS                 (b) AgCl or AgI
(c) PbCl2 or PbCl4              (d) BeCl2 or MgCl2
Solution:
(a) CuS                           (b) AgI

(c) PbCl4                         (d) BeCl­2

Wednesday 21 August 2013

Cytochrome P450


Figure
Figure 1 The fold of cytochrome P450s is highly conserved and shown in a ribbon representation (distal face). Substrate recognition sequence (SRS) regions are shown in black and labeled. α-Helixes mentioned in the text are labeled with capital letters.



Figure
P450 catalyti cycle
Figure
Figure 3 Crystal structure of the oxy complex of wild-type CYP101 (pdb code 1dz8) that catalyzes the 5-exo hydroxylation of camphor. Oxygen binding induces a rotation of the amide of Asp251, which ensues binding of two new water molecules, Wat901 and Wat902.
  • P450 catalytic mech focuses on state of heme iron and O
  • Steps:
  1. Oxygen binds reduced heme iron. Formation of oxygenated heme Fe2+-OO or Fe3+-OO-
  2. One-electron reduction of complex to a ferric peroxo state Fe3+-OO2-, which is earily protonated to form hydroperoxo Fe3+-OOH- 
  3. 2nd protonation of latter Fe3+-OOH- complex at distal oxygen atom to form unstable transient Fe-OOH2. Followed by heterolytic scission of O-O bond. Water mol released
  4. Reactions of remaining higher valent pophyrin metal-oxo complex, a ferryl-oxo Ã°-cation porphyrin radical and referred to as “Compound I"
  • intmts formed in reactions 2 to 4 have common features in all cytochrome P450s
  • similar iron-oxygen states are thought to be important in non-heme oxygen activation
2. Active-Site Structure of P450 Enzymes
  • P450 share common overall fold and topology
  • Conserved P450 strucural core formed by a four-helix bundle composed of 3 parallel helices D,L, and I and 1 antiparallel helix E.51
  • Prosthetic heme group is confined between distal I helix and proximal L helix and boung to adj Cysheme-ligand loop containing P450 signature aa seq. 
  • Absolutely conserved Cys is proximal or 5th ligand to heme iron
  • This sulfur ligand is thiolate 52
  • Prximal Cys forms 2 H bonds with neighbouring backbone amides
  • Further interaction with a side chain in some P450s, with Gln in CYP 152A1 or Trp in nitric oxide synthase
  • Mutations of these affect reduction potential or catalytic activity and stability of bond between heme iron and its 5th or 6th ligand in NOS
  • Long I helix forms a wall of heme pocket and contains signature aa seq (A/G)-Gx(E/D)T which is centred at a kink in middle of helix
  • Highly conserved thr preceded by an acidic residue is positioned in active site
  • may be involved in catalysis
  • SRS predetermine P450 substrate spec
  • Point mutations in SRSs affect substrate spec
  • SRSs are flesible protein regions
  • move on substrate binding in induced fit mech to favour substrate binding

3. Enzymatic Reaction Cycle of Cytochrome P450
  • Substrate binds to resting state of low spin (LS) ferric enzyme 
  • perturbs water coordinated as 6th ligand of hene iron
  • change spin state to high spin (HS) substrate-bound complex
  • HS F3+ has more positive reduction potential
  • In CYO101 much easier reduced to ferrous state
  • In other systems, spin shift is not obligatory part of cycle
  • Oxygen binding ---> oxy-450 complex
  • last relatively stable intmt in cycle
  • Complex reduced
  • formation of peroxo-ferric intmt
  • formation of its protonated form, hydroperoxo-ferric intmt
  • 2nd protonation of distal oxygen atom with heterolysis of O=O bond and formation of Compound I and water
  • Oxygenation of substrate to form product complex
  • P450 reaction cycle has at least 3 branch points
  • multiple side reactions can occur under physiological conditions
  • 3 major abortive reaction are
  1. autoxidation of oxy-ferrous enzyme with concomitant production of a superoxide anion and return of enz to resting state
  2. peroxide shunt. Coordinted peroxide or hydroperoxide anion dissoc from iron forming H2O2,completing unproductive 2-electron reduction of oxygen
  3. oxidase uncoupling. Ferry=oxo intmt is oxidised to water instead of oxygenation of substrate. Results in 4-electron reduction of O2 mol. Net formation of 2 mols of water.
  • These procseses are uncoupling
3.1. Substrate Binding
  • In general substrates for cyt P450 metabolism are hydrophobic
  • Substrate binding triggers change of spin state from LS to HS in heme iron
  • induce change in reduction potential from ca. -300 to ca. 100 mV more opsitive
  • In resting state, or in quilibrium with aerobic media, cyt P450 appear in ferric Fe3+ form
  • due to low reduction potential of Fe3+/Fe2+ couple
  • -400 to -170mV
  • Low redox potential is maintained by prsence of negatively charged proximal thiolate ligand
  • same negative shift of reduction potential can be induced in other heme proteins eg myoglobin 85 by replacing proximal His by Cys
  • LS-HS thermodynamis equilibria for Fe3+ and Fe2+ states of enz are coupled with 6th ligand binding equilibrium
  • Experimentally measured mitpoint potential of heme enz depends on substrate's ability to change heme ligand binding equilibria in ferric and ferrous enz
  • In purifies _450 systems, in absence of other stronger ligands, water coordination at 6th distal position can stabilise LS state of ferric iron
  • Reduced ferrous cyt P450s are predominantly in HS 5-coordinated state
  • water is much weaker ligand for Fe2+ heme
  • Difference in ligation state of ferric and ferrous P450 in absence of substrates causes add. stabiliosation of ferric state and lower midpoint potentials of substrate-free cyt P450s.
  • Substrate binding causes water mol coordinated to Fe 3+ to be usually displaced
  • indicated by shift of spin state of 5-coordinated heme iron to HS
  • Loss of 6th ligand of heme iron thermodynamically destabilises ferric state of cyt P450 with respect to Fe2+ state
  • midpoint potential of heme shifts to positive values
  • Stronger ability of substrate to perturb water ligation of ferric heme, more pronounced resulting positive shift of redox potential
  • Same factors affect reduced enz
  • In presence of CO or O2 which do not bind ferric iron porphyrins but are strong ligands for Fe2+ heme, reduction potential shifts to more positive values compared to those measured under inert atmosphere.
  • Coupling between substrate binding and change in reduction potential is indicated by diff in substrate binding free energy in ferric and ferrous state
  • caused by closed thermodynamic cycle.
  • Temperature and pH and presence of cosolvents may change parameters of ovserved high-spin-low-spin equilibrium 
  • via perturbation of 6th ligand binding 
  • change reduction potential of cyt P450
  • In most P450 systems ultimate reducing agent for catalytic cycle is NADPH, which has midpoint potential of -320 mV
  • reduction potentials of protein's redox partners are roughly in same range
  • cyt P450 should be reduced slowly before substrate binds
  • Prevents unproductive turnover of enz with waste of NADPH and formatio nof toxic superoxides and peroxides which is rendered more likely due to fast autooxidation of thiolate ligated heme proteins
  • Mammalian cytochrome CYP3A4 can bind 2 or 3 substrates mols
  • In this system spin shift caused by cooperative substrate binding can serve as allosteric switch
  • from slow turnowever at low [S] to faster turnover at high [S]
  • Observed cooperativity of product formation may be higher than cooperativity of substrate binding
  • Many of cyt P450s have broad spectrum of substrates
  • Induced fit model
  • Large str rearrangements induced by substrate binding observed by comparing Sray crystal st of cyt P450s 
  • Subsrate recognition sites are flexible and procide substrate access to heme
  • HEme otherwise buried in protein globule
  • In absence of charged and H bonding grps on substrate mols and in the active sites of most P450 enz. such binding mechs stabilise substrate in active centre
  • In many cases diff substrate analogues bind tightly to P450 enz because of poor solubility in water
  • not because of strong interactions at active site
  • Wired substrates bind CYP101 more tightly than natural substrate camphor
  • Camphor is tethered to fluorescent reporter grp by hydrophobic links of diff length
  • Hydrocarbon tether is dehydrated ---> favourable effect
  • Long hydrophobic tail is extended thru binding channel in protein globule up to surface, where fluoresecent grp is partly exposed to solvent
  • Same conform changes observed with substrates linked to a Ru complex
  • These tethered modified substrates do not dusplace water coordinated to heme iron
  • do not shift spin state to high spin, as camphor does
  • Spin shift reg mech is sensitive to str of bound substrate or analogue
  • enz recognises optimcal substrates this way when there are almost no spec functional grps at active cetnre to control substrate binding spec


3.2. Iron Spin Shift and the Heme Redox Potential
  • Some substrates of cyt P450s bind with very high affinity but do not display marked shirt in spin state of ferric he,e
  • A crystal str shows no water at 6th ligand position for wired substrate but contains water for another analogue
  • Redox potential as Fe3+/Fe2+ equilibrium is perturbed by changes in ligation state
  • or changes in ligation strength in course of reduction or oxidation
  • In presence of strong ligands for ferric heme redox potential is lower
  • Strong ligands for ferrous state will increase redox potential of heme enz
  • This concept is valid even without great change in spin state caused by ligand replacement
  • equilibrium constants for spin state equilibrium change for diff ligands
  • Interaction with other proteins may change heme reduction potential in cyt P450
  • In ferric CYP101 studies, spec conform changes of heme, proximal thiolate and key distal pocket residues caused by formation of complex with putidaredoxin
  • Interaction with adrenodoxin induces high spin shift in CYP11A1.
  • Spin state of heme iron in CYP101 in complex with Pdr coupled with spec change in ESR spectrum of reduced Pdr
  • may help function
  • Perturb CO stretch band in Fe-CO  complex in CYP 101
  • promote e donation to heme iron from axial sulphur ligand of Cys357
  • Conform hcange in P450 reductase important in reductase catalysis
  • Shift of heme iron spin state from low to high spin
  • Ligand concentration or temp can change position of thermodynamic equilibrium with coupled microscopic equilibria
  • Redox potential is thermodynamic measure of equilibrium between diff oxidation states
  • does not solely determine rate of heme reduction
  • Kinetics of reduction depends on spin state change
  • reorganisation energy diff involved in changes from a 6-coorfinated to 5-coordinated state
  • If reduction is accompanied by changes in coordination state and spin state
          === activation barrier is higher
          ---- rate of reaction is lower
  • Eg spni shift alone caused 200-fold increase of reduction rate in substrate-bound wildtpye CYP102 compared to substrate-free mutatns F393A and F393H of same protein
  • all 3 enz have similar reduction potentials
  • Catalytic turnover is higher in wt enz
  • kinetic control of later steps in kinetic cycle, mainly 2nd ET to ferrous-exy complex
  • Using resonance raman spectroscopic studies large changes of heme iron reduction potential attributed to conform changes of proponiate and vinyl grps
  • Spec conform change caused by substrate binding and formation of complex with ET partners important in reduction kinetics
  • Tight binding of steriodal P450 with its redox partner adrenodoxin increase rate of P450 reduction and product formation
  • Substrate binding may be rate-limiting step in some P450 systems
          --- low [S]

3.3. Oxygen Complex

  • Oxygen binding reduced P450 gives Fe2+-OO complex or Fe3+-OO- complex
  • Gross str of oxy-P450 similar to analogous xomplezes in Mb and Hb and heme enzymes, HRP, HO
  • In P450 this complex is diamagnetic and EPR silent like Fe2+-OO complex
  • only partial electron density transfer from iron to oxygen
  • Active proton delivery to bound oxygen or peroxide ligand in P450 mech of oxygen activation
  • Autoxidation rate depends strongly on reduction potential in mutants of CYP102
  • X-ray crustal str of Fe2+-OO complex of CYP101 was similar to analogous complexes of other heme enz
  • O is coordinated in bent end-on mode with angle Fe-O-O 142 degrees
  • no steric conflict with bound substrate mol

3.4. Formation of the Peroxo/Hydroperoxo Complex

  • Stability of peroxo state Fe3+-OO(H)- is marginal in heme enz
  • Prsence of strong proximal ligand (His, Cys or Tyr) and aq sol near neutral pH defines str of most of these complexes as end-on and low-spin state
  • Such complexes obtained in reactions with H2O2 with heme enz have low stability
  • convert fast to ferryl-oxo-species
  • Reactions of peroxide dianion with free Fe3+ porphyrins afford high spin Fe3+-OO2- complexes 
          --- with side-on-bound peroxide and iron displaced out of porphyrin plane towards bound ligand
  • To realise this str in heme protein must break proximal ligand bond to iron
  • Presence of strong proximal ligand favours low spin state in 6-coordinate Fe3+-OOH- comlexes
          --- important restirction on chem of oxygen activation
          --- characteristic of heme enz


  • if system fails to do 2nd protonation at distal oxygen site to promote O-O cleavage
          ---> uncoupling reacion
          --- transition from 5b to resting state of enz


3.5. Peroxoferric Intermediates in Heme

  • Enzymes : role of protein transfer
  • In heme enz porpyrin and proximal ligand provide heme with 5 coordination sites
  • Predent side on of peroxide
         --- common in non-heme metalloenz
         --- very rare in model metalloporphyrin complezes
  • Porphyrin donates e for O-O bond cleavage in peroxide ligated to heme iron
  • Porphyrin donates 1 e to peroxide ligand
          ----> heterolytic scission of O=O bond
          ----> cation radical forms on pophyrin ring
  • Oxyegn activation begins when O2 binds as an axial ligand to Fe2+ heme iron or H2O2 binding Fe3+ heme iron
  • Diff between oxidase/ oxygenase pathway an peroxidase/peroxygenase pathway
          --- diff in redox state of oxygen vs peroxide
          --- former pathway needs 2 add reduction steps (1 e reductions by exogenous e donor)
          --- must provide 2 protons delivered to peroxide dianion heme ligand 
  • Reultant equivalent of rearranged H2O2, iron-coordinated peroxo-water (Fe-O-OH2) is precursor to heterolytic O-O bond cleavage to form ferryl-oxo porphyrin complex and H2O product
  • H bond stabilised network of watre mols is important
  • uncoupling of cyt P450 CYP101 mutants with native substrate camphor and wt CYP101 when metabolising other substrates can be conceptualised via operation of distal pocket proton relay system
  • Relay composed of water mols stabilised in active centre of enz and 2 residues conserved in P450 systems
  • Famous acid-alcohol pair is Asp251-Thr252 in CYP101
  • Peroxidase and oxygenase enz use high valent ferryl-oxo pophyrin caton radical as catalyst of oxidative transformation of substrates
  • In P450 catalysis O2 binds reduced heme iron
  • ferrous-oxygen or ferric-superoxide complex accepts one more e fr protein redox partner
  • form peroxo ferric intmt
  • protonated to hydroperoxo-ferric intermediate
  • contrary to P450 and HO in peroxidases natural formation of active intmt involves HH2O2 as O donor
  • H2O2 binds ferrric heme 
  • brings 2 e and 2 protons for Compound I generation
  • Key step: proton transfer from proximal (closest to iron centre) to distal O atom of bound hydrogen peroxide
  • Induce formation of oxo-water by heterolytic splitting of O-O bond
  • may be promoted by ET from porphyrin to Fe-O bond
  • Water mol and ferryl-ozo porphyrin cation radical are formed
  • no other external source of e or protins
  • Imidazole side chain of distal His may be intmt catalyst
  • accept a proton from proximal oxygen on 1st step
  • donates proton to distal oxygen at 2nd step
  • pKa of Fe-coordinated Hoo(H) is 3.6-4.0 for HRP and CcP
  • Iron-coordinated peroxide is anion at neutral and alkaline pH
  • Unstable complex
  • Hydroperoxide is weak ligand
  • experimentally observed kcat is 10^7 M-1 s-1
  • Dissociation rate of HOO(H) may be as high as 10^3-10^4s-1
  • Typical for dissoc kinetics of diatomic ligands in heme proteins
  • lower than rates of water replacement as a ligand in octahedral porphyrin complexes at room temp
  • Protein redox partner reduces heme iron
  • P450 enz must catalyse transfer of 2 protons to distal O atom of bound peroxide anion
  • thru str arrangement and reulgation of proton relay system
  • involves water mols
  • stabilised in active centre thru H bonded network interacting with conserved acid-alchohol pair
  • Mutating residues in network shows enz kinetics and coupling ratio are sensitive to H bonding properties of sites
  • In absence of add catalysis of heterolytic scission thru 2nd protonation of distal O atom  hetereolysis or homlysis of O=O bond is governed by thermodynamic stability of reactant and product
  • If iron peroxide complex is mainly in high spin state in non heme metal enz with weak or moderate ligand field
         --- favourable homolysis of O--O bond
         --- spin state of transition state and product state is high spin
  • For heterolytic O-O scission with 2nd protonation of leaving O atom and foramtion of water mol
          --- 2nd e supplied to peroxide ligand by porphyrin moiety
          --- porhphyrin cation radical is formed
  • determine behaviour of actualy systems in heterolysis/homolysis ratio and product distribution
  • Parameters, pH of solvent, ligation of metal, str, redox properties of porphyrin and peroxide may play important roles
  • may favour spec catalytic pathway
  • by shifting HS-LS equilbrium
  • provide protonation of coordinated peroxide
  • weak O=O bond
  • homolytic decomposition of H2O2 favoured for HS and LS states
  • Porphyrin important as e donor for heterolytic scission of O-O bond in coordinated hydroperoxide to form