+ All Categories
Home > Documents > CHAPTER 1 INTRODUCTION - Shodhganga : a reservoir of...

CHAPTER 1 INTRODUCTION - Shodhganga : a reservoir of...

Date post: 05-Jul-2018
Category:
Upload: buituong
View: 240 times
Download: 0 times
Share this document with a friend
48
1 CHAPTER 1 INTRODUCTION 1.1 CHEMICAL KINETICS Chemical kinetics has been developed in a logical and coherent fashion over the years. However, an understanding of the way we achieved our present ideas on chemical kinetics is a very good basis to understand the subject. The great success of Newtonian mechanics in the area of mechanics and astronomy, which involved the idea of explaining phenomena by simple forces acting between particles. This led scientists in the nineteenth century to introduce such mechanical explanation to all areas involving natural phenomenon. In chemistry, for example, these concepts are applied to interpret chemical affinity or chemical mechanics. Hence understanding of chemical kinetics within this context is important. The concept of chemical kinetics evolved relatively late in terms of the overall studies of reactions and reactivity. The study of chemical kinetics can be traced back to Wilhemy (1850), who carried out the first study of inversion of cane sugar in the presence of acid. He formulated in terms of a first order mathematical expression to interpret the progress of the reaction. Unfortunately, this work went unrecognized until Ostwald (1884) drew attention 34 years later. It may seem strange today that such an idea of studying the variation of chemical affinity with time had not occurred earlier. Farber (1961) had tried to explain this and shown that there were some earlier attempts to study the time evolution of reaction even before Wilhemy (1850).
Transcript

1

CHAPTER 1

INTRODUCTION

1.1 CHEMICAL KINETICS

Chemical kinetics has been developed in a logical and coherent

fashion over the years. However, an understanding of the way we achieved

our present ideas on chemical kinetics is a very good basis to understand the

subject. The great success of Newtonian mechanics in the area of mechanics

and astronomy, which involved the idea of explaining phenomena by simple

forces acting between particles. This led scientists in the nineteenth century to

introduce such mechanical explanation to all areas involving natural

phenomenon. In chemistry, for example, these concepts are applied to

interpret chemical affinity or chemical mechanics. Hence understanding of

chemical kinetics within this context is important.

The concept of chemical kinetics evolved relatively late in terms of

the overall studies of reactions and reactivity. The study of chemical kinetics

can be traced back to Wilhemy (1850), who carried out the first study of

inversion of cane sugar in the presence of acid. He formulated in terms of a

first order mathematical expression to interpret the progress of the reaction.

Unfortunately, this work went unrecognized until Ostwald (1884) drew

attention 34 years later. It may seem strange today that such an idea of

studying the variation of chemical affinity with time had not occurred earlier.

Farber (1961) had tried to explain this and shown that there were some earlier

attempts to study the time evolution of reaction even before Wilhemy (1850).

2

The practical importance of such studies did not exist at the end of

eighteenth century, and it was only with the advent of chemical industry at the

beginning of the nineteenth century that chemists need to consider this

problem. Eventually, this became important for the development of industrial

research at the end of that century. An excellent discussion of this problem is

given by King (1982 and 1984), King and Laidler (1984) in their studies on

the history of chemical kinetics where they analyzed the impact of various

theoretical, experimental and conceptual works of Berthelot and Giles (1862)

and Harcourt and Essen (1866). These researchers are truly considered to be

the founder of this new branch of chemistry called chemical kinetics.

The formation of activated complex during the reaction was

proposed and reported by Arrhenius (1889) in which chemical kinetics was

slow to develop because complex phenomenon involved when molecules

collide to produce chemical reactions. He interpreted the effect of temperature

on reaction rates of a chemical reaction. For rates measured under standard

concentration conditions, Arrhenius expressed this effect by equation (1.1)

k = Ae (Ea/RT) (1.1)

where k is the rate constant under standard conditions and A and Ea are

constants, which are practically independent of temperature. A is called the

frequency factor and Ea is the activation energy. The Arrhenius law took a

long time to become accepted, and from this many other expressions were

derived and proposed to explain the dependence of rate on temperature.

Many of the conceptual and experimental difficulties would

disappear with the work of Vant’s Hoff, who introduced the concept of order

of reaction and proposed the mechanism of a chemical reaction based on the

basis of chemical kinetics. In the first phase of development, the theories of

chemical kinetics tried to resolve the problem of calculation of

3

pre-exponential factor and activation energy in the Arrhenius equation. Lewis

(1918) developed mathematical expressions that allowed the formulation of

collision theory for pre-exponential factor. Henry Eyring (Laidler and King

1983) came forward to develop the transition state theory based on

thermodynamics and statistical mechanics.

The energy differences between reactants and products in solution

are normally measured in terms of their equilibrium constants. The

equilibrium constant of a reaction is related to the energy as given in

equation (1.2)

G0 = RT ln Keq (1.2)

where G0 is the standard free energy, Keq is the equilibrium constant, R is

the gas constant and T is the temperature in absolute. However, the science of

kinetics does not end here. The next task is to look at the chemical steps

involved in a chemical reaction and develop a mechanism. Chemical kinetics

is not just an aspect of physical chemistry but it is a unifying topic covering

the whole of chemistry, many aspects of biochemistry and pharmaceutical

industries. In this context it covers the measurement of rates of reaction and

analysis of experimental data to give a systematic collection of information

which summarises all quantitative kinetic information of any given reaction.

This, in turn, enables comparison of reactions. The sort of information used

here is summarised in terms of the

factors influencing rates of a reaction,

dependence of the rate of reaction on concentration called the

order of the reaction,

4

rate expression, which is an equation which summarizes the

dependence of the rate on the concentration of substances

which affect the rate of reaction,

expression involves the rate constants which is a constant of

proportionality linking the rate with the various concentration

terms,

rate constant collects in one quantity all the information

needed to calculate the rate under specific conditions and

effect of temperature on the rate of reaction. Increase in

temperature generally increases the rate of a reaction. It gives

information to a deeper understanding of a reaction to occur.

1.2 CATALYSIS

The term catalysis was introduced by Berzelius (1835) in order to

explain various decomposition and transformation reactions. He assumed that

catalysts possess special powers that can influence the affinity of chemical

substance. In continuation, Ostwald offered a valid definition for catalysis in a

book of Gates (1992). A catalyst is a substance which alters the speed of a

chemical reaction without itself undergoing any chemical change and the

phenomenon is known as catalysis. Example

MnO2 2KClO3 2KCl + 3O2 (1.3)

The suitable catalyst for an industrial process mainly depends on the activity,

selectivity and stability (deactivation behavior).

5

1.2.1 Types of Catalytic Reactions

Catalytic reactions are classified into two broad types, viz.,

1. Homogeneous catalysis

2. Heterogeneous catalysis

1.2.1.1 Homogeneous catalysis

In this reaction, the reactants and catalyst are in the same phase.Homogeneous catalysts have a higher degree of dispersion thanheterogeneous catalysts since in theory each individual atom can becatalytically active. Due to their high degree of dispersion, homogeneouscatalysts exhibit higher activity than heterogeneous catalysts. The highmobility of the molecules in the reaction mixture gives more collisions withsubstrate molecule. The reactants can approach the catalytically active centerfrom any direction and a reaction at an active center does not block theneighboring centre. This allows the use of low catalyst concentration andmilder reaction conditions. Homogeneous catalyzed reactions are controlledmainly by kinetics and less by material transport because diffusion of thereactants to the catalyst. Hence the mechanism of homogeneous catalysis isrelatively clear and mechanistic investigations can readily be carried outunder reaction conditions by means of spectroscopic methods.

Metal ions, being Lewis acids, can be expected to assist a reactionby drawing the electrons of a bond to themselves and thereby weakening thebond to be broken. Though a proton can do this, metal ions which carry morethan a single positive charge are much more efficient than H+. Thus metalions can function as super acids. For example, polyvalent metal ion such asCu2+, Fe2+, Fe3+, etc., catalyze decarboxylation of -keto acids like oxolacticacid. In this reaction, transition metal ions are found to be better catalyst thannon-transition metal ion like Al3+ because of their ability to coordinate withreactant molecules.

6

Transition metal ions and their complexes catalyze a wide variety

of reactions like hydrogenation, oxidation and polymerization. Apart from

their ability to coordinate with the substrate molecules, the availability of

vacant d-orbitals enables the formation of sigma bonds with ligands which are

Lewis bases. The three t2g orbitals remain non-bonding. However, these non-

bonding electrons on the metal ion can also be donated to suitable -acceptor

ligands. This ability to form both sigma and pi-bonds with suitable substrates

constitutes one of the important factors in catalysis by these ions. Their ability

to assume a variety of oxidation states and different coordination numbers

enable them to act as catalysts in redox reactions. The original objective of

binding metal complexes to insoluble supports is to get over the problem of

separation of homogeneous catalyst from the reaction mixture. A number of

materials have been used as supports for complexes viz., silica, -alumina,

molecular sieves (zeolites) and clays.

1.2.1.2 Heterogeneous Catalysis

The reactants and catalyst (usually solid) are in the different phase

in heterogeneous catalysis. Heterogeneous catalysts provide a surface for the

chemical reaction to take place. In order for the reaction to occur, one or more

of the reactants must diffuse to the catalyst surface and adsorb onto it. After

reaction, the products must desorb from the surface and diffuse away from the

solid surface. Frequently, this transport of reactants and products from one

phase to another plays a dominant role in limiting the reaction rate.

Understanding these transport phenomenon and surface chemistry such as

dispersion is an important area of heterogeneous catalyst research. If diffusion

rates are not taken into account, the reaction rate for various reactions on

surface depends solely on the rate constant and reactant concentration. The

solid catalysts that are used in major industries for various purposes include

food industry (hydrogenation and isomerisation of olefins), fine chemicals

7

(catalytic hydrogenation), petroleum industry (catalytic cracking and

reforming process) and heavy inorganic chemicals (ammonia synthesis and

nitric acid). Some of the solid catalysts are clays, zeolites, mesoporous

molecular sieves and heteroploy acids.

1.3 OXIDATION

The combination of processes involved in the flow of electrons

from a reducing agent (reductant) to an oxidizing agent (oxidant) is oxidation.

The total number of electrons lost by one substance is the same as the total

number of electrons gained by another substance. Oxidation and reduction

always occur simultaneously and are really opposite sides of the same

reaction, which is often called redox reaction. Oxidation involves the loss of

electrons from the reducing agent. Since electrons carry negative charge,

oxidation results in an increase of positive valence. Therefore oxidizing agent

is the substance that brings the oxidation.

Oxidation state of an atom in a chemical compound is counted by

set of rules: (1) the oxidation state of a free element (uncombined element) is

zero, (2) for a simple (monoatomic) ion, the oxidation state is equal to the net

charge on the ion and (3) hydrogen has an oxidation state of +1 and oxygen

has an oxidation state of 2 when they are present in compounds. Exceptions

to this are that hydrogen has an oxidation state of 1 in hydrides of active

metals, e.g. LiH and oxygen has an oxidation state of 1 in peroxides, e.g.

H2O2. The algebraic sum of the oxidation states of all atoms in a neutral

molecule must be zero, while in ions the algebraic sum of the oxidation states

of the constituent atoms must be equal to the charge on the ion. For example,

the oxidation states of sulfur in H2S, SO2 and H2SO4 are 2, +4 and +6

respectively. Hence higher the oxidation state of a given atom, greater is its

degree of oxidation and lower is the oxidation state, greater is its degree of

reduction.

8

1.3.1 Importance of Oxidation

Oxidation reactions play an important role in organic synthesis and

there is currently demand for selective and efficient oxidation methods

(Backwall 2004). Pressure from society has placed restrictions on industrial

oxidation technology because of the need for sustainable and ecofriendly

processes. Today there is an increasing demand to use oxidants such as

molecular oxygen and H2O2 which are environmental friendly and do not give

rise to any waste products (Sheldon et al 2002).

The direct oxidation of organic substrates by either O2 or H2O2 is

rare as the energy barrier for electron transfer from the organic substrate to the

oxidant is usually high. This high energy barrier is the way nature protecting

organic compounds from destructive oxidation. Nature has also found

methods to make controlled aerobic oxidation under highly mild condition.

The unfavourable kinetics associated with direct aerobic oxidation is in the

respiratory chain, which is involved in many biological oxidations (Duester

1996, Gille and Nohl 2000, Berkessel 2006).

Oxidation reactions are of fundamental importance in nature and

are key transformations in organic synthesis (Julio and Backwall 2008). The

developments of new processes employ transition metals as substrate-

selective catalysts and stoichiometric friendly oxidants. Direct oxidation of

the catalyst by molecular oxygen or hydrogen peroxide is often kinetically

unfavored. Selective oxidation reactions based on palladium catalyst led to

excitement in 1950s since the direct reoxidation of palladium by molecular

oxygen is difficult in many cases (Stahl 2005). Therefore Pd catalyst is now

used in the fuel cells and biological enzymes for selective oxidation. Hence

metal catalyzed oxidation has been providing a better alternative route and

they provide new opportunities for industrially relevant reactions. Improved

oxidation methods also provide inventive methods to improve the substrate

9

conversion and product selectivity by increasing the solubility of oxidizing

agent in the reaction mixture. In modern organic synthesis, oxidation of

organic substrates has been carried out in aprotic solvents under mild and

neutral conditions to get maximum product.

1.3.2 Oxidation of Amino acids

Oxidation of amino acids is one of the most prevalent forms of

chemical reactions and is susceptible to modification by a wide array of

oxidants. Uncharacteristic oxidation reactions are of particular concern in

biotechnology and medicine. Therefore, it is important to understand the

amino acid metabolism (Bender 1975) with a model system or model reaction

towards oxidants.

Amino acids have both amino and carboxylic acid functional

groups and therefore they are both acids and bases. In certain compound

specific pH known as isoelectric point, the number of protonated amino

groups with a positive charge and deprotonated carboxylate groups with a

negative charge are equal, resulting a net neutral charge. These ions are

known as zwitter ions. Amino acids are zwitter ions in solid phase and in

polar solutions such as water depending on the pH but not in the gas phase.

Zwitter ions have minimal solubility at their isoelectric point. Simple

amino acids present in municipal wastewater can cause serious eutrophication

in water bodies. It is essential to remove them from wastewater. In textile

industry and tanneries the waste sent out to nearby area contain some simple

amino acids. It rises up the nutrients on that particular place. This may not be

suitable for living plants. It affects their growth and population. Therefore,

before discharge, they may be treated completely and converted into

ecofriendly product.

10

The importance of biologically active pyridoxal phosphate

(Vitamin B6) in the metabolism of amino acids has been well established. It is

the cofactor for enzymes catalyzing a number of reactions of amino acids

such as transfer of amino group to -ketoacids (Braunstein 1973, Metzler

1979, Walsh 1979). The initial reaction of an amino acid with pyridoxal

phosphate is the formation of an intermediate complex, a Schiff’s base, by

condensation between -amino group of the substrate and 4'-formyl group of

pyridoxal phosphate (Braunstein and Shemyakin 1953, Metzler et al 1954,

Metzler 1957, Braunstein et al 1960). The electron withdrawing effect of

heterocyclic nitrogen results withdrawal of electrons from all the three bonds

around the -carbon. The structural requirements for non-enzymic and

coenzyme activity of pyridoxal phosphate analogues are the heterocyclic

nitrogen for electron withdrawal, 4'-formyl group and phenolic hydroxyl

group (Groman et al 1972). Hence in the non-enzymatic model reactions,

heterocyclic nitrogen can be placed by an oxidant which will act as an

electron withdrawer.

It was observed that thermal decomposition of -amino acids in the

presence of aldehydes and ketones is relatively fast (Chatelus 1964). It was

also noted that in the oxidation of -amino acids, the product aldehyde

enhanced the rate of the reaction. Therefore a chemical model pyridoxal-

catalyzed amino acid metabolism can also be constructed using formaldehyde

and oxidant. The efficacy of the model can be measured from the oxidation of

amino acids. However, this study brings out the importance of Schiff’s bases

(aldimines) in the oxidative decarboxylation-deamination of amino acids.

1.3.3 Oxidation of Organic Substrates with Oxidizing Agents and

Metal Ions

Kinetic study of dioxygen complexes has been reported by Martell

(1983). Reaction scheme is described in which oxygen complexes of cobalt,

11

iron and copper may be involved in the oxidation of organic substrates by

electron transfer (oxidase models) and oxygen insertion (oxygenase models).

Many non-metal redox reactions take place by atom transfer or ion transfer

processes. However, much less attention has been paid to electron transfer

reactions. A detailed study of the electron transfer reactions was investigated

by Kumar and Margerum (1987). Based on the study, the following rate

expression was proposed for the oxidation of bromide by HOCl and OCl

THA T

--d[OCl ] - - = k [HA] [OCl ] [Br ]dt

(1.4)

where [OCl ]T = [OCl ] + [HOCl] and HA is the general acid.

The kinetics of oxidation of thiosemicarbazide, thiocarbohydrazide

and hydrazone by hydrogen peroxide in both perchloric and sulphuric acid

media were investigated by Thimme Gowda and Vasi Reddy (1990). The

rates revealed first order kinetics each with respect to [oxidant] and [substrate]

in all the cases.

In aqueous acid solution, the complex [Ag(H2L)]3+ (H2L= ethylene

bis (bigunaide)) quantitatively oxidized ethanol and isopropyl alcohol to the

corresponding carbonyl products at moderate rate (Banerjee et al 1990). The

rate law is expressed as in equation (1.5).

k-d[complex] 2 = k + [alcohol] [complex]+1dt [H ] (1.5)

During the reaction, deprotonation of the alcohols assisted by axial

coordination to [Ag(H2L)]3+ is suggested to be the source for inverse acid

dependence.

12

The kinetics of oxidation of propionic acid by Ce(IV) in aqueous

perchloric acid was studied by Alvarez-Macho and Mata-Perez (1992). The

analysis revealed that the rate of oxidation is proportional to propionic acid

concentration and the over all order is two. The activated complex consists of

different species such as propionic acid, Ce(IV), H2O and H+. Amels et al

(1997) reported the kinetics of oxidation of dimethyl sulfide (DMS) by

hydroperoxides such as hydrogen peroxide, peroxo formic acid, peroxo acetic

acid, peroxo nitrous acid and peroxomonosulfuric acid anion.

The kinetics of oxidation of mandelic acid by permanganate in

aqueous alkaline medium was studied spectrophotometrically (Rafeek et al

1998). The reaction showed first order kinetics over [permanganate ion] and

fractional order dependence over [mandelic acid] and [alkali]. There was no

significant change on the reaction rate upon the addition of aldehyde to the

product. An increase in the ionic strength and decrease in dielectric constant

of the medium increased the rate. The oxidation process in alkaline medium

under the conditions employed in the present investigation proceeded first by

formation of an alkali permanganate complex, which combined with mandelic

acid to form another complex. The latter decomposed slowly followed by fast

reaction between free radical of mandelic acid and another molecule of

permanganate to give the final products. There is a good agreement between

the observed and calculated rate constants under different experimental

conditions.

The kinetics of oxidation of glyoxylic acid by

peroxomonophosphoric acid (PMPA) in acidic medium was reported by Vijai

et al (2008). The stoichiometry corresponds to the reaction of one mole of

PMPA and one mole of glyoxylic acid. The reaction was found to be second

order.

13

CHO

COOH + H3PO5 HCOOOH + CO2 + H3PO4 (1.6)

Ag(I) catalyzed cold aqueous persulphate oxidation of primary

amines are converted into aldehydes or ketones. When applied to secondary

amines of the types (RCH2)2 NH and (R1R2CH)2 NH gave poor yields in

which -amino acids are converted into aldehyde. But the yields are not better

than with other inorganic oxidants (Bacon et al 1966).

(RCH2)2NH + S2O82 RCHO + RCH2.NH2

+ + H+ + 2 SO42 (1.7)

The reaction between malic acid and potassium peroxydisulphate in

the absence of a catalyst was found to be very slow. However, the rate was

enhanced by the introduction of Ag(I) ion as a catalyst (Agrawal et al 1970).

The total order of the reaction was nearly unity. The observed reaction and

rate equation are expressed in equations (1.8) and (1.9) respectively.

Ag+ + S2O82 Ag2+ + 2 SO4

2 (1.8)

22 0.3 +2 8

2 8-d[S O ] = k [S O ] [malic acid] [Ag ]

dt (1.9)

The kinetics of Ag(I) catalyzed oxidation of benzoin and four

substituted benzoins by peroxydisulphate in acid-water mixture was

investigated by Khandual (1990). The rate law was explained on the basis of

free radical mechanism using steady state principle

2+ 22 8

2 8

-d[S O ] = k [Ag ][S O ]dt

(1.10)

14

Maleic anhydride, a precursor for polyester resins, is made by the

oxidation of n-butane over vanadium phosphate catalyst (Coulston et al

1997). The rate of malic anhydride formation is proportional to the rate of

decay of V5+ species in the catalyst. Thus V5+ species are kinetically

significant for the production of malic anhydride and not just for the

production of by-products. The results also suggested that V5+ species may

play a role in the initial hydrogen abstraction from n-butane, the

rate-determining step in the reaction sequence. V4+ sites appeared to be

responsible for the by-product formation.

The oxidation of 6-aminocaproic acid by Ag(III) complex

(dihydroxydiperiodatoargentate(III)) was studied in alkaline medium (Huo

et al 2007). The reaction was first order with respect to Ag(III) complex and

the order with respect to 6-aminocaproic acid was found to be from one to

two. A plausible mechanism involving a pre-equilibrium adduct formation

between the complex and reductant was proposed from the kinetic study.

1.4 LITERATURE REVIEW ON THE OXIDATION OF AMINO

ACIDS WITH OXIDIZING AGENTS

Toennies and Homiller (1942) reported the action of hydrogen

peroxide dissolved in formic acid oxidation of tryptophan, methionine and

cystine. The rates of Caro’s acid oxidation of dimethylaniline and

p-substituted dimethylaniline were measured in aqueous solution at various

pH and temperatures. The rates were found to be proportional to the product

of concentration of both the reactants. The effects of pH and

p-substituents on the rate were estimated. The rate constant in acid medium

increased with increasing electron withdrawing power of the substituents of

dimethylaniline (Ogata and Tabushu 1958).

15

Amino acid oxidation was carried out with a variety of oxidizing

agents to produce a variety of products. Silver oxide oxidation produces

carboxylic acid with one carbon less (Clarke et al 1968). The kinetics of

oxidation of glycine and valine by chloramine-T in hydrochloric acid medium

was studied. The rate of disappearance of chloramine-T showed first order

dependence on both chloramine-T and amino acid and an inverse first order

with respect to [H+] (Gowda and Mahadevappa 1979). The kinetics of

oxidation of glycine, alanine, phenyl alanine, serine, threonine, aspartic acid

and glutamic acid by permanganate in aqueous medium were investigated and

proposed suitable mechanism for the reactions (Surender Rao et al 1979). The

rate law was found to be

0-d[Mn(VII)] = k + [amino acid] [Mn(VII)]

dt (1.11)

Laloo and Mahanthi (1990) studied the kinetics of oxidation of

glutamic acid and aspartic acid by alkaline hexacyanoferrate (III). The rate

depended on the concentration of substrate and oxidant but independent of the

concentration of alkali. During the reaction, there is loss of hydrogen atom

from the C-H bond in a slow step, giving a radical species which was

characterized by ESR spectroscopy. The reaction proceeds with formation of

-imino acid in a fast step, which undergoes hydrolysis to give the

corresponding -keto acid.

Kinetics of oxidation of twelve different -amino acids by

N-chlorosuccinimide (NCS) in aqueous alkaline media were studied and

compared with those of N-bromosuccinimide (NBS) oxidation

(Ramachandran et al 1990). Perusal of the results showed that NCS/NBS

reacted with -amino acid anion to produce -amino acylhypohalite which

then decomposed in the rate-determining step to give the products. The

intermediate -amino acylhypohalite was identified by UV-Visible absorption

16

spectra. But glycine behaved differently from other amino acids with both the

oxidants.

Aliphatic amino acids were smoothly oxidized by

peroxomonophosphoric acid (PMPA) in the pH range 6-10. Kinetic studies

revealed second order dependence on the amino acid, which was not observed

earlier (Panigrahi and Paichha 1991). However the dependence of rate on

[PMPA] was unity. The reaction path included oxidation step due to H2PO5

and HPO5 . Second order dependence on the amino acid was attributed to the

participation of two zwitter ionic molecules of the amino acid to form a

nucleophile.

Electrochemical oxidation of amino acids in the presence of

triphenylphosphine gave -amino aldehydes (Maeda et al 1992). Amino acids

with longer side chains, hydrogen abstraction could lead to the formation of

hydroxyl derivatives via oxy-radical-mediated reaction pathway which could

form alkoxyl and peroxyl radicals (Luxford et al 2002). Studies on the

modification of valine to hydroxide form of valine and leucine to hydroxy

leucine (Fu et al 1995 and Fu and Dean 1997) supported this study.

Kinetics of amino acid oxidation by sodium salt of 2-p-

phenylsulfonicacid-2-phenyl-1-piperylhydrazyl and 2,2-di-p-phenylsulfonicacid -

2-phenyl-1-piperylhydrazyl at isoelectric point of amino acids were studied in

the temperature range 298-318 K. The analysis of the results provided

information about the mode of action during the reaction. The mechanistic

pathway of amino acids oxidation occurred probably by an intermediate of

amino acid radical type which led to keto acids (Ionita et al 2000).

Oxidative decarboxylation of L-ornithine by permanganate in

sulfuric acid medium was found to be first order with respect to both oxidant

and substrate concentration (Usha and Choubey 2006). The kinetics of

17

oxidation of L-phenyl alanine by diperiodatoargentate (III) (DPA) in aqueous

alkaline medium was studied spectrophotometrically (Lamani et al 2009). The

reaction showed first order in [DPA], less than unit order dependence each in

[L-phenyl alanine] and [alkali] and retarding effect of [IO4 ] under the

reaction conditions. The active species of DPA was monoperiodato

argentate (III) (MPA). The reaction is shown to proceed via MPA-L-phenyl

alanine complex, which decomposed in the rate-determining step to give

intermediates followed by a fast step to give the products. The products were

identified by spot and spectroscopic studies.

1.4.1 Oxidation of Amino acids by Oxidizing Agents in the Presence

of Metal Ions

Silver catalyzed persulphate oxidation of amino acids generates

aldehyde with one carbon less (Bacon et al 1966). The reactions are found to

be acid catalyzed, and the kinetic data indicate the participation of water

molecules in the rate-determining step as a proton abstracting agent from the

substrate as per Bunnett's hypothesis. As Ag(I) was found to catalyze these

reactions, the oxidation of glycine and glutamic acid was studied. The rate

law was found to be

k k'' [amino acid] [Ag(I)]- dln[Mn(VII)] c =dt 1+ k [amino acid] = k [Ag(I)]

(1.12)

Studies on the oxidative decarboxylation of amino acids by

peroxydisulphate in the presence of Ag(I) revealed that aldehyde is formed

through radical intermediates (Minisci et al 1983, Zelechonok and Silverman

1992). It served as the model system to understand the catalysis by

monoamine oxidase (Gates and Silverman 1990, Silverman 1992). These

results stimulated interest to study the kinetics and mechanism of oxidative

decarboxylation of amino acids by peroxydisulphate and also in the presence

18

of various metal ions (Chandra and Srivasatava 1971, 1972, 1973, Kumar and

Saxena 1970 and Ram Reddy et al 1978a, 1978b, 1979). The mechanism

formulated in these studies involved the formation of SO4-..

S2O82 2 SO4

. (1.13)

The rate determining step is catalyzed by metal ions. The above

mechanism explained the zero order dependence of rate with respect to amino

acid concentration. The following rate equation is deduced for Cu(II)

catalyzed oxidation of glycine and alanine (Ram Reddy et al 1978b).

22 3 0.52 8

2 8-d[S O ] = k [S O ] [amino acid] [Cu(II)]

dt (1.14)

The kinetics of oxidation of dl-alanine by hydrogen peroxide in the

presence of Fe(II) ions was studied by Ashraf et al (1979). The reaction was

first order with respect to [alanine], [Fe(II)] and zero order in [H2O2]. A free

radical mechanism as shown below was postulated for this reaction in which

carbon dioxide, acetaldehyde and ammonia were the reaction products.

Fe2+ + H2O2 Fe (OH)2 + HO. (1.15)

19

R CH COOH+ OH. H2O

H2O2

OHO

+ NH4+

RCHO + CO2

R C COOH +.

R C COOH

NH3+

NH3+NH3+

NH3+NH3+

R C COOH

Scheme 1.1 Plausible mechanism for the oxidation of dl-alanine by H2O2

in the presence of Fe(II) ion

The kinetics of oxidation of glycine, alanine, phenyl alanine, serine,

threonine and aspartic acid by permanganate in aqueous medium were

investigated by Surender Rao et al (1979). The rate was found to be

-d[Mn(VII)] = k [amino acid][Mn(VII)]dt

(1.16)

The reactions were found to be acid catalyzed and the kinetic data indicated

participation of water molecules in the rate determining step as proton

abstracting agent from the substrate as per Bunnett’s hypothesis. As Ag(I)

was found to catalyze these reactions, the rate law was found to be

K k'' [amino acid] [Ag(I)]-d[Mn(VII)] c=dt 1+ K [Ag(I)]+ K [amino acid]

(1.17)

Kinetics of oxidation of amino acids viz., glycine, alanine, valine,

leucine, phenyl glycine and phenyl alanine by diperiodatoargentate (III) was

20

investigated in alkaline medium by Jayaprakash Rao et al (1985). The

oxidation products were found to be ammonia and the corresponding

keto acids. The presence of electron-withdrawing groups at the -carbon

increased the rate of oxidation while the rate decreased with increasing alkyl

chain length. A two electron transfer mechanism was proposed to explain the

results.

The rate of Mn(II) catalyzed O2 production in the presence of

alanine or leucine was about 4-fold the rate observed in the absence of amino

acids and accounts for about half of the H2O2 consumed. The other half of the

H2O2 was consumed in the oxidation of amino acids. In contrast, O2

production was increased nearly 18-fold in the presence of -methylalanine

and accounts for about 90% of the H2O2 consumed. Oxidation of amino acid

in this complex most likely proceeded by a free radical mechanism involving

hydrogen abstraction from the -carbon. Mn(II) could able to facilitate

Fenton-type reactions (Berlett et al 1990).

2 H2O2 + RCHNH3+COO RCOO + CO2 + NH4

+ + 2H2O (1.18)

Metal catalyzed oxidation (MCO) or non-enzymatic oxidation gave

rise to highly reactive intermediates such as hydroxyl radicals. The hydroxyl

radicals cause simple amino acids to lose a hydrogen atom at the -position

and form a carbon-centered radical, which in the presence of O2 generates

hydroperoxides (Fu et al 1995). These hydroperoxides are unstable in the

presence of redox active transition metal ions like copper and decompose

rapidly to -keto acids, aldehydes and/or carboxylic acids corresponding to

the oxidized amino acid after losing the amine group as NH3 and/or carboxyl

group as CO2 (Stadtman and Berlett 1991).

The oxidation of amino acids by Fentons reagent (H2O2 + Fe(II))

led to the formation of NH4+, -keto acids, CO2, oximes, aldehydes or

21

carboxylic acids containing one carbon atom less. Oxidation is almost

completely dependent on the presence of bicarbonate ion and is stimulated by

iron chelators at levels which are sub-stoichiometric with respect to [Fe(II)]

but is inhibited at higher concentrations. The stimulatory effect of chelators is

not merely due to solubilization of catalytically inactive polymeric forms of

Fe(OH)3 but also due to the conversion of Fe(II) to complexes. The results

suggested that an iron chelate and another form of iron are required for

maximum rate of amino acid oxidation. The metal ion-catalyzed oxidation of

amino acids is likely to be a caged process since the oxidation is not inhibited

by hydroxyl radical scavengers. The relative rates of oxidation of various

amino acids by Fenton system as well as the distribution of products formed

especially products of aromatic amino acids are significantly different from

those reported for amino acid oxidation by ionizing radiation. Several iron-

binding proteins, peptides and hemoglobin degradation products can replace

Fe(II) or Fe(III) in the bicarbonate dependent oxidation of amino acids

(Stadtman and Berlett 1991).

Several metals or metalloid ions exist in multiple oxidation states

and can undergo electron transfer reactions that are important in biological

and environmental systems. There are endogenous metal ions such as Fe(II),

Cu(II) and Co(II) that participate in the oxidation-reduction reactions with

species like molecular dioxygen, superoxide and hydrogen peroxide. These

reactions may be modulated by endogenous reducing agents such as

glutathione, ascorbate and tocopherol. The reactions can be described in terms

of thermodynamics using the standard electrode potentials. The favorable

reaction will depend on the concentration of the reactants and may depend on

the pH and/or the presence of organic ligands that form complexes with metal

or metalloid. As(V) can react with glutathione in buffered aqueous solution to

produce As(III) and oxidized glutathione. This reaction may be important in

the methylation reactions of arsenic. Arsenic species can decrease the red

22

blood cell levels of reduced glutathione, but the products of oxidation and the

mechanism of oxidation are more complex than those found in water alone.

Cr(VI) was found to interact with DNA after reacting first with a reducing

agent such as glutathione to form lower oxidation states of chromium. These

examples illustrate the importance of oxidation-reduction reactions for toxic

metals and metalloids (Carter 1995).

Tungsten catalyzed oxidative decarboxylation of N-alkyl amino

acids with hydrogen peroxide under phase transfer condition gave the

corresponding nitrones with good yield (Murahashi et al 1994). The oxidation

of organic compounds including amino acids by tungsten catalyzed H2O2 was

studied by Miochowski and Said (1977). The kinetic and mechanism of

permanganic oxidation of L-glutamine in sulfuric acid was carried out both in

the presence and absence of Ag(I) (Iloukhani and Bahrami 1999). The overall

rate expression for the oxidation may be written as

- d[Mn(VII)] = k [L- glutamine] [Mn(VII)]°dt (1.19)

The rate law is expressed as in equation (1.20) in the presence of Ag(I),

K k'' [L- glutamine] [Ag(I)]- d[Mn(VII)] c °= [Mn(VII)]dt 1+ K [Ag(I)] + K [L-glutamine]

(1.20)

The kinetics of Os(VIII) catalyzed oxidation of DL-methionine by

hexacyanoferrate (III) in aqueous alkaline medium was studied spectrophoto

metrically (Jose et al 2006). The analysis revealed that there is a decrease in

the dielectric constant of the medium with increase in the rate of the reaction.

The addition of products has no effect on the rate of reaction. Os(VIII) binds

to OH species in the step prior to equilibrium step to form hydroxide species

which reacts with [Fe(CN)6]3 in a slow step to form an intermediate species.

This reacts with a molecule of DL-methionine in a fast step to give sulfur

23

radical cation of methionine and yields sulfoxide product by reacting with

another molecule of [Fe(CN)6]3 .

The kinetics of Cu(II) autocatalyzed oxidation of threonine by well

recognized analytical reagent diperiodatocuprate (III) (DPC) in aqueous

alkaline medium was studied spectrophotometrically (Jose and Tuwar 2007).

The reaction between DPC and threonine in alkaline medium exhibited 2:1

stoichiometry (DPC: threonine). The reaction was first order each in [DPC]

and [threonine] and less than unity in [alkali]. Periodate has retarding effect

on the rate of reaction. Ionic strength has negligible effect on the reaction.

Increase in the dielectric constant of the medium with a decrease in the rate of

the reaction was observed. The main products were identified by spot test and

IR spectra.

Kinetics and mechanism of oxidation of leucine and alanine by

Ag(III) complex were studied spectrophotometrically in alkaline medium at a

constant ionic strength (Song et al 2008). The reaction was first order with

respect to Ag(III) complex and amino acids (leucine and alanine). The

second-order rate constant (k) decreased with increase in [OH ] and [IO4].

The oxidation of amino acid, L-tryptophan (L-trp) by diperiodatoargentate (III)

(DPA) in alkaline medium was studied spectrophotometrically (Tatagar et al

2009). The reaction between L-trp and DPA in alkaline medium exhibited 1:2

stoichiometry. The involvement of free radicals was observed in the reaction

based on the observed order and experimental evidences. The products were

identified by spot test and characterized by spectral studies. The reaction

constants revealed different steps of mechanism.

Kinetics of oxidation of L-cystine by hexacyanoferrate (III) was

studied in alkaline medium at 300 ºC (Nowduri et al 2009). The reaction was

followed spectrophotometrically at max = 420 nm. The reaction was found to

be first order dependence each on [hexacyanoferrate (III)] and [cystine]. It

24

was found that the rate of the reaction increased with increase in [OH-]. The

oxidation product of the reaction was found to be cysteic acid. The oxidation

of glycine and alanine by bis(dihydrogen-tellurto)argentite (III) ion was

studied by stopped-flow spectrophotometer (Huo et al 2009). The reaction

was first order in Ag(III) complex and less than unit order in glycine and

alanine. A plausible mechanism was proposed from the observed kinetic

results.

1.5 PEROXOCOMPOUNDS

The peroxo oxidants are the derivatives of hydrogen peroxide

(H-O-O-H) formed by the replacement of hydrogen atom by the groups of

sulphate, phosphate and carbonate. The weak peroxide bond (–O-O-) makes

the peroxides highly reactive with easily oxdisable molecules. They

spontaneously decompose in solution leading to more stable products. The

–O-O- linkage undergoes cleavage during the reaction and makes sensitive

towards trace amount of catalysts and promoters, which can accelerate the

decomposition. Hence the kinetics and mechanism of oxidation reactions of

amino acids with peroxo compounds are reviewed in the following section.

Peroxo oxidants such as peroxomonosulphate (PMS), peroxodisulphate

(PDS), peroxomonophosphate (PMP) and peroxomonocarbonate (PMC) have

gained paramount importance due to their utilisation as auxiliary reagents in

organic synthesis (Swern 1971, Sosnovsky and Rawlison 1971 and Adam

et al 1992).

1.5.1 Peroxomonosulphate (PMS) (Caro’s acid)

Among the oxidant, PMS (HSO5 ) is a dibasic acid having two

ionisable protons, one of them resembling sulphuric acid proton and the other

peroxide proton (Ball and Edwards 1956). The first pKa value of the peroxide

proton was found to be 9.4±0.2 which is lower than that of hydrogen peroxide

25

(pKa = 11.65). It is also significantly higher than that of peroxoacetic acid. IR

studies revealed that the -O-O- stretching frequency is higher than that of

H2O2 and the two -OH groups are structurally different (Arnau and Giguere

1970).

HSO4 + H2O HSO5 +2H+ + 2e (1.22)

The standard electrode potential (E°) for the couple HSO4 /HSO5

was estimated to be -1.75 V (Spiro 1979). Later Stele and Appleman (1982)

estimated the value as -1.82 V. This high potential value suggested that many

room temperature oxidation can be carried out with PMS. The potentiality of

PMS as a powerful oxidant is brought by its ability to oxidize many organic

and inorganic compounds (Kennedy and Stock 1960). The high oxidation

potential of PMS and the propensity of HSO5 to react via oxygen transfer

(Trost and Curran 1981, Johnson and Balahura 1987 and Edwards and Marsh

1989) make this molecule as a favourable one for the oxidation of various

organic compounds in aqueous solution.

PMS is a two equivalent oxidant and in a number of systems,

oxygen atom transfer occurs from the terminal peroxide position following

nucleophilic attack (at the peroxide moiety) by the substrate (Fortnum et al

1960, Edwards and Muller 1962, Johnson and Edwards 1966, Secco and

Venturini 1976 and Thomson et al 1979). Moreover the superior ability of

HSO4 to act as a leaving group permits HSO5 oxidation proceeded at a

faster rate than other peroxides. PMS could be a viable alternative to H2O2

and S2O82 for the control of sulphide induced corrosion in concrete sanitary

sewers and wastewater treatment facilities. The effectiveness of HSO5 for

H2S oxidation in wastewater from wastewater treatment plant was studied and

compared with H2O2 and S2O82 oxidation. In this report, PMS was found to

26

be more effective based on its molar stochiometry of oxidant used in the total

sulphide oxidation.

1.5.1.1 Decomposition of peroxomonosulphate

PMS in aqueous acetonitrile readily converted aryl and thio

benzoates to carboxylic acid and sulphonic acid respectively (Bunton et al

1995). The rate of reaction increased with increase in the water content in

acetonitrile. Studies on the oxidative hydrolysis of phosphorous(V) esters of

thiols (Blasko et al 1997) by PMS revealed that HSO5 converted

phosphorous(V) esters of thiols into phosphorous(V) and sulphuric acid.

Kinetic and mechanism of decomposition of Caro’s acid was

carried out in alkaline medium (Ball and Edwards 1956). The rate of

decomposition of Caro’s acid is given as in equation (1.22).

-5- d[HSO ]

dt = k [HSO5

-] [SO52-] (1.22)

The rate limiting step is the nucleophilic attack of SO52 on the

peroxide oxygen of HSO5 . Studies on the decomposition of PMS were also

reported by Goodman and Robson (1963). The results revealed that the

kinetics of decomposition was identical with peroxyaromatic acids. The

activation energy was calculated to be 11.9 Kcal/mole. The rate limiting step

proposed is the nucleophilic attack of SO52 on the sulphur atom of HSO5 .

The intermediate then decomposed to give S2O82 and HO2 .

The decomposition of PMS over a wide range of pH was

investigated by Kyrki (1963). The results confirmed the direct interaction

between SO52 and OH in highly alkaline medium. From the results

obtained, the rate equation (1.23) can be derived.

27

5-d[HSO ]dt

= k [SO52 ] [OH ] (1.23)

The mechanism proposed in acid medium involved the formation of

H2O2 by hydrolysis of HSO5 as given in equation (1.24).

HSO5 + H3O+ HSO4 + H2O2 + H+ (1.24)

The rate equation can be given as

5-d[HSO ]dt

= k [H+] [HSO5 ] (1.25)

The rate of decomposition of PMS in 0.5 M perchloric acid is very

slow at room temperature. But the decomposition rate of PMS is remarkably

enhanced by the presence of moderate concentration of Ag(I) (Thompson

1981). Plots of [PMS] vs. time were strictly linear for at least 90% of the total

reaction. The rate expression is

5-d[HSO ]dt

= k (1.26)

Detailed investigation was made on the formation of singlet

molecular oxygen for the ketone- catalyzed decomposition of

peroxomonosulfuric acid by infrared phosphorescence measurements (Lange

and Brauer 1996). Janakiram et al (1998) investigated the mechanism of

oxidation of aliphatic acetals by PMS. The reaction was found to be first order

in [PMS] and [acetal]. The oxidation reaction was independent of dielectric

constant of the medium. A hydride ion shift was proposed in the mechanism.

Kinetics and mechanism of thermal decomposition of PMS by phase transfer

catalyst (PTC) viz., tetrabutylammonium chloride (TBAC) and

tetrabutylphosphonium chloride (TBPC) was reported by Balakrishnan and

28

Kumar (2000). The effects of [PMS], [PTC], ionic strength of the medium (µ)

and temperature on the rate of decomposition of PMS was first order in

[PMS] for TBAC and half order for TBPC.

The kinetics and mechanism of decomposition of Caro’s acid in

aqueous sodium hydroxide in the presence of -cyclodextrin have been

reported. The rate of decomposition of PMS was enhanced by the formation

of -cyclodextrin peroxyanion by the interaction between SO52 and

-cyclodextrin anion. The self decomposition of PMS in alkaline medium

catalyzed by -cyclodextrin involved molecular intermediate rather than free

radical intermediate (Ramachandran and Lathakannan 2002). The kinetics of

ketone-catalyzed decomposition of Caro’s acid was studied in aqueous

alkaline medium at 25 ºC. The rate followed simple second-order kinetics,

first order each in [ketone] and [PMS]. The rate constant values were

independent of hydroxide ion concentration over the range from 0.05 to

0.15 M. The experimental results suggested that the nucleophilic addition of

SO52 ion at the carbonyl carbon led to the formation of oxirane intermediate

in the rate-determining step. Oxirane reacted rapidly with another SO52 to

give the parent ketone, oxygen and SO52 (Selvarani et al 2005). PMS is an

inexpensive, safe and environmentally benign oxidant for C-H bond

oxygenation. It could catalyze acetoxylation, etherification of arene and

alkane C-H bonds. Hence PMS is proved to be effective in acetic acid and/or

methanol. So these transformations are applied to a wide variety of substrates

(Desai et al 2006). The kinetics and mechanism of Mn(II) catalyzed

decomposition of PMS in highly alkaline medium was attempted by Sundar

et al (2008). The value of pseudo first order rate constant (kobs) decreased with

increase in [PMS]0 and increased with increase in [OH-] and [Mn(II)].

Analysis of the results revealed that the formation of manganese peroxide

(MnO2) as molecular intermediate which decomposed rapidly. The rate law is

29

-2

252 5

-- d[SO ] -= k [complex] + k [SO ]1dt (1.27)

1.5.1.2 Oxidation of organic substrates by PMS

The oxidation of halide (Fortnum et al 1960), nitrite (Edwards and

Muller 1962) and chlorite ion (Johnson and Edwards 1966) by PMS was

reported in the literature. The rate limiting step in these oxidation reactions

was found to be nucleophilic attack on the peroxide oxygen. The oxidation of

nitrite by PMS was also studied by Sharma et al (1992a). The product of

oxidation was identified as nitrate.

HSO5 + NO2 SO42 + NO3 + H+ (1.28)

Oxidation of aldehyde by PMS in the presence of alcohols

produced high yield of esters of corresponding acids (Nishihara and Kubota

1968). Further studies confirmed that the rate equation (1.27) is invalid in the

presence of H2O2, S2O82 and metal ions. Isotopic labeling studies revealed

that the peroxo bond in HSO5 was broken in alkaline medium (Koubek et al

1964). Review on the self decomposition of PMS and peroxides were reported

in the literature (Curci and Edwards 1970).

The oxidation of azide by peroxomonosulphate was studied in acid,

neutral and basic solutions, and the oxidation of [Cr(NH3)5N3]2+ by PMS was

studied in acidic medium. The reaction with azide obeyed the rate law:

k1 [N3 ] [HSO5 ] in acid and neutral solution and the rate law: k2 [N3 ]

[SO52 ] in base. The reaction with [Cr(NH3)5N3]2+ obeyed the rate law:

k3 [Cr(NH3)5N3]2+ [HSO5 ] (Thompson et al 1979). This study revealed that

the reaction of azide complex with PMS is the convenient route for the

synthesis of nitrosyl complexes. The reaction proceeded through transfer of

30

terminal peroxide oxygen of PMS to reductant, resulting in the formation of

N2, N2O in azide and [Cr(NH3)5NO]2+in [Cr(NH3)5N3]2+ respectively.

The kinetics of Baeyer-Villiger oxidation of biacetyl and benzyl by

PMS and PMPA was studied in different pH range at 308 K. The rate of

oxidation was strongly pH dependent and the rate increased with increase in

pH. From the pH data the reactivity of different peroxo species in the

oxidation was determined by Panda et al (1988). The detailed kinetic study of

PMS with dimethyl sulphoxide was studied by Pandurangan and

Maruthamuthu (1981). The following products viz., aldehydes (Renganathan

and Maruthamuthu 1986a, Ramachandran et al 1986 and Naseruddin et al

1987), ketones (Manivannan and Maruthamuthu 1986 and Panda et al 1988),

sulphides (Betterton 1992 and Bacaloglu et al 1992), dimethyl aniline (Ogata

and Tabushi 1958), cyclohexene and cyclooctene (Taquikhan et al 1990),

indoles (Balon et al 1993 and Carmono et al 1995) and aniline (Jameel and

Maruthamuthu 1998) were reported. Oxidation of aldehydes to the

corresponding acids by oxone (Caro’s acid) in aqueous acetone was also

reported by Webb and Ruszkay (1998).

The first order (k0) and heterogenetic (k1) rate constants showed

first order dependence on [PMS] and 1/[H+]. The reaction was catalyzed by

the addition of chelating ligand glycine and k1 showed first order dependence

on [glycine] at a fixed pH. This catalysis was ascribed to complexation

whereby the redox potential for Mn (gly)n(2-n)+ was lower than for Mnaq

2+ ,

facilitating the oxidation (Lawrence and Ward 1985). Oxidation of Mnaq2+ to

Mn(II) by PMS in acetate buffer was autocatalytic and obeyed the rate

expression of the general form

0 1 x- d[Mn(II)] = k [Mn(II)] + k [Mn(II)] [MnO ]

dt (1.29)

31

The kinetics of oxidation of aliphatic aldehydes such as

formaldehyde, acetaldehyde, propionaldehyde, n-butyraldehyde and

trichloroacetaldehyde by peroxomonosulphate was carried out in aqueous

perchloric acid medium at a constant ionic strength of 1.2M in the

temperature range 283-333 K (Renganathan and Maruthamuthu 1986a). The

reaction of all aldehydes was found to be a total second order kinetics, first

order each with respect to [PMS] and [aldehyde]. The rate law is expressed as

+5a b

--d[HSO ] = k [PMS] [aldehyde] [H ] + k [PMS] [aldehyde]dt

(1.30)

The kinetics of oxidation of hydroxylamine by PMS in acetate

buffered solution was investigated by Sharma et al (1992b). The reaction

proceeded through a free radical mechanism. Under the experimental

conditions, PMS produced two pairs of possible species viz., OH + SO4.

and OH + SO42 . The reactivity order was S2O8

2 < HSO5 > H2O2. The

kinetic reaction and rate law are as given in equation 1.31 and 1.32.

2 HSO5 +NH3OH+ 2 SO42 + N2 + 3 H2O + 3H+ (1.31)

5-- d[HSO ]

dt= k K [PMS] [NH3OH+] (K + [H+])-1 (1.32)

The oxidation of bisulphate ion by PMS to form sulphate ion was

studied in the pH range 3.8-7.9 (Connick et al 1993). The proposed

mechanism involved the formation of pyrosulphate ion (S2O72-) as an

intermediate, which undergoes hydrolysis to form sulphate and hydrogen

ions. The overall reaction is given as

HSO5 + HSO3 2SO42 + 2H+ (1.33)

32

The rate law derived is given in equation (1.34)

5-- d[HSO ]

dt = (kaaH

+ + kbaH-1 +kc) + [HSO5 ] [HSO3 ] (1.34)

where aH+ is the activity of hydrogen ion and aH

-1 is activity of hydrogen

anion.

Oxidation of hypophosphorous acid by PMS in acid medium was

investigated by Dubey et al (2002). The reaction was first order with respect

to [hypophosphorous] and [PMS]. The oxidation of indole-3-acetic acid by

PMS in acetonitrile medium was also reported by Chandramohan et al (2002).

The reaction followed total second order, first order each with respect to

[indole-3-acetic acid] and [PMS]. The rate of the reaction was not affected by

[H+]. Variation of ionic strength (µ) did not show any significant effect on the

reaction rate. Increase in the percentage of acetonitrile decreased the rate. The

rate of the reaction proceeded through a non-radical pathway. Kinetics of

oxidation of ascorbic acid (AH2) by PMS was determined in acidic, neutral

and alkaline conditions over the temperature range 286-301 K (Raja et al

2003). The reactions were found to obey total second order kinetics, first

order each with respect to [PMS] and [ascorbic acid].

- 2- d[AH ]dt

= k2 [PMS] [AH2] (1.35)

The stoichiometry of the reaction revealed the absence of self

decomposition of PMS. The addition of neutral salt (NaClO4) found to

increase the reaction rate. A suitable mechanism was proposed based on the

formation of free radical intermediates such as hydroxyl, sulphate and

ascorbate. Accelerated bleaching, photobleaching and mineralization of non-

biodegradable azo-dye, Orange II was observed with PMS in the solution of

Co2+ ions. The results revealed that the bleaching rate of Orange II in the dark

33

was found to follow first order kinetics with respect to [Co2+] and reaction

was found to proceed by a chain radical branched mechanism (Fernandez et al

2004). PMS, an efficient oxidant for the photo catalyzed degradation of a

textile dye, acid red 88 was studied by Madhavan et al (2009). The rate of

photodegradation of dye decreased with increase in dye concentration. Total

organic carbon (TOC) analysis revealed rapid mineralization of acid red 88 in

the presence of PMS.

Oxidation of indole by PMS in aqueous acetonitrile was studied by

Meenakshisundaram and Sarathi (2007). Analysis of the results showed that

HSO5 and SO52 are the respective electrophile in acidic and basic media.

Nucleophilic attack of the ethylene bond on the persulphate oxygen was

envisaged to explain the reactivity. The reaction failed to initiate

polymerization and hence radical mechanism was ruled out. The values of

thermodynamic parameters suggested a bimolecular process. Catalytic

activity was not significantly observed for the reaction system in the presence

of Ag(I), Cu(II) and heteroaromatic N-bases. Oxidation of tris (1, 10-

phenanthroline) Fe(II) with PMS ion was first order with respect to both

substrate and oxidant. The rate is accelerated by alkali metal ion due to the

formation of an ion pair between M+ and SO52- ions and not between M+ and

HSO5- ions (Mehrotra and Mehrotra 2008).

1.5.1.3 Oxidation of organic substrates by PMS in the presence of

metal ions

Evidences have been reported for the catalytic decomposition of

Caro’s acid by specific substances in aqueous phosphate buffer. Co(II) and

Mo(VI) are capable and effective catalyst although other metal ions can also

act as catalyst for the decomposition. Cobalt catalyzed decomposition of

Caro’s acid was found to be second order kinetics (Ball and Edwards 1958).

34

Homogeneous catalyst systems are extensively used in the manufacture of

important chemicals. Compared to heterogeneous catalysts, the number of

chemicals manufactured using homogeneous catalyst are quite less. However,

there are reports for the catalytic oxidation of alkenes to aldehydes or ketones

in homogeneous reaction using Pd(II) and Cu(II) catalyst (Smidt et al 1962).

Studies on the uncatalyzed decomposition of PMS in the pH range

6-12 revealed that oxygen is evolved and terminal peroxide oxygen is

incorporated into this oxygen product (Koubek et al 1964). In strongly acidic

medium the product formed is hydrogen peroxide with both oxygen atoms

originating from the peroxide moiety in PMS. However, in the metal ion

catalyzed decomposition neither of the above modes are competitive and

instead a redox process is operative. Metal ion catalyzed decomposition of

PMS in acidic and weakly alkaline medium are widely reported in the

literature (Ball and Edwards 1958, Mariano 1968, Billing et al 1970,

Thomson 1981, Gilbert and Stell 1990a, b and Bennet et al 1991). Metal ions

are observed to influence the rate of decomposition of PMS to a greater

extent. The catalytic effect of different metal ions on the rate of

HSO5 decomposition was studied in acid medium. The results revealed that

the catalytic effect of Ag(I) is profound compared to other metal ions such as

Cu(II), Ti(III) and Co(II) (Rettmer et al 1970).

Detailed studies on Ce(IV) induced decomposition of PMS in

acidic medium (Billing et al 1970) revealed that the peroxy radical generated

from HSO5 is SO5. rather than HSO5

. as shown below.

Ce (IV) + HSO5. Ce (III) + H+ + SO5

. (1.36)

SO5. is proposed as an intermediate in the radiolysis of S2O8

2

(Atkins et al 1963) and in other radiolysis reactions as well (Maruthamuthu

35

and Neta 1977, Huie and Neta 1984 and Deister and Warneck 1990).

However it is reported that SO4. is formed during the decomposition of PMS

by low valency metal ions such as Fe(II) and Ti(III) (Gilbert and Stell 1990a).

Mn+ + HSO5 Mn+1 + OH + SO4. (1.37)

OH. is the proposed intermediate during the decomposition of PMS

at pH ~ 2.0 by Cu(I) which is produced in situ by the reaction between Ti(III)

and Cu(II) as shown in equation (1.38).

Ti (III) + Cu (II) Ti (IV) + Cu (I) (1.38)

However, Fe(II) and Ti(III) react with HSO5 also produce SO4. (Gilbert and

Stell 1990a).

Kinetics of oxidation of some sulphoxides with oxone in the

presence of Ru(III) have been reported (Meenakshisundaram and

Sathiyendiran 2000). The catalytic activity of Ru(III) was demonstrated with

several diaryl, dialkyl and alkyl aryl sulphoxides, all of which were found to

undergo oxidation under homogenous condition. The rate increased

substantially with increasing water content in aqueous acetic acid. The rate

data were consistent with the mechanism involving electron transfer from

electrophilic perhydroxyl oxygen of oxone to sulphoxide.

The kinetics of oxidation of glycolic acid and -hydroxy acid by

PMS were studied in the presence of Ni(II) and Cu(II) ions in the acidic pH

range of 4.05-5.89. The metal glycolate and not glycolic acid was oxidized by

PMS. The rate was found to be first order in [PMS] and metal ion

concentrations. The oxidation of nickel glycolate is zero order in glycolic acid

and inverse first order in [H+]. The increase of glycolic acid decreased the rate

in copper glycolate, and the rate constants initially increased and then

36

remained constant with pH. The results suggested that metal glycolate reacts

with PMS through a metal-peroxide intermediate, which transformed slowly

into a hydroperoxide intermediate by the oxygen atom transfer to hydroxyl

group of the chelated glycolic acid (Shailaja and Ramachandran 2009).

The kinetics of redox reactions of PMS with Fe(IV), Ce(III),

chloride, bromide and iodide ions were reported (Gabor Lente et al 2009).

Ce(III) is only oxidized upon illumination by UV light and Ce(IV) is

produced in the photoreaction. Fe(II) and V(IV) are most probably oxidized

through one electron transfer producing sulphate ion radicals as intermediate.

The halide ions are oxidized by two electron process, which most likely

include oxygen-atom transfer. Comparison with literature data suggested that

activation entropies may be used as indicators distinguishing between

heterolytic and homolytic cleavage of the peroxo bond in the redox reaction

of PMS. PMS ion in the presence of Ni(II) lactate and formaldehyde in the pH

between 4.0 and 5.9 undergo self-decomposition and evolve oxygen. The

reaction is second order in [PMS]. Experimental results revealed that

hemiacetal of Ni(II) lactate catalyzed the self-decomposition (Murugavelu

et al 2009).

1.5.2 Literature Review on the Oxidation of Amino acids by PMS

and Metal Ions

The kinetics of oxidation of amino acids by PMS in the presence

and absence of formaldehyde were studied. Analysis of the results revealed

that the rate can be represented at constant [H+] and in the absence of

formaldehyde as

5-d[HSO ]dt

= ka [amino acid] [PMS] (1.39)

37

in the presence of formaldehyde and at constant [H+], the rate law can be

represented as

5-d[HSO ]dt

= kb [amino acid] [HCHO] [PMS] + ka [HCHO] [PMS] (1.40)

The results showed that formaldehyde catalyzed reaction occur

approximately 105 times faster than uncatalyzed reaction and this is attributed

to the formation of Schiff base (Ramachandran et al 1984b).

Kinetics of oxidation of -amino acids by PMS has been already

reported in the literature and the mechanism is given in Scheme 1.2. The rate

was observed to be first order in [oxidant] and [amino acid] and inverse first

order in hydrogen ion concentration. The kinetic results exhibited that

aldehyde, the oxidation product enhanced the rate of oxidation of all amino

acids except valine. This exceptional behaviour can be attributed to steric

factors due to the presence of methyl group in the -position (Ramachandran

et al 1984b).

38

R' - C - R

NH2

COOH

PMSR' - C - R

NHOH

COOH

[O]R'

R

C

C

O

O

N

O

H

-CO2

= N-OH

[O]

[O]

RCOOH

C

R

R'

= OC

R

R'

If R' = H

Scheme 1.2 Plausible mechanism for the oxidation of amino acids by PMS

Kinetics and mechanism of oxidation of -amino acids by PMS in

acetic acid/sodium acetate buffered medium (pH 3.6-5.2) was reported by

Ramachandran and Vivekanandan (1984c) as given below. The results

showed that SO52 is more reactive than HSO5 . This high reactivity is

attributed to nucleophilic attack of peroxide at the amino group.

39

R- CH -COO-+

-O- O- SO3-

NH3+

R

CHC O-

O

N+H

H

H-O- O- SO3

-

R- C - H + SO42- + H2O + CO2

NH2+

RCHO + NH3

hydrolysis

Scheme 1.3 Plausible mechanism for the oxidation of amino acids by

PMS in acetate buffer

Kinetics of oxidation of amino acids by PMS in aqueous alkaline

medium at 308 K was studied. Based on the experimental results, it could be

concluded that electrophilic attack of HSO5 occurred at the amino nitrogen.

The break down of the intermediate is influenced by the nature of substituents

at the amino carbon atom (Ramachandran et al 1994). Therefore, the observed

rate constant for the disappearance of PMS is

obs 3 k AA k-k K +k K K [OH ]1 1 2 1 2

-[OH ] (1.41)

40

Oxidative decarboxylation of -amino acids with acetone and PMS

was reported by Paradkar et al (1995). The interesting observation is the

isolation of carboxylic acid, ketone and oxime with one carbon atom less than

the parent amino acid. The proposed mechanism involved the formation of

N-hydroxy amino acid (RR’CH-(NHOH)-COOH) in the first step, which

further oxidized to -nitrosocarboxylic acid with excess PMS. The -nitroso

carboxylic acid decarboxylated via a cyclic transition state leading to the

formation of oxime.

Kinetics and mechanism of decarboxylation of -amino acids by

PMS in acetic acid/sodium acetate buffered medium was studied extensively

for different amino acids by Sayee Kannan and Ramachandran (2003).

Comparison of the results in glycine and N-methylglycine revealed that

autocatalytic effect is more pronounced in N-methylglycine. This suggested

that the formation of Schiff base is not the reason for autocatalysis as reported

earlier. The authors proposed that the formation of hydroperoxide is

responsible for autocatalysis.

RCH (N+H3) COO + HOOSO3 RCHO + NH4+ + CO2 + SO4

2

(1.42)

RCHO + HOOSO3 Hydroperoxide (I) + SO42 (1.43)

RCH (N+H3) COO + I 2RCHO + NH4+ + CO2 (1.44)

Sundar et al (2007) reported the kinetics and mechanism of

oxidation of lysine by oxone in acetate buffered medium (pH = 3.6-5.2) at

308 K. The rate of disappearance of oxone at constant [lysine] and pH is

-d[oxone] = k [lysine][oxone] + k [acetate][oxone]1 2dt (1.45)

41

The experimental results revealed that autocatalysis route is absent because

the formation of the product is 6-amino-2-oxohexanoic acid.

Kinetics and mechanism of oxidation of -alanine by PMS in the

presence of Cu(II) ion at pH 4.2 (acetic acid/sodium acetate) was reported by

Sayeekannan et al (2008). Autocatalysis was observed only in the presence of

Cu(II) ion and this was explained due to the formation of hydroperoxide

intermediate. The rate constant values (k1obs) are independent of [acetate] but

the rate constant values (k2obs) decreased with the increase in acetate

concentration.

22 T 1

21 1 2 T

k K[Cu(II) [H ](K +[H ])obsk =[H ](K +[H ])K K K ( -alanine)a

(1.46)

The kinetics of Mn(II) catalyzed oxidative decarboxylation of five

different amino acids such as alanine, valine, leucine, phenyl alanine and

2-methyl alanine by PMS in alkaline medium were studied by Kutti Rani et al

(2009). The observed rate was found to be first order in [PMS], [amino acid]

and [Mn2+] and inverse first order in [OH-]. The effects of added product,

ionic strength and dielectric constant were investigated. The (MnO(O)) or

manganese peroxide was proposed as the intermediate to enhance the

reaction rate and confirmed by IR and GC-MS.

1.6 SILVER(I) CATALYZED OXIDATION OF ORGANIC

SUBSTRATES BY OXIDISING AGENTS

Several industrial catalysts have been developed and

commercialized based on Ag(I), yet many may be launched in the years to

come. The selected literature survey revealed the broad scope of potentially

promising applications of Ag(I) catalysts in alkylation, esterification,

42

transamination and decarboxylation reactions. Due to the unique

physicochemical properties, Ag(I) can be profitably fit in homogeneous

systems providing a broad operational choice. In many cases Ag(I) can

provide high activity and selectivity. Hence Ag(I) can play a vital role both in

acidic and alkaline medium. Ag(I) catalyzed oxidation of secondary aliphatic

amines and -amino acids by persulphate was investigated by Bacon et al

(1966). It is a useful preparative method to convert secondary aliphatic

amines into aldehydes. It is also proved to be successful in oxidation and

depends upon the ability of the amine to coordinate with Ag(I) or Ag(II) ions

in competition with hydroxyl ions or other potential ligands in aqueous

system. Ag(I) catalyzed oxidation of malic acid by peroxydisulphate was

studied by Agarwal and Bhattacharya (1970). The experiments showed that

the reaction in the absence of catalyst is slow. But the rate is enhanced by the

introduction of Ag (I) as a catalyst.

Uncatalyzed oxidation of organic substrates by peroxydisulphate

ions is usually quite slow but the reaction rate is enhanced in the presence of

Ag(I) (Chandra and Srivastava 1972). Surender Rao et al (1979) investigated

Ag(I) catalyzed oxidation of amino acids by KMnO4 in aqueous medium. The

study proposed the mechanism based on the electron transfer process from

metal to oxidant which inturn to reductant. Usha et al (1977) reported Ag(I)

catalyzed oxidation of amino acids such as glycine, -alanine, -alanine,

aspartic acid, glutamic acid and threonine by Co(III) in sulphuric acid. The

rate law is obtained by assuming the formation of an adduct between amino

acid and Ag(I) in a fast step. This adduct reacted with Co(III) in a slow step

yielding Ag2+-substrate adduct, which ultimately undergoes internal oxidation

in a fast step to give the products.

43

The kinetics of oxidation of water with Bi(V) in presence of Ag(I)

was investigated in a mixture of HClO4 and HF (Inani et al 1990). The

reaction was overall second order, first order each with respect to Bi(V) and

Ag(I). However, the rate is independent of hydrogen ion concentration. A

comparative analysis of these results with the results obtained for PDP and

Ag(I), and PDS and Ag(I) was made to correlate the rate constants and redox

potentials of the oxidant couples. Silver (I) catalyzed oxidation of aspartic

acid by Ce(IV) was studied in acid perchlorate medium (Indu Sharma et al

1995). Ce(IV) exists mainly as a mixture of several species such as Ce4+,

Ce(OH)3+, Ce(OH)22+, (CeOCe)6+ and (HOCeOCeOH)4+ in perchloric acid.

From the reports, it is suggested that Ce(IV) attacked carboxyl group bonded

to Ag(I) instead of amino group bonded to Ag(I) and the formed polymeric

species have significantly high molecular weight. However, such polymeric

species create problems in the analysis of kinetic data in perchloric acid

medium. Therefore, Ag(I) catalyst was introduced in such reactions to help

the fragmentation in order to make easier for kinetic data calculation.

The kinetics and mechanism of permanganic oxidation of L-

glutamine in sulfuric acid was carried out both in the absence and presence of

Ag(I) using spectrophotometric technique (Iloukhani and Bahrami 1999). The

effect of complexing agents like sulphate and pyrophosphate ion was found to

be absent, from which Mn(VII) was confirmed as reactive species. In

presence of Ag(I), the order in [permanganate] was unity and [L-glutamine]

and [Ag+] were fractional. This is due to one involving two electron

mechanism and the other one electron transfer.

44

Kinetics of Ag(I) catalyzed autoxidation of aqueous sulphur(IV) in

acetate buffered medium obeyed the rate law in equation (1.47) as given by

Gupta et al (2000).

2 + -1- d[S(IV)] D [Ag(I)] [S(IV)] [H ]=dt (B + C [S(IV)])

(1.47)

The rate is independent of [O2] but strongly inhibited by EtOH. The

study revealed that Ag(I) catalyzed reaction is strongly inhibited by

autoxidation. Caraiman et al (2003) investigated the reactivities of Ag(I) and

Cu(II) complexed with glycine in the gas phase towards three neutral

molecules. Amino acids are known to exist in the neutral form (Iijima et al

1991) in the gas phase while zwitter ionic form in solution phases (Wada et al

1982). A modest catalytic effect was identified theoretically for the influence

of CO and NH3 in the interconversion between charge solvated and metal salt

adduct ions. Ag(I) catalyzed exchange of coordinated cyanide in

hexacyanoferrate(II) by phenylhydrazine in aqueous medium was studied by

Naik et al (2007).

1.7 SCOPE OF THE PRESENT INVESTIGATION

It is important to understand the mechanism of oxidation of amino

acids proceeding through the formation of Schiff base intermediate with

pyridoxal phosphate in living systems. Biological reactions such as trans-

amination, racemization and decarboxylation in living systems are suggested

to proceed via Schiff base intermediate. As this Schiff base is a tridentate

ligand with high coordinating capability compared to either pyridoxal

phosphate or amino acids which are bidentate, it readily forms complex with

any redox metal ions. In order to understand the mechanism of oxidation, in

45

the present study pyridoxal is replaced by a metal ion, namely, Ag(I) and

oxidant, PMS.

Oxidation of amino acids has been carried out with peroxo oxidants

such as H2O2, PDS, PMS and PMP with an objective of designing model

system to understand enzymatic oxidation of amino acids. Among the

oxidants, PMS is important. Though the oxidation of amino acids by PMS has

been studied, it is found that the reaction is slow. It is reported that this

oxidation could be catalyzed by metal ions such as Cr(III), Fe(II), Mn(II) and

V(IV). Ag(I) is reported to catalyze oxidation of amino acids by

peroxodisulphate, oxidation of aspartic acid by cerium and oxidation of amino

acids by KMnO4. Although oxidation of amino acids by PMS has been fully

exploitated, the same reaction has not been studied in the presence of Ag(I).

This particular study has been carried out to understand the influence of the

catalyst in the oxidation. Hence in the present investigation the role of Ag (I)

catalyzed oxidation of amino acids by PMS is undertaken.

Amino acids can associate with alkali, alkaline earth and quaternary

ammonium ions. Such interactions are called ion-pairing. In this respect Li+,

Na+ and Ag+ affinities are often found parallel to each other. However, Ag+

affinity is stronger and easier when compared to alkali and alkaline earth

metals. This is due to specific d-orbital interaction. So Ag(I) adducts play an

important role in the present investigation. The present investigation is based

on the following facts (i) development of highly efficient oxidation protocols

and (ii) oxidation of amino acids by PMS catalyzed by Ag(I). It is proposed to

study the effect of Ag(I) in the oxidative decarboxylation of amino acids by

PMS. The experimental results of the kinetics and mechanism of oxidation of

amino acids by PMS in presence of Ag(I) are discussed in the forthcoming

46

chapters. The following amino acids are chosen for the present study and the

structures are given below.

Glycine Alanine

Phenyl alanine Valine

Leucine Serine

47

Scope and objectives of the present investigation are

1. Kinetics of oxidation of amino acids such as glycine, alanine,

phenylalanine, leucine, valine and serine by PMS in the

presence of Ag(I) in the pH range 1- 3

influence of ionic strength and dielectric constant on the

rate of the reaction

autocatalytic behaviour and free radical intermediates

effect of product and addition of sulphate ion

influence of temperature on the rate of the reaction

calculation of activation energy

identification of adduct and number of electron transfer

involved in the oxidation using UV-Visible spectroscopy

and cyclic voltammetry

analysis of the products of oxidation by FT-IR

spectroscopy

2. Kinetics of oxidation of amino acids such as glycine, alanine,

phenyl alanine, leucine, valine and serine by PMS in the

presence of Ag(I) in the pH greater than 9.

influence of ionic strength and dielectric constant on the

rate of the reaction

autocatalytic behaviour and free radical intermediates

effect of product and addition of sulphate ion

influence of temperature on the rate of the reaction

48

calculation of activation energy

the identification of adduct and number of electron

transfer involved in the oxidation using UV-Visible

spectroscopy and cyclic voltammetry

analysis of the products of oxidation by FT-IR

spectroscopy

3. Kinetics of oxidation of amino acids such as glycine, alanine,

phenyl alanine, leucine, valine and serine by PMS in the

presence of Ag(I) in the pH range 3.5 - 5.5

influence of ionic strength and dielectric constant on the

rate of the reaction

autocatalytic behaviour and free radical intermediates

effect of product and addition of sulphate ion

influence of temperature on the rate of the reaction

calculation of activation energy

identification of adduct and number of electron transfer

involved in the oxidation using UV-Visible spectroscopy

and cyclic voltammetry

analysis of the products of oxidation by FT-IR

spectroscopy


Recommended