0% found this document useful (0 votes)
92 views57 pages

Transition and Inner Transition Metal Chemistry: CML 524

Transition metal complexes can form metal-metal bonds of varying orders. Rhenium forms an unusual quadruple bond between two rhenium atoms in the [Re2Cl8]2- complex, with bond lengths of only 2.24 angstroms. This involves the overlap of d-orbitals to form sigma, pi, and delta bonds between the metals. The delta bond, though weak, further stabilizes the short and highly stable quadruple bond in this complex.

Uploaded by

Rahul Malik
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
92 views57 pages

Transition and Inner Transition Metal Chemistry: CML 524

Transition metal complexes can form metal-metal bonds of varying orders. Rhenium forms an unusual quadruple bond between two rhenium atoms in the [Re2Cl8]2- complex, with bond lengths of only 2.24 angstroms. This involves the overlap of d-orbitals to form sigma, pi, and delta bonds between the metals. The delta bond, though weak, further stabilizes the short and highly stable quadruple bond in this complex.

Uploaded by

Rahul Malik
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 57

Transition and Inner Transition Metal Chemistry:

CML 524

• Metal-metal bonded Reactivity of Introduction to


compounds and Transition metal Bioinorganic
Chemistry
transition metal Complexes:
• Bioinorganic
cluster compounds • Substitution chemistry of iron:
• Uses of lanthanide Reactions hemoglobin,
complexes: as shift • Electron transfer myoglobin,
reagents, as strong reactions cytochromes
magnets, and in • Photochemical • Non-heme Iron
Enzymes
fluorescence reactions of
• Bioinorganic
• Use of transition transition metal chemistry of zinc
metal complexes in complexes and cobalt
catalysis • Copper dependent
enzymes

Book: Sharpe and Houscroft, Inorganic Chemistry, 4th Edition


J. E. House, Inorganic Chemistry, 3rd Edition
Metallic Bond

In metals, the bonding is essentially covalent. The bonding


electrons are delocalized over the whole crystal, giving rise to
the high electrical conductivity that is characteristic of metals.

➢ Solid state packing of atoms in metal

(left) a cubic close-packed (face-centred cubic) lattice and


(right) a hexagonal close-packed lattice

The metallic radius is half of the distance between the


nearest neighbor atoms in a solid state metal lattice, and is
dependent upon coordination number.
Metal-Metal bond(s) in compounds

Zr4+

Monomeric [ZrF7]3-

➢ Transition elements (especially 4d and 5d) can form Bu3SnH


multiple bonds among themselves in cluster complexes. PR3 Low valent
➢ Compounds with the metal atoms in low to medium ZrCl4
state of Zr
positive oxidation states, and ligands of the same kinds
normally found in classical Werner complexes, e.g.,
halide, sulfate, phosphate, carboxylate or thiocyanate
Zr-Zr : 320–354 pm
ions, water, amines and phosphines. Compounds of this [Zr6Cl14(PnPr3)4]
(1 pm = 0.01 Å)
type include metal-metal bonds of orders ranging from
about 1/2 to 4.0.
The structure (determined by X-ray diffraction) of
[Zr6Cl14(PnPr3)4] [F.A. Cotton et al. (1992) Angew. Chem. Int.
Ed., vol. 31, p. 1050] Colour code: Zr, yellow; Cl, green; P,
red; Me and nPr groups are omitted.
Those compounds with the metal atoms in formal oxidation states of zero or close to it,
including negative ones. For the most part these are polynuclear metal carbonyls, or very
similar compounds. In these compounds the M-M bonds are usually long, weak and of order
one.

➢ Second and third row


transition elements can
form multiple bonds
among themselves in
cluster complexes.

According to F. A. Cotton, “I shall define as first order (or ipso facto)


evidence that a metal−metal bond exists, an internuclear distance between the
metal atoms that is equal to or less than the sum of their estimated single
bond radii.”
History of M-M bond congaing compounds

Cr2(O2CCH3)4(H2O)2 Ta6Cl4 · 7H2O K3W2Cl9


1844 1913 1935

[Re3Cl12]3−
ReI−ReII = ReI−ReIIʹ = 2.47 (0.01) Å
Re3Cl9(PR3)3
F. A. Cotton et al., J. Am. Chem. Soc., 1963, 85, 1349
According to Cotton “We viewed this
extraordinary structure with misgivings.
The state of refinement was poor and
uncertain. The presence of isolated protons
‘on centres of symmetry’ was
unaccountable.”

"the dimeric group [Re2Cl8]4−. Eight chlorine atoms


constitute a square prism ...the rhenium atom has for its
neighbours one rhenium atom, at a distance of 2.22 Å,
and four chlorine atoms at a distance of 2.43 Å. It may
be surmised that four hydrogen atoms are situated
between Cl atoms on centres of symmetry…and serve to
bond the [Re2Cl8]4− groups…to each other.”
P. A. Koz’min et al, Zhur. strukt. Khim., 1963, 4, 55 F. A. Cotton et al, Inorg. Chem., 1965, 4, 330
Quadruple bond
• A quadruple bond can only occur with transition metals, and
orbitals of angular momentum 2 or more are required for such
bonding.

• The overlap of d orbitals can form:


σ bond (dz2 + dz2) and σ* (dz2 − dz2)
𝜋 bond (dxz + dxz or dyz + dyz) and
𝜋* (dxz − dxz or dyz − dyz)
δ bond (dxy + dxy) and δ* (dxy + dxy)

Energy of the orbitals: σ < 𝜋 << δ < δ* << 𝜋* < σ*

Diagram of the overlaps of d-orbitals and the resulting energy


levels as they are involved in the formation of M–M multiple bonds
in a X4M–MX4 structure.
MO presentation of Re-Re multiple bond

𝜎*
𝝅* [Re2Cl8]2−
δ*

σ bond from 𝜋 bond from


overlap of dz2 overlap of dxz
orbitals orbitals
δ

𝝅 Re−Re distance: 2.24 Å


Orbital configuration:
𝜎 σ2 𝜋4 δ2

δ bond from overlap


of dxy orbitals
Stability of M-M quadruple bonds

Bond order is being used in an ordinal and not a metrical sense; it is simply a statement of the net number of
electron pairs—or halves thereof—that are serving to bind the two atoms together. It does not explicitly or
implicitly provide a measure of bond strength, except in the broadest qualitative sense. The four components, σ,
two 𝜋, δ, vary considerably in their contributions to total bond strength, that of the δ component being very small
(<10 per cent)

𝑛𝑏 − 𝑛𝑎
Bond order =
2

where nb and na designate the number of electrons occupying bonding and antibonding orbitals, respectively.

• The quadruple bond consists of one σ bond, two π bonds, and one δ bond.

• The quadruple bonds are only about 2.1 Å in length.

Bond order = 4; quadruple bond is exist


between Re atoms The energy of δ bond is about 60 Kcal/mole
The position of the chlorides are in eclipsed position rather
than at a staggered position.
F. A. Cotton et al, Science 1964, 145, 1305−1307
[Re2Cl8]2−
Re−Re distance: 2.24 Å
Orbital configuration: σ2 𝜋4 δ2
The dependence of the overlap on the angle of internal
Bond order = 4; quadruple bond is
rotation is given by cos2𝜒. Therefore, considerable
exist between Re atoms deviation from perfect eclipsing can occur without serious
First description of a metal−metal multiple loss of δ-bonding. Indeed a rotation of 30°, that is, two
thirds of the way towards the fully staggered
bond, by Prof. F. Albert Cotton in 1964.
conformation, causes a loss of only half of the overlap
Electronic configuration in the energy level

d3−d3 d4−d4 d5−d5

Diagram showing the relationship of electronic configurations for M—M bonds of orders 3 and 4
Variation of Bond Orders

• Bond orders may vary, in steps of 1/2, from 1/2 to 4. Those to the left of 4 may be called “electron-poor”
and those to the right “electron-rich.”

• There is another important structural motif for multiple metal–metal bonding. Molybdenum and tungsten
form a large number of triply-bonded compounds of the trigonal type L3M≡ML3, with D3d symmetry,
shown by the representative example, Mo2(NMe2)6.
Examples of Compounds with Bonds of Various Orders
Bond distances for M—M bonds of various orders.

There is a single Mn–Mn bond present in


(OC)5Mn–Mn(CO)5 is 2.92 Å in length and has a
strength of less than 100 kJ/mol. In contrast,
Re−Re in Re2Cl82− is 2.24 Å in length and has a
strength of several hundred kJ/mol.
Peligot’s method to synthesize Cr2(O2CCH3)4(H2O)2

• Cr2(O2CR)4 molecules have


strong tendency to coordinate
electron pair donors at the axial
positions.

• These compounds shows weak


paramagnetism; typically
corresponds to 0.3 to 0.5 BM.
Peligot’s method to synthesize Cr2(O2CCH3)4(H2O)2

Axial coordination occurs even in


these unsolvated compounds by
association of the molecules to
form infinite chains.

The question of how long the Cr–Cr


bond would be in an isolated
Cr2(O2CR)4 molecule, that is, when
axial coordination of any kind is
entirely absent.

The formation of infinite chains of Cr2(O2CR)4 molecules by oxygen


bridge bonding.
Peligot’s method to synthesize Cr2(O2CCH3)4(H2O)2

Axial coordination occurs even in


these unsolvated compounds by
association of the molecules to
form infinite chains.

The question of how long the Cr–Cr


bond would be in an isolated
Cr2(O2CR)4 molecule, that is, when
axial coordination of any kind is
entirely absent.

The formation of infinite chains of Cr2(O2CR)4 molecules by oxygen


bridge bonding.
Unsolvated Cr2(O2CR)4 compounds

dCr−Cr = 2.256(4) Å
Cr2(O2CCPh3)4·C6H6
Benzene can’t participate in sigma donation. Still the
Cr−Cr distance is relatively long. Why?

Cr2(O2CR)4, R = 2,4,6-tri-
isopropylphenyl
Cr2(O2Cbiph)4
dCr−Cr = 1.9662(5) Å
dCr−Cr = 2.348 (2) Å
CH3CN
Cr2(O2CR)4(NCCH3)2

dCr−Cr = 2.3892(2) Å
The Cr−Cr bond distance in
these two compounds are
extraordinarily short compared
to other known metal-metal
bond distances, even when
compared to other quadruple
bonds.

Cr−Cr distance: 1.849 (2) Å


Cr2(TMP)4
The molecular structure of Cr2(2,6-
dimethoxyphenyl)4, Cr2(DMP)4

Cr−Cr distance: 1.847 (1) Å


Cr2(2-MeO-5-MeC6H3)4
Cr2(2-oxophenyl)4
Cr−Cr distance: 1.828 (2) Å Cr−Cr distance: 1.830 (4) Å Cr2(O2CCH3)2(2-Me3COC6H4)2
Cr−Cr distance: 1.862 (1) Å
In terms of Cotton’s ‘formal shortness ratio’,
FSR, for a bond A–B as follows:
The FSR for the Cr–Cr bond in Cr2(DMP)4 is
found to be 1.847/(2×1.186) = 0.779 , for the
Re–Re bond in [Re2Cl8]2− is 0.869

When the Cr2(DMP)4 and


Cr2(TMP)4 compounds were
first reported, the Cr–Cr
bonds were described as
“super-short” because lengths
<1.9 Å were so far below
those in any previously
known metal-metal bonded
compounds.
Connie C. Lu et al, Angew. Chem. Int. Ed. 2013, 52, 4449 –4452
Transition and Inner Transition Metal Chemistry:
CML 524

• Metal-metal bonded Reactivity of Introduction to


compounds and Transition metal Bioinorganic
Chemistry
transition metal Complexes:
• Bioinorganic
cluster compounds • Substitution chemistry of iron:
• Uses of lanthanide Reactions hemoglobin,
complexes: as shift • Electron transfer myoglobin,
reagents, as strong reactions cytochromes
magnets, and in • Photochemical • Non-heme Iron
Enzymes
fluorescence reactions of
• Bioinorganic
• Use of transition transition metal chemistry of zinc
metal complexes in complexes and cobalt
catalysis • Copper dependent
enzymes

Book: Sharpe and Houscroft, Inorganic Chemistry, 4th Edition


Meissler and Tarr, Inorganic Chemistry, 3rd Edition
Parameters of a chemical reaction
A + B → C Rate = k[reactants]n

➢ Kinetics of a reaction Reaction coordinate


directly related to the
activation energy

The activation energy is


the energy difference
between the reactants
and the transition state

➢ Kinetics of a reaction
directly related to the Reaction coordinate
activation energy
Rection Intermediate
Transition state
Chemical reactions of inorganic complexes
➢ Laboratory scale reaction

➢ Industrial scale reactions Hydroformylation


R
Hydrogenation C CH2
H
RHC CH2 + H2 RCH2CH3 HCo(CO)4 CO,
200 bar, H2
110°C
Methanol to acetic acid process
R
CH CH2
CH3OH + CO CH3COOH
H HC
O
Reactions of coordination complexes
(1) Substitution at the metal center,
(2) Oxidation–reduction,
(3) Reactions of the ligands that do not change the attachments to the metal center.

1. Substitution at the metal center

A well-studied class of substitution reactions involves aqueous metal ions ( [M(H2O)m]n+ ) as reactants. These reactions can
produce colored products used to identify metal ions:

Ligand substitution is governed by the M-L bond strength of reactants and product.
Kinetic Lability and Inertness of Complexes
1. Labile Complex.
➢ A labile complex has a very low activation energy for ligand substitution.
➢ Taube suggested a reaction half-life (the time it takes for the concentration of
the initial compound to decrease by one half) of one minute or less as the
criterion for lability.

Complexes such as [Cr(H2O)6]2+, which react rapidly, essentially


exchanging one ligand for another within the time of mixing the reactants,
are classified as labile. Prof. Henry Taube
Nobel Prize in 1983,
Inner sphere reaction mechanism

2. Inert Complex. Compounds that react more slowly are called inert complex.
➢ an inert compound does not resist ligand substitution; it is simply slower to
react, with a higher activation energy for ligand substitution.
➢ Reaction half life is less than a minutes

Example: [Co(NH3)6]3+
Factors affecting kinetic inertness or lability

1. Oxidation state of the metal ion


Predict the hydrolysis rate or find out the most inert one:
Inertness increases with oxidation states [AlF6]3-, [SiF6]2-, [PF6]- , [SF6]

2. Size of the metal ion


Inertness increases when size decreases Find out the most inert one among the hexa-aqua
complex of Na+, K+, Cs+, Rb+

3. Effect of chelating ligand to the metal ion


[Co(en)3]3+ will be more inert than [Co(NH3)6]3+

4. Coordination number around the metal


Coordinatively unsaturated or low coordination number (CN 4)
center
are labile in nature
Water exchange in transition metal complex (kinetics)
Aqueous solution of transition metal ions

➢ A dynamic equilibrium in aqueous solution of metal complex: Exchange rate vary widely as a function of the metal ion

➢ Rate constants are known for many substitution reactions, and general trends in the speeds of these reactions have been correlated to the
electronic configuration of the starting complex. The rate constants for water exchange differ by more than 13 orders of magnitude for
[Cr(H2O)6]3+ and [Cr(H2O)6]2+ ! It is also intriguing that [V(H2O)6]3+ undergoes water exchange roughly 6 times faster than does
[V(H2O)6]2+, even though ligand loss from a V(III) ion might be expected to be more difficult than from a V(II) ion on the basis of an
electrostatic argument.
Kinetics of water exchange reaction
Correlation of Kinetic inertness and Lability of transition metal complex from their
electronic configuration

▪ Inert octahedral complexes are generally those with high


ligand field stabilization energies, specifically those with d3
or low-spin d4 through d6 electronic structures.
▪ These configurations have no electrons in the eg orbitals, and
at least one electron in each t2g orbital
▪ The d3 inert classification is exhibited in Table; [Cr(H2O)6]3+
undergoes water exchange exceedingly slowly relative to the
high-spin d4 [Cr(H2O)6]2+, and [V(H2O)6]2+ reacts slower
than [V(H2O)6]3+.
▪ With strong-field ligands, d8 atoms often form inert square-
planar complexes. Compounds with other d configurations
tend to be labile, with a wide range of substitution reaction
rate constants.

CFSE value
CSIR NET 2017, chemical sciences

Ans key. 1

GATE 2018, Chemistry

Ans key. D

CSIR NET 2018, chemical sciences

Ans key. 3
Thermodynamic stability of complexes
➢ These kinetic terms must be distinguished from the thermodynamic descriptions stable and unstable.

➢ Thermodynamic stability can be predicted by looking at the Stability constant (Keq) or formation constant of
product and reactant
+ +
Keq ++
[𝑀 𝐻2𝑂 6] 𝑛 𝐿 = M H2O 5L n H2O

Keq = kf/kb
> 1…….kf>kb; product is more stable
< 1…….kf<kb; reactant is more stable

DGo = -RT lnKeq


DGo < 0; Stable product and unstable reactant
DGo > 0; Stable reactant and unstable product

• A species such as [Fe(H2O)5(F)]2+ is very stable; it has a large equilibrium constant for the formation, but it is also labile.
On the other hand, hexamine cobalt (3+) is thermodynamically unstable in acid. But [Co(NH3)6]3+ reacts very slowly and
is therefore inert (the activation energy for the above reaction is high). Both thermodynamic and kinetic parameters are
necessary to understand chemical behaviour.
Mechanism of Substitution reactions
➢ In inorganic substitutions, the limiting mechanisms are dissociative (D), in which the intermediate has a lower coordination number than
the starting complex (eq. 1), and associative (A), in which the intermediate has a higher coordination number (eq. 2).

(eq. 1)

(eq. 2)

➢ Between the two extremes is interchange, I, in which the incoming ligand assists in the reaction, but no detectable intermediates appear.
When the degree of assistance is small and the reaction is primarily dissociative, it is called dissociative interchange, Id . When the
incoming ligand begins forming a bond to the central atom before the departing ligand bond is weakened appreciably, it is called
associative interchange, Ia.
Dissociation mechanism and kinetics
In a dissociative ( D ) reaction, the first step is loss of a ligand to form an intermediate with a lower coordination number. Subsequent
additions of either a new ligand (Y) or the leaving group (X) are two possible reaction pathways for this intermediate

Steady-state approximation on the intermediate formation. This approximation assumes that a vanishingly small (and constant)
concentration of the intermediate, ML5, is present during the reaction by assuming that the rates of formation and consumption of the
intermediate are equal. If these rates are the same, the change in the [ML5] must equal zero (and this species cannot accumulate) during
the reaction. Expressed as a rate equation

Final rate law


Limiting Case I.

Limiting Case II. At high [Y], the system will show saturation
kinetics in which the reaction rate depends only
upon [ML5X].

Association mechanism and kinetics


In an associative reaction, forming an intermediate with an increased coordination number is the rate-determining step. This
first step is followed by a faster reaction in which the exiting ligand is lost

The steady-state approximation, which assumes [ML5XY] to be exceedingly small, results in second-order kinetics regardless
of the concentration of Y:

J. H. Espenson, Chemical Kinetics and Reaction Mechanisms, McGraw-Hill, New York, 1981, pp. 45–48
Interchange mechanism and kinetics
An interchange ( I ) reaction in its simplest form is a direct replacement of the leaving group with the incoming group that does not
proceed via an intermediate, but rather a single transition state leading to the conversion of reactants to products

Case I: irreversible replacement

If the substitution reaction is irreversible, an interchange mechanism will exhibit second order kinetics, being first order in [ML5X] and
first order in [Y].

Case II: Reversible replacement

Rate law:
Thermodynamics of Reactions

➢ Kinetic effects are related to thermodynamic effects by linear free-energy relationships (LFER)

The Arrhenius equation for temperature dependence of rate constants and the equation for temperature dependence of
equilibrium constants justify this correlation

A plot of ln(k/T) against 1/T (an Eyring plot) is


linear; the activation parameters DH and DS can be
determined as shown in Figure (right)

A large negative value of DS is indicative of an


associative mechanism,
Experimentally determined thermodynamic values: Experimentally determined rate constants:

Rate expression for Associative mechanism


Plots of kobs against concentration of the entering group for the reactions of
trans-[PtCl2(py)2] with [SCN] or with I-; both reactions were carried out in
MeOH and so there is a common intercept

U. Belluco et al. (1965) J. Am. Chem. Soc., vol. 87, p. 241


Correlation with binding constant (Keq)

Keq ➢ Hydrolysis of cobalt-ammonia complex

DG0 = -RT lnKeq

The log of the rate constant is plotted against the log of the equilibrium
constant for the acid hydrolysis reaction of [Co(NH3)5X]2+ ions.
Mechanisms of Base Hydrolysis of -cis-[Co(en)2Cl2]+. (a) Retention of configuration in dilute hydroxide. (b) Inversion of configuration in
concentrated hydroxide.
Stereochemistry of product

Dissociation Mechanism and


Stereochemical Changes for
trans - [M(LL)2BX]. (a) Square-
pyramidal intermediate (retention
of configuration). (b) Trigonal-
bipyramidal intermediate (while
only formation of the isomers is
shown, a racemic mixture of cis
isomers is expected since and
trigonal-bipyramidal
intermediate are equally likely
upon X dissociation). (c) Less
likely trigonal-bipyramidal
intermediate (two possible
products).
Dissociation Mechanism and
Stereochemical Changes for
cis-[M(LL)2BX]. (a) Square-
pyramidal intermediate
(retention of configuration).
(b) Trigonal-bipyramidal
intermediate. (c) Unlikely
trigonal bipyramidal
intermediate.
Associative mechanism
Conjugate base mechanism
Some cases in which second-order kinetics suggest an associative mechanism are believed to proceed via a conjugate base mechanism ,
12 called SN1CB for substitution, nucleophilic, unimolecular, conjugate base. 13 These reactions depend on amine, ammine (NH3), or
aqua ligands that can be deprotonated to form amido or hydroxo species that subsequently release a ligand dissociatively. It is often the
ligand trans to the amido or hydroxo group that dissociates. Octahedral Co(III) complexes seem particularly predisposed to this
mechanism. The small radius of low-spin Co(III) may render it sufficiently Lewis acidic to stabilize the five coordinate intermediate via p
interactions.

➢ Predicting reaction mechanism (stereochemistry of product)

OH- OH-
+

Reactions with [Co(tren)(NH3)Cl]2+ isomers show that the nitrogen atom trans to the leaving group is most likely to be deprotonated in
the conjugate base mechanism. The reaction in Figure a is 104 times faster than that in Figure b.
Conjugate base mechanism

The dominant kinetic product in both


reactions is hypothesized to proceed by a
trigonal-bipyramidal intermediate or
transition state with the amido ligand in
the trigonal plane that is ultimately trans
to the hydroxo ligand.

The reaction in Figure a has a much lower barrier to obtain a trigonal-bipyramidal geometry (with only
slightly widening of two N—Co—N angles than does the Figure b reaction, which requires more
dramatic rearrangement of a square-pyramidal structure.

The activation energy associated with


hydrolysis of the square pyramidal
intermediate in Figure b , where the
hydroxo ligand would engage an orbital
orthogonal to the amido nitrogen, is
sufficiently high that the (slow)
rearrangement to a trigonal-bipyramidal
structure, followed by hydrolysis is still
the faster pathways.

D. A. Buckingham, P. J. Creswell, A. M. Sargeson, Inorg. Chem., 1975 , 14 , 1485


Water exchange mechanism of the transition metal salts
First order rate constants, k, for reaction vary greatly among the
first row d-block metals (all high-spin Mn+ in the hexaaqua ions):

➢ The rates of water exchange in high-spin hexaaqua ions follow the sequences:

For a series of ions of the same charge and about the same size undergoing the same reaction by the same mechanism, we may reasonably
suppose that collision frequencies and values of DS‡ are approximately constant, and that variations in rate will arise from variation in DH‡.
Let us assume that the latter arise from loss or gain of CFSE on going from the starting complex to the transition state: a loss of CFSE
means an increase in the activation energy for the reaction and hence a decrease in its rate.
L

L L
L L
L L
L L L Dissociative
L Associative +
L M
X +X L L -L L
L
L L L
L
L
L
Considering dissociative mechanism

Changes in CFSE (DCFSE) on converting a high-spin octahedral complex into a square-based pyramidal (for a dissociative process) or
pentagonal bipyramidal (for an associative process) transition state, other factors remaining constant
• Rudolph A. Marcus Electron transfer theory
Occurrence of electron transfer reaction How electron transfer occurs ?

Potential Energy surface


Electron-transfer reactions

In 1953, Taube (who received the Nobel Prize for Chemistry in 1983) made the classic demonstration of an inner-sphere
reaction on a skilfully chosen system in which the reduced forms were substitutionally labile and the oxidized forms were
substitutionally inert.

Intermediate species
Reaction Pathway of inner sphere reaction

These reactions proceed in three steps


(1) a substitution reaction that leaves the Most inner-sphere processes
oxidant and reductant linked by a bridging exhibit second order kinetics
ligand overall, and interpreting the
(2) the electron transfer, frequently data is seldom simple
accompanied by transfer of the ligand
(3) separation of the products
This transfer of chloride to the chromium has been
established experimentally. The products can be
separated by ion exchange techniques, and Cr(III)
complexes are inert. The isolated Cr(III) appears
exclusively as [(Cl)Cr(H2O)5]2+. To show that chloride
transfer from a coordination complex to [Cr(H2O)6]2+ is
feasible to promote inner-sphere electron transfer, the
[Cr(H2O)6]2+ >[Cr(H2O)5Cl]2+ Cr(II)/Cr(III) exchange
reaction, which results in no net change, has been
studied using radioactive 51Cr. All the transferred
chloride in this reaction comes from [Cr(H2O)5Cl]2+,
with none entering the Cr(II) coordination sphere even
when excess Cl- is present in the solution.
Reaction Pathway of inner sphere reaction: Isolable intermediate

➢ In the reaction between [Fe(CN)6]3- and [Co(CN)5]3-, the intermediate (which is stable enough to be precipitated as the Ba2+ salt) is
slowly hydrolysed to [Fe(CN)6]4- and [Co(CN)5(OH)2]2- without transfer of the bridging ligand.

Bridging ligand can enhance the


possibility of inner sphere
reaction
Stable intermediate

Rate of electron transfer:

➢ Nature of birding ligand


For the reaction below with a range of ligands X, the rate determining
step is electron transfer, the rates of reaction depend on X?
The increase in k along the series F, Cl, Br, I correlates with increased ability of the halide to act as a bridge.

The value of k for [OH]- is similar to that for Br,


but for H2O, k is very small and is also pH-
dependent. This observation is consistent with
H2O not being the bridging species at all, but
rather [OH]-, its availability in solution varying
with pH.

Bridging ligand can enhance the


possibility of inner sphere
reaction
Outer-sphere mechanism
When both reactants in a redox reaction are kinetically inert, electron transfer must take place by a tunnelling or outer-sphere mechanism

Clearly, the reactants must approach closely for the electron to migrate from reductant to oxidant. This reductant–oxidant pair is called the encounter or precursor
complex

There is no overall reaction and therefore DGo = 0, and K = 1.

Electron transfer between [Co(NH3)6]2+ (Co–N:211 pm) and


[Co(NH3)6]3+ (Co–N:196 pm) requires not only changes in bond
lengths, but also a change in spin state: [Co(NH3)6]2+ is high-spin
d7 (t2g5eg2) and [Co(NH3)6]3+ is low-spin d6 (t2g6eg0). Transfer of
an electron between the excited states shown in Fig. 26.10 leads
to a configuration of t2g5eg1 for {[Co(NH3)6]3+}* and t2g6eg for
{[Co(NH3)6]2+}*.

You might also like