You are on page 1of 42

Chapter 1 Introduction

Electron transfer (ET), one of the simplest of chemical events, profoundly

affects chemical reactivity by inverting normal electron densities in an electron

donor-acceptor pair, thus activating previously inaccessible reaction modes.

To the extent that most energy, and all life, on earth derive ultimately from

photosynthesis, our very existence depends on this fundamental process.

It is hardly surprising, then, that examples of electron transfer activation abound

within every sub-discipline of science (Figure 1).

ET»( fir at ETat ETai


li<til>d>liquid polymer-liquid umiconductor modified
inlerfaces inlerfacei elecirodegE elecirodc*
T X
ETsI chemHumineicenns
BirmandEo mtui elsclrodes
rate 1940$ inverted solw eneigy
effect convenion
ET in »olid« wid polyme«l
ETat coltoids -;—r
photoiynihetit
quantum
•ndmicelkt
ot^anicETs cheinistty
calcttlaiions
solvent <iyn*fflit:s Electron TmrnCer
andET in (he
latelSWs&igsO-t crots-reactions
•on pairs.
ivcomwtHKioa, nMhytand
it«e«pe other tiamfert
nugpietic ettectt coupled ET and
proion transfer long range BT
onET in rigid media
charge transfer ETactosi rigid
organic Mtlgedi ETinproteiaa |
spectra

Bgui« I: Developinents in theckeiron transfer field

The intimate mechanisms of electron transfer between metal-ion

complexes in solution have been a source of intense study more than five

decades. Systematic investigations that emphasized the interplay among

thermodynamics, electronic structure, stereochemistry and kinetics of inorganic

systems, in particular those involving transition metal complexes did not appear
Chapter 1 Introduction

in the literature until the early 1950s. Perhaps the single, most in-fluential paper

at that time was Henry Taube's classic review [1] in which the substitution

properties of metal-ion complexes were surveyed and their lability or inertness

was related to their electronic structures, that is, the foundation for structure-

reactivity relationships in inorganic chemistry have been constructed.

A great deal of work in inorganic reaction mechanisms originates in Taube's

review [2]. In this review, a direct application of the relationship between rates

of substitution and electronic structure was the study of the reaction between

[Co(NH3)5Cl]^^ and [Cr(H20)6]^^. In this study, geometric details about the

mechanisms of oxidation-reduction reactions were elucidated, inner-sphere

pathway for electron transfer was clearly defined, and a whole new field of

chemistry, began. Since then, a number of reviews [1-4, 11, 12, 16-80] on

electron transfer reactions have appeared, mainly on the reactions of transition-

metal complexes (particularly, Co(III), Cr(III), Ru(III) and Rh(III)).

The identity of these reactants is preserved due to their substitution inertness.

Taube and coworkers [2-5] have made an extensive review of the

theoretical aspects of electron transfer [2, 3], electronic delocalization in mixed

valence molecules [4] and the effect of organic ligands as bridging groups in

electron transfer reactions [5]. Sutin et al. [6-14] has given a good account of

the structural effects on electron transfer [6-8] with special relevance to fi-ee

energy barrier on reactivity pattern [9], role of nuclear fi-equency factor [10, 11],

'Photochemical and Thermal Reactions of Some Aggregated Systems'


Chapter 1 Introduction

energetic and dynamic of solvent reorganization [12-14] in electron transfer

reactions. Haim et al. [15, 16] has surveyed the mechanism of electron transfer

reactions in terms of the bridged activated states. Sykes et al. [17, 18] indicated

the formation of Co(III)-Cr(III) intermediates in the Cr(II) reduction of bi- and

tetra-nuclear Co(III) complexes.

Thermodynamic and kinetic aspects of cation-macrocylic complexes

[19, 20] and their relation to structure [21] have been extensively reviewed.

Davies [22-24], Housecroft [25, 26] and Hay [27] made a thorough survey on

the redox chemistry of metal complexes in solution. Electron exchange and

electron transfer in metal polynuclear complexes in proteins [28] and across

vesicle bilayers [29] were studied and have been used as a tool to determine

active centers in proteins [30-34]. Evans [35], Andrieux and coworkers [36]

made an extensive survey of the electrochemistry of electron transfer reactions

involving organic substrates and organometallics.

The electron transfer field subsequently expanded considerably into

diverse areas such as reactions at interfaces [37-46] (metal-liquid,

semiconductor-liquid, liquid-liquid, liquid-polymer-metal, inorganic nano-

building blocks and redox-active dendrimers), stereoselectivity [47],

photoinduced electron transfers [48-63] (solar energy conversion,

photosynthesis, supramolecular assemblies and charge transfer spectra of donor-

acceptor systems), and solvent dynamics of charge transfer systems [48, 59, 64].

'Photochemical and Thermal Reactions of Some Aggregated Systems'


Chapter 1 Introduction

Various quantum mechanical aspects of electron transfer problems were treated

in detail and ab initio molecular orbital calculations of exchange rates were

initiated [65-71]. A large amount of knowledge about redox processes occurring

in metalloenzymes has been acquired through the investigation of biologically

relevant electron transfer reactions. In the biological systems, particular

attention has been given to the evidence of the long-range electron transfer and

the factors governing the rate and direction of the transport [72, 73]. Electron-,

energy- and atom-transfer reaction between DNA and metal complexes has also

been reviewed [74-77].

The redox chemistry of Co(III) complexes has attained considerable

importance due to its methyl transfer capacity as witnessed in methylcobalamin

[78-80], which transfer its coordinated methyl group as anion, free radical or as

cation depending upon the nature of the metal-ion being methylated.

During such biomethylation, the metal-ion in methylcobalamin has been found

to be reduced to Co(II) or Co(I) state. The overall understanding of a chemical

change requires the identity, nature of intermediates formed between reactants,

products to be identified, and knowledge of the process of bond making and

breaking and of electron transfer to be established. Inorganic reaction

mechanisms are thus of intrinsic interest.

'Photochemical and Thermal Reactions of Some Aggregated Systems'


Chapter 1 Introduction

1.1. General aspects

The field of electron transfer has been most active since its beginnings in

1953. During the past several years new avenues of inquiry in the areas of

intramolecular electron transfer, intervalence transfer, excited-state electron

transfer, radical ion electron transfer, and bioinorganic electron transfer have

continued to challenge some of the best minds in chemistry.

The essential similarities or difference between oxidation-reduction

reactions are revealed only if one examines the elementary electron transfer

steps themselves after having carefully identified them.

Elementary electron transfer reactions between metal complexes are

classified into two distinct categories:

1. Inner-sphere reactions (IS) and

2. Outer-sphere reactions (OS)

These are the two distinct mechanisms recognized for redox reactions,

which will be outlined in this section and then described in detail in the sections

that follows. The essential difference between these mechanisms lies in the key

role played by substitution prior to electron transfer in the inner-sphere route.

Outer-sphere redox reactions simply involve electron transfer.

'Photochemical and Thermal Reactions of Some Aggregated Systems'


Chapter 1 Introduction

(a) +a + o

electron transfer through

(b)

reductant ^ ' oxidant


direct electron transfer

Figure 2. The two mechanisms for the transfer of electrons from reductant to oxidant,
(a) The irmer-sphere mechanism, (b) The outer-sphere mechanism.

In soi inner-sphere reaction, the oxidant and reductant are Hnked by a

common bridging ligand through which the electron transfer occurs.

The earliest reaction of this kind was demonstrated by Henry Taube and

coworkers [81, 82] on the reduction of substitutionally inert ammine complexes

of Co(III) by substitutionally labile [Cr(H20)6]^^ in acidic aqueous solution.

i2+
[Co(NH3)5a]^-^ + [Cr(H20)6]2+ =s^ [(NH3)5CoClCrl 4+

5ir

2+
[Cr(H20)5Cir + [Co(H20)6r^+ 5NH4^

'Photochemical and Thermal Reactions of Some Aggregated Systems'


Chapter 1 Introduction

From the appearance of CI" in the inert Cr(III) product, it was deduced

that when electron transfer occurs both metal centers must have been bonded

simultaneously to CI" as the bridging ligand.

When the ligands of both reactants are tightly held and there is no change

in the coordination sphere of the two complexes, the reaction proceeds by outer-

sphere electron transfer. In an outer-sphere reaction, electron transfer occurs

without the utilization of a bridging ligand. For example [83-86],

[Co(NH3)5X]^^ + [Ru(NH3)6]^^^ [Ru(NH3)6]^^ + Co^^ + 5NH3 + X "

The great simplicity of this mechanism (Figure 2) makes positive proof

very hard to obtain. There are a number of inorganic reactions where it is

extremely difficult to imagine an inner-sphere mechanism operating, and which

are therefore firmly believed to go by way of an outer-sphere mechanism.

Such reactions are those in which both the oxidant and reductant are

substitutionally inert and yet electron transfer is very rapid.

1.2. Electron-Exchange reactions

The electron-exchange reactions of the type M^"^'^ / M"^ is generally

studied by isotopic labeling of one of the reactants:

Fe^^ + *Fe^'' ^ Fe^^ + *Fe^^

The electron-exchange reactions between +2 and +3 oxidation states of

iron, chromium and cobalt have been extensively studied [87]. In the case of

'Photochemical and Thermal Reactions of Some Aggregated Systems'


Chapter 1 Introduction

Fe^^/Fe^^ reaction, details of mechanism are not available, since both Fe^^ and

Fe are labile. However, the electron-exchange reactions studied in the

presence of anions [88] (other than perchlorates) showed an increase in rate,

which was attributed to the participation of anions in the electron-exchange

reaction leading to an inner-sphere process.

Sutin and coworkers [89-91] have demonstrated an inner-sphere

mechanism for Fe^^/FeCl^* and Fe^^/Fe(NCS)^^ systems. The electron-exchange

reactions in [Fe(EDTA)]^7[Fe(EDTA)]- and [Fe(phen)]^V[Fe(phen)]^^ couples

[92, 93] are found to be fast, and so an outer-sphere mechanism predominates.

The faster rate is probably related to the fairly extensive delocalization of

electrons from the central metal ion.

The electron-exchange between spin-free Co(II) and spin-paired Co(III)

complexes may be represented as follows:

D
[Co(NH3)6]^^ (low spin d^) [Co(NH3)6]2+(high spin d')

71 4-Hh4
[Co(NH3)6]2^ (high spin d^) [Co(NH3)6]^^ (low spin d^)

Figure 3. Schematic diagram of the d orbitals of [Co(NH3)6]^^'^^ showing the


multiple electronic changes required in the self-exchange process.

'Photochemical and Thermal Reactions of Some Aggregated Systems'


Chapter 1 Introduction

It can be seen that the electron-exchange reaction involves not only the

transfer of an electron but also the rearrangement of the other d electrons of the

two reactants (Figure 3). It is partially spin-forbidden and should proceed very

slowly. However, there will be no spin multiplicity restriction if the exchange

can occur between two spin-free or two spin-paired configurations. Baker,

Basolo and Neuman have proposed [94] that this is the case in the

[Co(H20)6] and [Co(phen)3] exchanges, respectively. In the first

system, it is probable that only a small amount of energy is required to excite the

spin-paired [Co(H20)6]^^ to the spin-free state; and in the second system, the

excitation energy for the formation of a spin-paired Co(phen)3^"^ from the spin-

free state is probably also small. The field strength for ammonia,

ethylenediamine, EDTA, and PDTA lie between those of water and

phenanthroline. The slow rates of the [Co(NH3)6]^^^^^ [Co(en)2]^^^^^,

[Co(EDTA)]"^^", [Co(PDTA)]'^^" exchanges may reflect the large multiplicity

restrictions and inner-shell reorganization energies expected for ligands of

intermediate field strength [87].

1.3. Outer-sphere electron transfer reactions and Theoretical


consideration

The coordination spheres of the two metal centers remain intact during

outer-sphere reactions, since substitution into the inner-sphere is slower than

electron fransfer. Typical outer-sphere reductants are [Ru(NH3)6] and

[Cr(bipy)3]^^ [95-97]. Reduction reactions [98, 99] involving V(II), Cr(II) and

'Photochemical and Thermal Reactions of Some Aggregated Systems'


Chapter 1 Introduction 10

electron-exchange reactions in [Fe(CN)6]'^/[Fe(CN)6]^~ and [Co(NH3)6]^V

[Co(NH3)6]^^ systems were also shown to follow outer-sphere path [100, 101].

A theoretical treatment of outer-sphere electron transfer reactions was given

by Marcus [102, 103],

The various steps proposed in the outer-sphere reactions are:

1. Collision of the donor with the acceptor to form the precursor complex
D+A • [D II A]

2. Thermal activation of the precursor complex to reach non-equilibrium


optimized nuclear and electronic configuration for ET to occur.

[D||A] •[D||A]*

3. Electron transfer

[D||A]^ •[D^llA-]^

4. Relaxation to the ground state of the successor complex

[D"||A-]* • [D^ll A-]

5. Dissociation to the free product ions

[DM|A-]* • D^ + A"

where, || denotes that no chemical bonds have been made or broken.

The first step involves the formation of a precursor complex from the reactants.

In the second step, the precursor complex reorganizes itself to form the activated

complex, i.e., inner-coordinated shells of the reactants and solvent molecules

'Photochemical and Thermal Reactions of Some Aggregated Systems'


Chapter 1 Introduction 11

reorganize, in order to adjust to the configuration appropriate to the activated

complex. In the third step, electron transfer takes place during the

reorganization of the precursor complex as required by the Franck-Condon

principle. The fourth step involves the deactivation of the activated complex

resulting in the formation of the successor complex. In the fifth step,

dissociation of the successor complex takes place to form the separated

products. In terms of the above scheme, the rate constant, A:obs for the outer-

sphere electron transfer reaction [103] is given as,

;tobs=Zpet-«'«^^^J.ef-^°*^^Tl {a)

where, Z is the collision frequency between the molecules in solution, and p is

transmission coefficient, G)(r) is the free energy associated in bridging the

reactants together, as defined in the following equation.

expCpaAl'-'O expCpaoI'^)
(r) = X exp(-p rl'^O (*)
2Dsr +
l-KPa^l'/O l+(paDl'/0

where, ZD and ZA are the charges on the two reactants, e is the unit of electric

charge, r is the intemuclear distance in the collision complex, I is the ionic

strength, p is the Debye-Huckel length equal to (87iNe^/1000DsRT)''^' and cJAand

CTD is equal to the radius of the reactant aA and ao, plus the radius of the main ion

of opposite charge in the reactant's ion atmosphere, and Ds is the static

dielectric constant of the medium. Equation {b) reduces to the equation (c)

'Photochemical and Thermal Reactions of Some Aggregated Systems'


Chapter 1 Introduction 12

when the distance of closest approach of the metal centers is equal to the sum of

the radii of the reactants [r =CT= (an + aA)] and the radii of all the ions are equal.

Further AG* [in equation (a)] is the minimum free energy increase above the

thermal energy, RT required in the vibrational and solvent trapping mode to

enable electron transfer to occur with energy conservation.

AG = - ^ 1 +
4 \ X

AG° is the standard free energy of reaction (and equals zero for a self-exchange

reaction), A, is a "reorganization term" composed of solvational (A,o) and

vibrational (A,i) components, 4AG* = X,i + A,o ,= X

Marcus [103] derived an expression for XQ, being the free energy

difference between reactants (D,A) and products (D*, A'), assuming the optical

(Dop) and static (Ds) dielectric constants of the solvent, and the charge

transferred (Ae) from one reactant to the other,

Xo= {/^ef [(1/Dop) - (1/Ds)] [ (1/ 2ai) + (1/1^2) - (1/r)]

where, ai and a2 are the molecular radii of the reactants and r the intermolecular

separation between them. The expression for the vibrational term A,i is given by

j V + fj

'Photochemical and Thermal Reactions of Some Aggregated Systems'


Chapter 1 Introduction 13

where, f/ andf/ =/** normal mode force constants in the reactants and products

respectively. Breathing vibrations are often employed and / = mean of the

breathing force constants

Aq j = change in equilibrium value of the j * normal coordinate, and when


breathing vibrations are employed

Ad = the difference of the metal-ligand distance between oxidized and


reduced complex

The free energy term, A,o gives the solvent contribution to the energy of

optical transition:

Red, Oxd -^ red^. Ox"

The experimentally observed rate constant, ^obs is related to the electron

transfer rate constant (ket) and the association constant between reactants (KA) as

follows:

^obs = KA ket = KA Vet exp {- (A,+AGet> 4A, RT }

where, Vet, the pre-exponential term, and KA were shown to be as follows:

Vet = (2 7uV^/h).(7t/A,RTf^

where, v is the electronic coupling term, and the association constant KA is,

KA = (4 7t NrVsOOO). e ^-<'''>''^'^

There more common approach of testing the predicted dependence of /^obs on

AG* has been based on the so-called Marcus cross-relation, which relates the

rate constant for a net reaction,

Oxi + red2 • red,+ 0x2

'Photochemical and Thermal Reactions of Some Aggregated Systems'


Chapter 1 Introduction 14

with the equihbrium constant, K12 = e ("^°*^^'^) of the cross ET reaction, and kn,

is the rate constant of the cross redox reaction for the self-exchange redox

reaction with the rate constants, kn and kn.

^11
Oxj + redi red, + Oxj

Ox2 + red2 ^22


redj + 0x2

The various free energy terms could be written as follows;

^12 — (^11 +X,22y 2 ;

hi2 =(A,i,ii+Xi,22)/2;

^0,12 = ( V 1 1 •'•^ 0,22)/ 2

Solving for A., for each individual self-exchange reaction, e.g.,

A,ii = 4RT (InVetKA)i 1/^11, inserting the expressions for A-n and A,22 to get A,i2, and

incorporating A,i2 into the rate constant, A:obs, the final equation becomes,

ki2 = (ku k22 ^nfiiY'^ W12

where/is defined as. 2


W12-W21
lnKi2+-
RT
ln/12 =
1^11^22 ^11"^ W22
4 In y+4
(VetKA)ll(VetKA)22f RT

W12 = exp [- (wi2 + W21 - wii - W22) / 2 RT]

W12, W21 are the electrostatic free energy changes associated with the formation

of the products and reactants, respectively.

'Photochemical and Thermal Reactions ofSome Aggregated Systems'


Chapter 1 Introduction 15

1.4. Inner-sphere electron transfer reactions

The inner-sphere electron transfer occurs when one of the metal

centers (usually the reductant) is coordinated to a ligand on the other complex.

The reaction sequence consists of the following three steps.

1. Formation of the bridged (\i) complex - Precursor complex formation

M«U + XM'"'L'5 i L^Mn-X-M-niL's + L


^ Precursor complex

2. Activation of Precursor complex and electron transfer

L5M" -X-M'"^L'5 = ^ ^ L5M"^-X-M'"L'5

Precursor complex Successor complex

3. Dissociation to separated products

L5M™ - X - M'" L'5 J' - Products

Depending on which step controls the overall rate of reaction, the inner-sphere

mechanism can be classified into three types.

Type I: Reactions, in which precursor complexes are readily formed, and the

rate of electron transfer is slow, come under this type (i.e., kp > kgt), Here,

rate = ket [L5M" - X - M'^^L's]

'Photochemical and Thermal Reactions ofSome Aggregated Systems'


Chapter 1 Introduction 16

The inner-sphere reduction of ^r<3ra-[Co(en)2(H20)Cl]^^ by Fe(II) [104] is a

typical example. This reaction proceeds by rapid substitution on the labile

Fe(II) ion. The electron transfer and subsequent Co - CI bond rupture yield

Fed and Co as product. Similarly, in the Fe(II) reduction of the

nitrolotriacetate complex of pentaammine Co(III), Fe(II) forms the kinetically

stable binuclear precursor complex in a rapid equilibrium process followed by a

slow electron transfer to Co(III) [105].

Type II: Due to slow substitution on the reductant, precursor complex formation

is the rate-determining step, and electron transfer takes place immediately after

the formation of precursor complex, ket > kp

rate = kp [L5M" - X] [M'"^ L'5]

This behaviour is observed in the inner-sphere reduction by V(II) [106].

Type III: In this case, the reactants and successor complex are in equilibrium

and the rate of decay of the successor complex is rate determining step.

This happens when both metal centers are inert to substitution,

when, (kp/k _p) » 1 and (ket/k ^0 » 1

rate a [L5M'" - X - M'" L' 5]

when, (kp/k - p ) « 1 and (ket/k _et)« 1

ratea[L5M"-X][M'"'L'5]

when, (kp/k _p) » 1 and (k^/k ^ t ) « 1

ratea[L5M"-X-M'"'L'5]

'Photochemical and Thermal Reactions of Some Aggregated Systems'


Chapter 1 Introduction 17

The [Co(EDTA)]^" reduction of [Fe(CN)6]^" is an example of this type of

reaction [107]. Movius and Linck reported [108] that the Cr(II) reduction of

cis-[Ru(NH3)4Cl2]^ and ci5-[Ru(NH3)4(H20)Cl]^^ proceed by this mechanism.

Above all, changing the bridging group in an inner-sphere reaction may

alter the rate of electron transfer by changing the interaction energy, the stability

of the precursor complex as well as the standard free energy change for the

reaction. It would be meaningful therefore to discuss the various types of

bridging ligands and their effects on the electron transfer reaction mechanism.

1.5. Halide Bridging

A very important fact of the inner-sphere process is the bridging ligand

that forms part of the coordination spheres of both the oxidizing and the

reducing metal ions. The bridging ligand must function as a Lewis base toward

both metal centers; it must have two pairs of electrons that can be donated to

different metal centers [109].

Inner-sphere reactions are generally suggested by the following criteria:

1. When the rate of electron transfer is equal to or slower than the rates of

substitution of the reactants. This is a corollary of the requirement for outer-

sphere reactions, that a reaction must be outer-sphere if it takes place faster

than the rate of substitution at the metal centers. However, outer-sphere

reactions are found with labile reactants.

'Photochemical and Thermal Reactions of Some Aggregated Systems'


Chapter 1 Introduction 18

2. When the rates do not fit with the common tests for an outer-sphere

mechanism such as the Marcus relationship. In general the reactions must

be faster than predicted by the Marcus relationship. However, as has been

observed, outer-sphere reactions need not conform to the Marcus relationship

and a number of inner-sphere reactions show relationships similar to those

for outer-sphere reactions.

3. The presence of a suitable bridge on one of the reactants. Minimal

requirements that the bridging group must have an available lone pair of

sufficient basicity to coordinate to the labile coordination position of the

reaction partner.

Group 16 and 17 donor atoms (halogens, oxygen, sulphur) usually have

electron pairs available to form a bridge to a second metal ion, but Group 15

atoms (nitrogen, phosphorus, arsenic) have a single lone pair which is fully

occupied in bonding to metal. Nevertheless, an available lone pair may not be

essential requirement for inner-sphere electron transfer [109]. Thus, just as CH3

can act as bridge between F and 0H~ in the SN2 base hydrolysis, so it may act as

a bridge between Cr(II) and Co(III) in the reduction of methyl-Co(III) species by

[Cr(H20)6]^^. It is, obviously, necessary for the electronic structure of the

bridging ligand to be such that electron transmission from one donor site to the

other is possible. When there is no available bridging ligand, or no reasonable

'Photochemical and Thermal Reactions of Some Aggregated Systems'


Chapter 1 Introduction 19

through ligand electron transfer path, then electron transfer must take place by

outer-sphere mechanism. The equation,

A ^ X + B - > A + X . B ^ {d)

denotes an inner-sphere reaction with transfer of the bridging group from

oxidant to reductant; and the equation,

A \ X + B ^ A . X + B^ {e)

denotes the alternative outer-sphere process involving the same reactants.

As noted above, other things being equal, the rates of reaction tend to correlate

with the thermodynamic driving force. Thus, if the group X forms a particularly

stable complex with B^, in preference to A, a bridged mechanism will be

favoured; conversely, if in a series of suitably comparable systems, rapid rates

are observed whatever the bridging group specifically favours B^ rather than A,

this is evidence of a bridging mechanism.

It is well known that most metal ions exhibit a marked trend in affinity for

the halide ions [110], and the terms 'hard' and 'soft' have been used [111] to

denote cations which bond preferentially in the order F~ > CI" > Br~ > T and

vice versa. From equation {e), it is clear that for an outer-sphere mechanism, if

the oxidant A^ is 'harder' than the corresponding reduced form A, then rates

with a common reductant B will fall in the sequence I" > Br" > CI" > F".

Most metal ions are indeed 'harder' in the higher valencies than in the lower

[110], and the sequence just mentioned is the one commonly observed [112].

'Photochemical and Thermal Reactions of Some Aggregated Systems'


Chapter 1 Introduction 20

This can be understood by consideration of the orbitals involved in the

electron transfer process (Scheme 1). Along the unique z-axis, the electron is

transferred from a d^ orbital (a*) on Fe(II) to the c// orbital (CT*) in Co(III)

which is involved in the bonding to the bridging atom. The electron is

transferred as a result of the orbital overlap (resonance transfer), which is

modified by the presence of the bridge orbital. The energy of the electron donor

orbital is directly related to the strength of the Co-X binding interaction, and so

electron transfer will be strongly coupled to the Co-X strength. Further, the

group trans to the bridge will also have a major effect on the energy of the

orbital. Weakening the Co-trans ligand bond will favour electron transfer.

Strong field ligands raise the energy of the orbital and inhibit electron transfer;

weak field ligands enhance the electron transfer [109].

Co-X-Fe
Scheme 1

For Cr(II) reduction of [Co(NH3)5X]^^ complexes, transfer of X~ ligand

to the newly formed Cr(III), inner-sphere activated complex of the type

[(H20)5Cr(II)-X-Co"(NH3)5]'*^ are suggested [113]. Here the bridging efficiency

of halide ions I" > Br" > CI" > F~ for Cr(II) reduction deserves special mention

[86]. The same order is found with [Co(CN)5]^' [114] and may be termed the

'Photochemical and Thermal Reactions of Some Aggregated Systems'


Chapter 1 Introduction 21

'normal' order of bridging efficiency. In the reactions with Fe(II) [115,116] an

'inverted' order, F"> C r > Br~> T is observed. The normal order is found in

the reactions of Cr(II) with [Cr(NH3)5X]^'" [117] or CrX^"^ [118] and the inverted

order when Fe(II) reacts with Co(III) complexes [119]. The bridging efficiency

thus seems to depend primarily upon the nature of the reducing agent and to a

lesser extent on the nature of the oxidizing agent.

Halpem and Rabani [120] have suggested that for the reactions,

Co-X + red ^ Co + X-red

the reactivity order is rightly determined by the strength of the bond being

broken (Co-X) and the bond being formed (X-red). The bond strength in the

present cases increase in the order: Y > Br~ > 01" > F~. For highly reactive

reductants like Cr(II) bond-breaking is more important than bond-making and

the reactivity order should be [Co(NH3)5l]^^ > [Co(NH3)5Br]^^ > [Co(NH3)5Cl]^^

> [Co(NH3)5F]^'". For reductants of low reactivity like Fe(II) [116, 119], bond

making with the reductant is of more importance because it stabilizes the

transition state of the reactants and hence the reactivity order is influenced by

the strength of the bond being formed, i.e., F" > CI" > Br" > F. The observed

reactivity orders are consistent with this interpretation. Sutin [9], Diebler and

Taube [115] suggested that electron transfer may proceed by means of

CT-interaction when Cr(II) is used as a reducing agent and for Fe(II) reduction,

5i-interaction may be more important than a-interaction.

'Photochemical and Thermal Reactions ofSome Aggregated Systems'


Chapter 1 Introduction 22

1.6. Effect of Non-Bridging Ligand

The ligands coordinated to the oxidant or reductant, but not involved in

bridging the reactants are defined as non-bridging ligands. Since the role of

non-bridging ligands appears to be electronic in nature, their effect on the

reduction rate should be similar in inner-sphere and outer-sphere reactions.

For the inner-sphere reactions much of the early work on non-bridging effects

was prompted by Orgel's theory [121] and the ready availability of numerous

geometric isomers of Co(III) and Cr(III) ammine and amine complexes.

For inner-sphere reactions in which the d^ orbital accepts the reducing electron,

the theory predicts that the rate of electron transfer is inversely proportional to

the ligand field strength of the group trans to the bridging ligand. Ogard and

Taube [117] extended these observations to include stretching of the trans

metal-ligand bond to achieve energy matching between donor and acceptor

orbitals. The first systematic test of these ideas was reported by Benson and

Hahn [122]. They found that the rate of oxidation of Fe(II) by some Co(ni)

complexes containing chloride as the common ligand showed tremendous

sensitivity to a change in non-bridging ligand.

Following this, numerous reports have appeared on non-bridging effects.

These include the Fe(II) reduction of c/5-[Co(en)2(NH3)Cl]^^ and

cw-[Co(en)2pyCl]^^ [123], Fe(II) reduction of c/5-[Co(en)2(A)Br]^^ [124]

'Photochemical and Thermal Reactions of Some Aggregated Systems'


Chapter 1 Introduction 23

(A = NH3, H2O), V(II) reduction of cw-[Co(en)2(A)Cl]"^ (A = CI", H2O) [125],

Cr(II) exchange reaction of cw-[Cr(en)2(A)Cl]"^ (A = CI", H2O) [126].

Bifano and Linck [123, 127] suggested that the sigma-bonding strength of

a ligand, rather than its position in the spectrochemical series, should be a useful

criterion of efficiency as a non-bridging ligand. This hypothesis was tested by

reduction of a series of cw-[Co(en)2ACl] complexes, where A refers to one of

a series of amines of varying base strength. Although the reduction with Fe(II)

indicated a dependence of rate on pKb of the ligand, reduction of a more


— '5-4-

extensive series of complexes by Ru(NH3)6 indicated that other factors were

also involved.

It would be interesting to study the effect of non-bridging ligand with

[Ru(NH3)6]^^, an outer-sphere reducing agent. To check whether the effect of

non-bridging ligands is seen only in the case of inner-sphere mechanism or can

be observed even for outer-sphere mechanism, Patel and Endicott [128] studied

the reduction of a series of cw-[Co(en)2(A)Cl]"*, where A = NH3, py, etc., by

[Ru(NH3)6]^^. They observed that the specific rates of outer-sphere

[Ru(NH3)6]^^, and inner-sphere Fe(II) reductions exhibit nearly the same

sensitivity to the ligation of the Co(III) oxidant [123]. This similarity in the

relative rates of reduction of cw-[Co(en)2ACl]^'^ complexes suggests that the

specific rates of these electron transfer reactions might be factorable into terms

which involve the properties of the oxidizing agent and the reducing agent

'Photochemical and Thermal Reactions of Some Aggregated Systems'


Chapter 1 Introduction 24

separately independent of mechanisms (i.e., whether outer-sphere or inner-

sphere);
i.e., ki2 «/n(i) (cobalt):yn(2)(reductant)

This is the approximate form of the free energy relationship suggested by

Marcus for outer-sphere reactions and this relationship has found some

application to inner-sphere as well [128].

Hicks, Toppen and Linck [129] reported values for the reduction of

Co(III) complexes containing azide ligand by V(II) by an inner-sphere path as

long as the non-bridging ligands are not too rate enhancing. Another interesting

study [130] involves the reduction of superoxide bound to two Co(III) ions by

Fe(II).

4+ 3+
^NH2^ ^NH,^
2+ 3+
Fe + LsCq Col^ •^ Fe + L2Cq C0L2
^0-0^ ^oV

As L is changed from en to bipy (or phen), the rate increases significantly;

supporting the view that 71-bonded ligands bipy and phen are rate enhancing.

1.7. Orbital symmetry and Electron transfer

The probability of electron transfer increases with increase in interaction

energy and when orbitals of same symmetry are involved in the reaction, the

interaction energy is large. If the ligands have filled or empty 71-orbitals (pTC, d7c.

'Photochemical and Thermal Reactions of Some Aggregated Systems'


Chapter 1 Introduction 25

or 7c-system), those orbitals will interact with the d orbitals of the metal ion

having tag symmetry. The d orbitals of the central metal ion having Cg symmetry

will overlap with the ligand orbitals having a symmetry. Evidently electron

transfer between tag orbitals will be favoured by 7i-bridging orbitals as in the

case of [Fe(CN)6]'*~' ^~, while electron transfer between Cg orbitals will be

favoured by a bridging ligand, for example, exchange reaction between

[Cr(H20)6]'" and [Cr(H20)5X]'".

When the Cg reductants, Cr(II) [84, 86] and [Co(CN)5]^" [114] react with

eg oxidants, the reactivity order for X is CI" > N3" » CH3CO2". The Cr(II)

reduction of [Fe(H20)5X]"^ indicates Na" > CI", but they react near the

substitution limit for Cr(II). Fe(II) is thought to react by inner-sphere path and

shows for Cg oxidants, Na" > CI" [116, 131]. The reduction by V(II) of

[Co(NH3)5N3]^^ is thought to be by an inner-sphere path [129] and that of

[Co(NH3)5Cl]^^ is outer-sphere [132, 133].

The ability of a ligand to mediate an electron transfer has been ascribed to

a matching of the symmetry of metal and ligand orbitals. When the symmetry

of the orbitals of the metal ions that donate and accept the electron are the same,

ligands with orbitals of matching symmetry may provide a lower energy

pathway for electron transfer. Thus if the reductant donates an Cg electron to an

Cg orbital, a ligand such as chloride (a sigma carrier) would be a better bridging

'Photochemical and Thermal Reactions of Some Aggregated Systems'


Chapter 1 Introduction 26

ligand than azide or acetate. Electron transfer involving a tig orbital would be

better accommodated by an azide or an acetate, presumably because of better

overlap of the t2g orbital with the 71 system of azide or acetate. Existing data are

in agreement with these suggestions, but they are insufficient to test the concepts

adequately [15].

1.8. Reducing Metal Ions

Lower valence transition metal ions act as powerful reductants.

The reducing efficiency and mechanism of the reduction depends largely on the

nature of the metal ion reductant and ligands surrounding it. A brief account on

the most well studied reducing metal ions would be pertinent here.

1.8.1. Iron(II)

The reduction of Co(III) complexes by Fe(II) continues to be an active

area of redox studies [134-150]. The redox potential for Fe /Fe couple is

+ 0.74 V [134]. Wada and co-workers have reported the effect of CH3OH [131]

and DMSO [135] on the Fe(II) and Fe(III) electron-exchange reaction.

They rationalize their results in terms of the hydrogen-atom transfer mechanism.

Ohashi and coworkers [136] have reported an increase in rate as the mole

fi-action of ethanol is increased for the Fe(II) reduction of

cw-[Co(en)2(NH2CH2CH20H)Cl]^^. Fe(II) reduction of several CoA4BX^

complexes where A is NH3, V2 en, Vabipy, V2 phen, etc., and B is a base have

been reported [137].

'Photochemical and Thermal Reactions of Some Aggregated Systems'


Chapter 1 Introduction 27

Kurimura et al. [138-143] reported the Fe(II) reduction [138] of

[Co(EDTAH)Cl]", the corresponding Br" complex and related complexes with

modification of the EDTA moiety as ligand, the reduction of macromolecular

Co(III) complexes by low molecular weight Fe(II) chelates [139-142] and

polymer-bound Fe(II) chelates [143]. Krop et al. [144] proposed a bridged

inner-sphere mechanism, for the Fe(II) reduction of halogeno macrocyclic

Co(III) complexes containing thioether donor atoms.

Candlin and Halpem [145] examined the volume of activation for a

number of reactions with Co(III) complexes, as a possible means of

distinguishing between inner-sphere and outer-sphere mechanism.

More positive A V values have been observed for the inner-sphere process.

Davies [146] reported the Fe(II) reduction of Co(III) in both LiC104-HC104 and

NaC104-HC104 mixtures, van Eldik et al. [147, 148] made use of volume of

activation as a criterion for assigning the nature of the mechanism involving

Fe(II) complexes [147] and Fe(II) [148] as the reductants. The effect of pressure

[149] and the role of distance on intramolecular electron transfer rates between

Co(III) and Fe(II) have also been studied in detail [150]. The thermal and

photoinduced inner-sphere electron transfer reactions of [Fe(CN)5H20] " and

[Co(NH3)4(pzc)]^"', [Co(en)2(pzc)]^'^, and [Co(NH3)4pz]^^ have been measured as

a function of temperature and pressure [151]. The large positive volume of

activation for the thermal process (AV* = 27-37 cm^ mol"') decrease

'Photochemical and Thermal Reactions of Some Aggregated Systems'


Chapter 1 Introduction 28

dramatically for the photoinduced intramolecular pathway (3-11 cm mol" ),

which is interpreted as a two-step chemical mechanism. The rate constants for

the inner-sphere electron transfer in the [(NC)5Fe(|i-pzc)Co(NH3)4]~ complex

decrease with an increase in the electrolyte concentration [152]. This trend has

been attributed primarily to a decrease in the thermodynamic driving force of the

redox reaction.

Two lipophilic redox couples, [Fe(4,4'-(t-bu)2bpy)3]^"^''^"^ [153] and

[Co(N-Ci6H33[14]aneN4)Cl2]°^^ [154], have been used to transport electrons

across a CH2CI2 liquid membrane. The former system employed aqueous

Ce(IV) as the oxidizing phase, with the reducing phase involving a number of

reductants. The rate of electron transport was controlled by the slower of the

two-redox processes. In the latter system, electrons were transported between

Cr(II) (aq) and Fe(III) (aq) in aqueous HCl phases. The electron transport is

stopped when CI" is replaced by CIO4', owing to the high reduction potential of

the [Co(N-Ci6H33[14]aneN4)(C104)2]^ species. The catalysis of Fe(II) - Co(III)

complex reactions by poly(vinylsulphonate) has been reported [155].

An inner-sphere halide bridged mechanism is proposed [156] for the

Fe(II) reduction of /ram-[Co(DH)2pyX] where DH~ is dimethylglyoximate

anion and X is the halogen ion. Rate constants for both uncatalyzed and base-

catalyzed pathways show an increase in the order CI" < Br" < F. A similar

'Photochemical and Thermal Reactions of Some Aggregated Systems'


Chapter 1 Introduction 29

mechanism is proposed [157] for the trans-azido and /ram-thiocyanato

complexes.

1.8.2. Ruthenium(II)

The redox potential for the [Ru(H20)6]^^^^^, [Ru(NH3)6]^'^^^^,

[Ru(en)3]^^^^^, [Ru(bpy)3]^^^^^ couples are + 0.23 V, + 0.067 V, + 0.19 V,

+ 1.29 V respectively [134].

Endicott and Taube [85] studied the reaction of [Ru(NH3)6]^^ with

pentaammine and tetraammine Co(III) complexes. Since [Ru(NH3)6]^^

undergoes substitution very slowly compared to the rate at which it is oxidized,

these reactions are undoubtedly of the outer-sphere type.

Reduction of a number of [(NH3)5CoX] complexes, where X is a halide,

by [Ru(en)3]^'', [Ru(NH3)6]^^ and [Ru(NH3)5H20]^^ have been reported [158].

For the first two reductants, an outer-sphere mechanism is proposed.

For [Ru(NH3)5H20]^^ however, reductions of [Co(NH3)5F]^^ and [Co(NH3)5Cl]^^

are irmer-sphere involving substitution at the ruthenium center. A feature of

Ru(II) chemistry is the case with which Ru(II) complexes reduce C104~ to CIO3"

[159]. Intramolecular electron transfer from Ru(II) to binuclear Co(III) centers

has also been reported [160]. The rates are dependent on the Co(III)/(II)

reduction potential but insensitive to the bridging ligand, indicating an adiabatic

process.

'Photochemical and Thermal Reactions of Some Aggregated Systems'


Chapter 1 Introduction 30

Controversy continues concerning the mechanisms of quenching of

[ Ru(bpy)3] and its derivatives by Co(III) complexes. Energy transfer

competes with electron transfer in the reaction with Co(III) cage complexes

[161]. An electron transfer mechanism for quenching of excited-state ruthenium

polypyridine complexes by Co(III) complexes is deduced [162] from rate

comparisons with the corresponding reductions by [Ru(NH3)6]^^. However,

universal assumption of an electron transfer mechanism has been criticized

[163] for these reactions and it has-been shown that with [Co(en)3] energy

transfer predominates, while with [Co(NH3)6] energy transfer is competitive

with the electron transfer process. Part of the evidence is that quenching by

[Co(en)3] is faster than with [Co(NH3)6] , the inverse of the order expected if

the electron transfer were the sole mechanism. Electron transfer is the operating

mechanism with [Co(NH3)5Cl] as quencher [164, 165]. Studies with

[Co(ox)3] ~ have also been reported [166]. Ascorbate reduction [167] of

[*Ru(bipy)3]^"^ gives [Ru(bipy)3]"^, which can be used to reduce [Co(phen)3]^^ to

the blue [Co(phen)n]^ ion. This ion reacts in aqueous medium to give a hydride

intermediate in the generation of H2 gas. A large number of papers dealing with

water-splitting reactions involving mainly the effects of catalysts on

[*Ru(bipy)3]^^ / (MV^"") system have appeared [168-174].

Long-range electron transfer from photo excited [Ru(phen)3] to

tris(polypyridine) complexes of Co(III), Rh(III), and Cr(III) is mediated by

'Photochemical and Thermal Reactions of Some Aggregated Systems'


Chapter 1 Introduction 31

DNA [175]. The DNA polymer provides an efficient intervening medium for

donor-acceptor coupling, with surface-bound complexes displaying greater rate

enhancements than intercalatively bound species.

1.8.3. Chromium(II)

Cr(II) is the most powerful reducing agent of all the available reductants.

The standard electrode potential for the couple [134]

J+ .2+ : „ - c O _
Cr'" + e-->Cr"" isE" = - 0 . 4 0 V

Normally Cr(II) reduction proceeds by an inner-sphere mechanism, but,

the reaction of Cr(II) with [Co(NH3)]6 is believed to be outer-sphere, because

there is no [H^]~^ dependence [176]. This has been attributed to the absence of

any basic ligand, which can provide a pair of electron for coordination with the

reductant.

Cr(II) reduction of crs-[Cr(N3)2(H20)4]^, c/s-[Co(NH3)5(N3)]^* and

cw-[Co(en)2(OOCH)2] indicated the transfer of two ligands from oxidant to

reductant. A double-bridged activated state was proposed to account for the

product distribution [177-179].


H
* I
NNN^ / O-C =o
Cr Cr (en)2Co
\
o-c = o
H
Lewis and coworkers [180-183] also made a similar observation on the Cr(II)

reduction of [Co(bipy)2(acac)]^'^, [Co(acac)3] and [Co(bipy)3]^"^. Neves and

'Photochemical and Thermal Reactions ofSome Aggregated Systems'


Chapter 1 Introduction 32

Franco [184] studied the electron transfer kinetics of Cr(II) with tetra-nuclear

isomeric species of Co(III) complexes containing pyridine 2,6-, 3,5-, 2,4- and

2,5-dicarboxylate bridging ligands. The 2,6-isomer reacted by an outer-sphere

path; whereas, the other three isomers reacted by an inner-sphere path with

attack of Cr(II) at the pyridine nitrogen.

Candlin et al. [84] compared kinetics data for the reduction of a series of

pentaammine Co(III) complexes with V(II), Cr(II), Eu(II) and [Cr(bipy)3]^''.

The variation in rate with V(II) and [Cr(bipy)3] was found to be much smaller

than with Cr(II) and the former two show similarity in variation of the rates and

also [Cr(bipy)3]'^'^ has no free coordination site unlike Cr(II) for substitution,

which clearly suggests [Cr(bipy)3] reacts by outer-sphere path. In contrast

the reactivity of Eu(II) towards the halopentaammine Co(III) complexes was

found to decrease [Co(NH3)5F]^^ > [Co(NH3)5Cl]^^ > [Co(NH3)5Br]^^ >

[Co(NH3)5l]^^, a trend which has been interpreted in terms of inner-sphere

mechanisms and a similar trend is observed with Fe(II) also.

In the Cr(II) reduction of [(NH3)5Co02CCOC6H5]^^ a chemical

mechanism involving the formation of an intermediate Co(III)-Cr(III) bound

phenylglyoxylate radical species is proposed [185]. The spectrophotometrically

detected intermediate is formed rapidly and decays with a first-order rate

constant of 93 s'^ in IM HCIO4. Similar observations have been made with the

binuclear oxidant [(NH3)3Co(^-OH)2(^-02CCOC6H5)Co(NH3)3]^^ where decay

'Photochemical and Thermal Reactions of Some Aggregated Systems'


Chapter 1 Introduction 33

is slower, 4 s"\ and with [(H20)5Cr02CCOC6H5]^^ as oxidant. In the latter case,

the intermediate reacts with excess Cr(II) and may be bleached catalytically by

the free keto-acid formed by aquation of the reactants. Detection [186] of O-

bonded Cr(III) amino acid products in the Cr^^ reduction of N,0-chelated amino

acid complexes of Co(III) suggests that an inner-sphere mechanism is operating

in contrast to an earlier report [187]. The rates are faster than those of

[(NH3)5Co(0-aminoacidH)] complexes due to greater accessibility of the

carboxylate group on chelation and lowering of the charge.

Reduction of [Co(EDTA)]~ by Cr(II) shows [188] a reactivity pattern

corresponding to other inner-sphere reactions of this oxidant and its derivatives

[189]. The product is a carboxylate-bound EDTA complex of Cr(III).

Nitrogen-bonded Cr(III) products are detected [190] in the reductions of a

number of pentaammine Co(III) tetrazole complexes by Cr(II), The rate laws

show inverse [H^] dependences and rate constants consistent with an inner-

sphere mechanism in all cases.

1.8.4. Vanadium(II)

The reactivity of the reducing agent, V(II) is clearly understood from its

redox potential [134], viz.,

V^^ + e'->V^'' isE^ = - 0 . 2 6 V

Since V(II) is substitution-inert, substitution into the V(II) coordination

sphere determines the nature of the mechanism. If the rate constant is with in

'Photochemical and Thermal Reactions ofSome Aggregated Systems'


Chapter 1 Introduction 34

40 M"' s~\ then the reaction mechanism [9] is classified as substitution-

controlled inner-sphere. A large number of V(II) reductions come under this

class [84, 98, 132]. Since V(III) is substitution-labile, product analysis does not

help in predicting the nature of the activated complex and so indirect criteria like

AH* and AS** have to be considered for assigning the mechanism. However,

Hicks et al. [129] reported the V(II) reduction of [Co(NH3)5N3]^'", where in they
0-4-

observed the formation of VN3 following an inner-sphere mechanism.

Several other reactions of V(II) are sufficiently rapid and are unambiguously

classified as outer-sphere [191-196].

1.8.5. Titaniuiii(III)

A variety of mechanistic pathways have been observed in reduction of

metal complexes by Ti(III). Fraser and co-workers [197] were the first to use

Ti(III) as a reducing agent. They reported the reduction of halopentaammine

Co(III) complexes by Ti(III). Following this, Orhanovic and Earley [198]

reported the reduction of [Co(NH3)5Cl] and cis- and ^a«5-[Co(en)2Cl2] by

Ti(III). Birk [199], Bakac and Orhanovic [200] reported the Ti(III) reduction of

azido, isothiocyanato and thiocyanato pentaammine Co(III) complexes.

They clearly discussed the differentiation between two more detailed

inner-sphere mechanisms operating for the reduction of the thiocyanate and

azido complexes.

'Photochemical and Thermal Reactions of Some Aggregated Systems'


Chapter 1 Introduction 35

An important feature in the case of Ti(III) reduction is the observation of

Ti(III) hydrolysis equilibrium. The conjugate base TiOH^"^ reacts faster than the

acid form Ti . Bakac, Marcec and Orhanovic [201] in order to find out a

reliable value of hydrolysis constant kh for Ti^^-TiOH^^ equilibrium, studied the

Ti(III) reduction of [Co(phen)3] and [Co(terpy)2] and observed an outer-

sphere mechanism, the value of kh being 3.54 x 10' M. Martin and Gould

[202] reported [H^]"^ dependence for the Ti(III) reduction of oxalato

pentaammine Co(III) and proposed a binuclear activated complex in which

TiOH forms a chelate with oxalatopentaammine Co(III) complex.

The kinetics of the reductions of three pentacyanocobaltate(III)

complexes, [Co(CN)5X]^- where X = N3", SCN", and NCS", by Ti(III) in acidic

aqueous solution display inverse acid rate dependences. Using rate-constant

ratios, the reactions are suggested [203] to proceed by inner-sphere mechanisms

involving substitution on TiOH^"^. The kinetics of the reductions of

[Ru2(CH3C02)4]^ and [€0(0204)3]^' by Ti(III) oxalato complexes have been

studied in aqueous solution. The reactive Ti(III) species is [Ti(C204)2]~, with

the former reaction proceeding by an outer-sphere mechanism, while the latter

reaction proceeds by an irmer-sphere pathway involving substitution on the

reductant [204].

Acid dependencies in the reduction of [(NH3)5CoLH]^^ complexes, where

LH is a keto-substituted carboxylic acid, by Ti(III) reveal much about the

'Photochemical and Thermal Reactions of Some Aggregated Systems'


Chapter 1 Introduction 36

mechanism [205]. Remote keto substitution, in the P and y positions, results in

[H^]~* terms in the rate law with rates comparable with those found for

unsubstituted derivatives, while additional [H^]~^ terms with a-keto substituents

and enhanced rates suggest attack by TiOH^^ on the gem-diol hydrated form of

the ligand. With phenyl-substituted keto acids, such as LH = C6H5COCO2H, an

acid-independent term reflects precursor-complex formation by reaction of

[Ti(H20)6]^''. Reduction of [Co(NH3)5(Hnta)]^ by Ti(III) also involves [206]

the formation of a binuclear intermediate, detected spectrophotometrically, with

a stability constant of lO''^^. A brief report of the reduction of [Co(NH3)4C204]'^

by Ti(III), [Ti(C204)]^, and [Ti(C204)]' has appeared [207] in an effort to

establish the role of bridging oxalate. The reactions are inner sphere and have

rates in the order 0.04 M"' s"*, 10 JVT^ s"^ and 400 M"' s"', respectively, in

0.15 M [it] at 1.0 M ionic strength.

1.9. Mechanistatic details of inner-sphere reactions

Most of the iimer-sphere reactions that have been studied kinetically

correspond to the case where precursor formation is rapid and unfavourable, and

electron transfer within the precursor complex is rate determining.

The measured second-order rate constant is then equal to Qpket where Qp is the

equilibrium constant for the formation of the precursor complex and ket is the

rate constant for inter-molecular electron transfer. Under such circumstance, the

role of the bridging ligand is dual [15, 208]. It brings the metal ions together

'Photochemical and Thermal Reactions of Some Aggregated Systems'


Chapter 1 Introduction 11

(thermodynamic contribution) and mediates the transfer of the electron (kinetic

contribution).

The rate constant for electron transfer between two metal ions may be

analyzed in terms of an electronic factor and a nuclear factor. The potential

energy surfaces of the precursor and successor complexes are functions of the

respective nuclear configurations. The precursor and successor complexes have

different sizes and charge distributions and therefore their most stable

configurations occur in different regions [209, 210]. The intramolecular

electron transfer process that transforms precursor into successor complex

corresponds to the crossing of the system from one surface to the other.

The first order surfaces for the system are shown schematically in Figure 4.

The nuclear factor represents the change in nuclear positions, both inner-sphere

(metal-ligand distance) and outer-sphere (orientation of solvent dipoles)

necessary to bring the precursor complex from its equilibrium configuration to

that appropriate to the intersection region. At that point, the energies of the

system before and after electron transfer are equal. Whether elecfron transfer

obtains (electronic factor) when the intersection region has been reached

depends on the interaction energy H12. If H12 is sufficiently large (0.5 kcal),

every time the required nuclear configuration is reached, the electron is

transferred, that is, the system follows the path given by the lower surface.

This process is referred to as adiabatic electron transfer. If H12 is too small, the

'Photochemical and Thermal Reactions ofSome Aggregated Systems'


Chapter 1 Introduction 38

system upon reaching the required configuration tends to follow the Hn surface

and only occasionally stays in the lower surface. This process is referred to as

non-adiabatic electron transfer. The magnitude of H12 depends on the distance

between the metal ions and on the symmetry of the donor, acceptor and carrier

orbitals. Electron transfer between metal centers that use the bridging ligand for

coupling the metal orbitals corresponds to the resonance transfer mechanism.

Under some circumstances, the bridging ligand can become involved chemically

in the reaction sequence. The electron (or hole) is transferred to the bridging

ligand, and the process results in the formation of a chemical intermediate of a

finite lifetime. In a subsequent step, the reduced (or oxidized) bridging ligand

transfers the electron (or hole) to the final acceptor and the overall electron

transfer is consummated. This is referred to as chemical, radical or stepwise

mechanism.

Figure 4. Schematic representation of zero andfirst-ordersurface for electron


transfer. E = Energy; q = nuclear coordinate; P = precursor complex;
B = transition state for electron transfer within binuclear complex;
S = successor complex; Hn = precursor complex surface; H22 = successor
complex surface; H12 = splitting at intersection region.

'Photochemical and Thermal Reactions of Some Aggregated Systems'


Chapter 1 Introduction 39

1.10. Surfactants and Metallosurfactants

Surfactants, sometimes called surface-active agents or detergents, are

among the most versatile chemicals available. They are amphiphilic molecules

consisting of a hydrophilic head group and a hydrophobic (lipophilic) tail, and

are thus able to interact with both polar and non-polar compounds. Amphiphilic

molecules self-assembled in aqueous solutions were introduced ninety years ago

by McBain [211]. He observed unusual changes in electrical conductivity upon

changing the soap concentration and coined the name micelle for the aggregates.

Micelles are one of the colloidal association structures, in which vesicles,

microemulsions and liquid crystals are included. Micelles are formed over a

fairly narrow concentration range of surfactant and this range is called critical

micelle concentration (CMC). They are molecular or ionic aggregates whose

size is about 3-100 nm and which arise after the attainment of the CMC.

Amphiphilic molecules exist as the monomeric form at low

concentrations, and they tend to aggregate (micelle) with increasing

concentration, although micelle formation is dependent not only on their

hydrophobic and hydrophilic balance of molecules, but also on the environment

where the surfactant exists. In a specific concentration range above the CMC

(usually within the limits of variation of the concentration by 1-2 order of

magnitude) micelles are spherical in shape. This is the lowest concentration at

which spherical micelles forms [212]. The formation of micelles was often

'Photochemical and Thermal Reactions of Some Aggregated Systems'


Chapter 1 Introduction 40

characterized by a discontinuity in system properties, such as conductivity,

surface tension, light scattering, self-diffusion and the molality of dissolved

compounds. Experimentally, the CMC can be determined from the inflection

plots of some physical property of the solution as a function of concentration.

The aggregation and surface properties of a surfactant in solution are very

sensitive, and are influenced or controlled by the solvent polarity, temperature,

pressure, pH, and the presence of various foreign substances (cosolvent).

The CMC can serve as a measure of micelle stability in a given state and the

thermodynamics of micellization can be determined from a study of the

CMC-temperature dependence [213-215].

Surfactants are important because of their bacteriostatic properties.

They have been introduced into several commercial products such as antiseptic

agents in cosmetics and as germicides [213]. They have also found a wide range

of applications because of their unique solution properties such as detergency,

solublization and surface wetting capabilities in diverse areas such as mining,

petroleum, and pharmaceutical industries, chemical as well as biochemical

research [214] and as catalyst in several organic and inorganic reactions [215].

Micellar effects on electron transfer reactions of metal ion complexes

attracted the attention of a large number of workers due to their importance in

biological processes. It has been observed [216-220] that several redox

reactions in micellar media are influenced by the hydrophobic and electrostatic

'Photochemical and Thermal Reactions of Some Aggregated Systems'


Chapter 1 Introduction 41

forces and for a given set of reaction the observed rate depends on the extent of

association between the reactants and micellar aggregates. Most of the chemical

reactions investigated are of simple organic and inorganic systems.

On the otherhand, relatively few works have been carried on the electron

transfer reactions for transition metal ions complexes with lipophilic ligands

[76, 221-227]. We are interested in the synthesis and micelle forming properties

of many metal complexes containing lipophilic ligands (metallosurfactants) for a

long-time [228-237].

Metallosurfactants are a special type of surfactants, where a coordination

complex (containing a central metal ion with surrounded ligands coordinated to

the metal) acts as the surfactant (Scheme 2).

hydrophilic head group

t
o4=o
M"' : Co(III) ^ V '
hydrophobic tail group
hydrophobic tail group

Metallosurfactants Sodium dodecylsulphate


Scheme 2

In these surfactants, the metal complex part containing the central metal ion with

its primary coordination sphere acts as the head group and the hydrophobic part

of one or more ligands acts as tail part. Like any other well-known surfactants

for example, sodium dodecyl sulphate (SDS), these surfactant-metal complexes

'Photochemical and Thermal Reactions of Some Aggregated Systems'


Chapter 1 Introduction 42

also form micelles at a specific concentration called critical micelle

concentration (CMC) in aqueous solution [228, 232]. There are but a few

reports [238-243] on the synthesis, isolation and characterization of surfactant

transition metal complexes, in contrast to numerous reports of the formation and

study of such surfactants in solution without isolation. It is argued that the high

charge and size of the head group of the complex having long paraffin tails;

detergent-like characteristics are able to penetrate biological membranes and

destabilize the exterior membrane of the organism [76]. Moreover, transition

metal ions complexes with lipophilic ligands in micellar or vesicular aggregates

attracted considerable attention as catalyst for the cleavage of esters and amides

and as biomimetic models of hydrolytic metalloenzymes [244-246]. Recently

there has been increasing interest in the use of these organized media to study

the fundamental photochemical reactions of metal complexes in relevance to the

conversion of solar energy into other useful forms of energy through

photochemical reactions [52, 54], So it is reasonable to examine

metallosurfactants in this content.

'Photochemical and Thermal Reactions ofSome Aggregated Systems'

You might also like