You are on page 1of 10

I.

Results and Discussion

Different complexes were analyzed spectroscopically in the exercise. Furthermore, their


spectra were generated and the corresponding splitting parameter (Δo) was calculated. Electronic
energy transitions by the absorption of light is the cause of the spectacular colors of transition
metal complexes. These energy levels of an atom are dependent on the total angular momentum
including the spin; hence, electronic configuration cannot solely describe these energy level. In
quantum mechanics, there are abbreviated symbols that describe the total angular momentum
quantum numbers in a polyelectronic system and this is referred as the term symbols.
LS (L=total orbital angular momentum; S=total spin) coupling is the usual basis of the term
symbols. A term symbol describes an atomic state and has a general form
_2𝑆+1 𝐿𝐽

where S is the total spin quantum number, L is the total orbital quantum number in spectroscopic
notation, and J is the total angular momentum quantum number.
The equation 2S+1 refers to the number of possible states of J for a given L and S, assuming
that L is larger or equal to S. This representation is called the spin multiplicity (Atkins et al, 2010),
which corresponds to singlets, doublets, triplets, quartets and quintets states with multiplicity of 1,
2, 3, 4 and 5, respectively. In general, the lowest energy corresponds to the configuration with the
highest multiplicity (Atkins et al, 2010).
When L is equal to 0, 1, 2 and 3, it represents the nomenclature S, P, D and F, respectively.
This is derived from the characteristics of the spectroscopic lines correspond to the orbitals s, p, d
and f (sharp, principal, diffuse, and fundamental, respectively). The rest of the values of L follows
the alphabetical order, except J is omitted (i.e. when L=4, the symbol is G; L=5, the symbol is H)
(Atkins et al, 2010).
The Hund’s rule is the basis for the calculation of the term symbol for the ground state of
an atom . This is the state where both S and L are in their maximum possible values. For the ground
state term symbol, the most stable electron configuration of the desired atom is considered. The
overall angular momentum is not affected by the full shells and full subshells, so they are not
considered.
Following the Pauli Exclusion Principle, the electrons are distributed in the available
orbitals. The orbital with the highest ml value will be filled first as stated by the Hund’s rule. The
overall value of S can be calculated by getting the half of the number of unpaired electron (i.e. if
there are 3 unpaired electrons, the value of S will be S=3/2 or 1.5). Conceptually, this is possible
because according to Hund’s rule, unpaired ground state electrons have parallel spin (same value
of ms) which is conventionally equal to +½. When a second electron fills a singly occupied orbital,
the spin of the second electron will be equal to -½; hence, the spin of the two electrons will cancel
out, the total ms will be zero and it will have no effect to the total spin quantum number.
The overall value of L is computed by adding the ml values for each electron. For example,
if there are two electrons in the same orbital, add twice that ml of the orbital.
For instance, Cu2+ in the complex [Cu(H2O)6](NO3)2 has a ground state d-orbital
configuration of d9. Therefore, there are 9 electrons to place in the subshell d (l=2). There are 5
orbitals (ml=2, 1, 0, -1, -2) that can hold up 10 electrons. The first 5 electrons can take ms=+½, but
the Pauli Exclusion Principle forces the next 4 electrons to have ms=-½ because they go to the
already occupied orbitals. With this, only one electron is unpaired, so S will be equal to ½.
Therefore, the spin multiplicity is equal to 2. For the value of L, there are 2 electrons occupying
the orbitals with ml equal to 2, 1, 0 and -1 and there is only one electron in ml=-2. Computing these,
the value of L will be equal to L=(2x2)+(1x2)+(0x2)+(-1x2)+(-2x1)=2, which corresponds to the
D notation. Its ground state term symbol is 2S+1L=2D. This shows that at ground state, Cu2+ has 2
possible states, read as doublet. This spin is due to one unpaired electron, as a result of Hund’s
rule. It also has a total orbital quantum number notation of D, which corresponds to the
characteristics of the spectroscopic lines of the diffuse (d) orbital. The following table shows the
ground state atomic and spectroscopic term for the metal ion for each of the given complex.
Table 2.1 Determination of the ground state and spectroscopic term symbols for the given
complexes.
Solution Complex Metal d-orbital Ground state
no. ion configuration atomic term
of metal ion symbol
1 [Cu(H2O)6](NO3)2 Cu2+ d9 2D

2 [Cu(en)2(H2O)2](NO3)2 Cu2+ d9 2D

3 [Cu(en)3](NO3)3 Cu3+ d8 2
D
4 [Ni(H2O)6](NO3)2 Ni2+ d8 3F

5 [Ni(en)6](NO3)2 Ni2+ d8 3F

6 [Co(H2O)6](NO3)2 Co2+ d7 4F

7 [Co(en)3](NO3)2 Co2+ d7 4F

8 [Co(NH3)5(H2O)](NO3)2 Co2+ d7 4F

9 [CoCl(NH3)5]Cl2 Co3+ d6 5D

9 KMnO4 Mn7+ d0 -
10 K2Cr2O7 Cr6+ d0 -
11 K3[Co(C2O4)3]·3.5H2O Co3+ d6 5D

12 K3[Fe(C2O4)3]·3H2O Fe3+ d5 6S

13 K3[Cr(C2O4)3]·3H2O Cr3+ d3 4F

14 [Ti(H2O)6]2+ Ti2+ d2 3F

15 cis-[CoCl2(en)2]+ Co3+ d6 5D

16 trans-[CoCl2(en)2]+ Co3+ d6 5D

Various spectroscopies of transition metal coordination complexes, in particular optical


spectra or colors, can be explained by the crystal field theory (CFT). According to CFT, in the
formation of metal complexes, the attraction between the positively charge metal cation and the
negatively charged on the non-bonding electrons of the ligand initiates the interaction between a
transition metal and ligands. In this case, there is a change in the energy of the five degenerate d-
orbitals when the metal ion becomes surrounded by an array of point charges of the ligands. The
loss of degeneracy of the d-orbitals happens when the electrons from the ligands come closer to
some of the d-orbitals and farther away from others as the ligand approaches a metal ion. As a
result, there will be a repulsion between the electrons in the d-orbital and those in the ligand.
Therefore, there will be a splitting in energy in the d-orbitals as the d-orbital electrons that are
closer to the ligands will now have a higher energy than those further away.
In the exercise, the given metal complexes follow the octahedral geometry, in which there
are six electron pairs (oxalate ion is bidentate) bonded to the metal center. In the octahedral
complexes, the dxy, dxz and dyz orbitals will have a lower energy than the dz2 and dx2-y2, because the
latter orbitals are closer to the ligands and experience greater repulsion. The two d-orbitals with
higher energy are collectively termed as eg while the other three that have lower energy are referred
to as t2g. The splitting of the d-orbitals into two sets has an energy difference of Δo.
The ligands greatly affect the size of the gap of Δ. To know how much a ligand affects the
splitting of the d-orbitals, one may use the spectrochemical series, in which ligands are listed
according to the size of the splitting that they produced. According to the series, CO gives the
greatest size of the splitting while I- gives the least (Atkins et al, 2010):
I- < Br- < S2- < SCN- <cl- < NO3- < F- ̴ urea < OH- < CH3COOH < C2O42- < H2O < NCS-
< CH3CN < pyridine ̴ NH3 < en < dipyridine (bipyridyl) ̴ o-phen < NO2- < PR3 < CN- <
CO
Ligands that cause small splitting of the d-orbitals are called as weak field ligand, such as
halogen ions (Miessler et al, 2014). On the other hand, ligands that give large splitting and high Δ
magnitude are referred to as strong field ligand, such as CO and CN- (Miessler et al, 2014). For
complexes with strong field ligands, it is not possible for electrons to be in the high energy orbitals.
In this case, the electrons completely filled the lower energy orbitals first before they go to the
higher energy orbitals (does not follow Hund’s rule). Complexes that has strong field ligands are
called low spin complexes (Miessler et al, 2014). On the other hand, complexes with weak field
ligand favor electrons to be in the higher energy orbitals than to have two electrons in the same
low-energy orbital (Miessler et al, 2014). This configuration follows Hund’s rule as each d-orbital
must be filled first by one electron before pairing happens. Such metal complex is called high spin.
The splitting energy parameter of the d-orbitals is called the ligand field splitting energy
(Δo) and it can be calculated using the Tanabe-Sugano diagrams. The Tanabe-Sugano diagrams
are used for the analysis metal of complexes in terms of spectroscopic data such as absorption in
the UV, visible and IR electromagnetic spectrum as they are useful to approximate the value of
Δo. The Tanabe-Sugano diagrams also considers the high spin and low spin configuration of the
complexes; therefore, these diagrams can also be useful to predict the size of the ligand field
necessary to cause high-spin to low-spin transitions.
The parameters included in the Tanabe-Sugano diagrams are the ligand field splitting
parameter Δo, the Racah parameter B, and energy E (Atkins et al, 2010). In the x-axis, the Δo is
divided by B. On the other hand, in the y-axis, E is also scaled by B. Moreover, the y-axis also
includes the term symbols for a specific dn ion in order of increasing energy. Each electronic state
corresponds to one line in the diagram. The mixing of terms with the same symmetry causes the
bending of some certain linesn (Atkins et al, 2010). Spin-forbidden electronic state energy levels
are also included in the diagram as they can appear in the spectrum although electronic transitions
are only allowed if the electrons do not change from spin up to spin down or vice versa when
moving from one energy level to another. Symmetry label is given by each states, but “g”
(centrosymmetric) and “u” (antisymmetric) are usually omitted because it is understood that all
states are gerade. The Hund’s rule determines the relative order of energies in the diagram. The
following table shows the corresponding splitting of the spherical, free ion terms symbols for an
octahedral complexes (Atkins et al, 2010).
Table 2.2 Splitting of Term Symbols from Spherical to Octahedral Symmetry
Term Degeneracy States in octahedral field
S 1 A1g
P 3 T1g
D 5 Eg+T2g
F 7 A2g+T1g+T2g
G 9 A1g+Eg+T1g+T2g
H 11 Eg+T1g+T1g+T2g
I 13 A1g+A2g+Eg+T1g+T2g+T2g
A=non-degenerate state; E=twofold degenerate state; T=threefold degenerate state; 1=twofold
rotational symmetry around the coordinate axis; 2=twofold rotational symmetry around an axis
between the coordinate axes
There are vertical lines drawn at a specific Δo/B value at some certain Tanabe-Sugano diagrams.
This discontinuity in the slopes of the energy levels of the excited states occurs when Δ equals to
spin pairing (Atkins et al, 2010). Complexes to the right of this line are low-spin, while the left are
for the high-spin. The d2, d3 and d8 configurations has neither low-spin nor high-spin. Moreover,
the Tanabe-Sugano diagrams do not consider the d1, d9 and d10 configurations. In the d1 complex,
there is no electron repulsion as there is only one electron occupies the t2g orbital ground state
while in the d10 configuration, there is no d-d electron transitions because the d orbitals are all
completely filled.
The Racah parameters, in general, represents the total electrostatic repulsion between
electrons and they are expressed in three variables: A, B, and C (Atkins et al, 2010). A is for the
average total interelectronic repulsion, which is constant among d-electron configuration and has
no effect in relative energies; hence, it is not considered in the Tanabe-Sugano diagrams (Atkins
et al, 2010). B and C, on the other hand, is for the individual d-electron repulsions (Atkins et al,
2010). However, C is only important in some cases such as predicting spin-forbidden transitions;
therefore, Tanabe-Sugano diagrams only consider the Racah parameter B (Atkins et. al, 2010).
The arrangement of electrons in the d-orbital is not only limited by electron repulsion
energy, but also to the splitting of the orbitals due to the ligand field for a centrosymmetric ligand
field. The combination of the electron repulsion and splitting energies leads to the high-spin and
low-spin states. The Tanabe-Sugano diagram shows the splitting of spectral parameters with the
increase of the ligand field strength. In this way, it is now possible to determine the energy of the
different configuration states, depending on the strength of the ligand attached to the metal ion.
In order to interpret the given spectra in the Tanabe-Sugano diagram, the states to be
considered should be related to spin-allowed transitions. Thus, the states to be considered must
have the same spin multiplicity as the ground state in the Tanabe-Sugano diagrams. This will
corresponds to the transitions from the ground state to the excited state. For example, the d8
configuration has a ground state of 3A2g according to the Tanabe-Sugano diagram. Therefore, the
allowed transitions are as follow:
_3 𝐴2𝑔 → _3 𝑇2𝑔

_3 𝐴2𝑔 → _3 𝑇1𝑔

_3 𝐴2𝑔 → _3 𝑇1𝑔 (𝑃)

In the exercise, the ligand field splitting parameter for each given complex was calculated
using the Tanabe-Sugano diagrams. Once the Δo was determined, it was used to compute the
corresponding energy absorbed by each complex. The following table relates the color of the
complex and its corresponding value of Δo and energy.
Table 2.3 Relationship of Δo and energy to the color of complex
Complex Color λmax (nm) Δo (cm-1) Energy (kJ/mol)
[Cu(H2O)6](NO3)2 Light blue/sky 739, 283 24433.74439 292.2924928
blue
[Cu(en)2(H2O)2](NO3)2 Indigo to violet 347, 274 27388.94301 351.5069061
[Cu(en)3](NO3)3 Clear royal 589, 457, 362 16759.69815 200.4905131
blue
[Ni(H2O)6](NO3)2 Light clear 636, 536, 457, 15570.67060 186.2665849
green 344
[Ni(en)3](NO3)2 violet 536, 459, 358 22672.45642 271.2228097
[Co(H2O)6](NO3)2 Light red 636, 459, 364 32695.61045 391.1263691
[Co(en)3](NO3)2 Red-orange 636, 480 20584.27755 246.2426431
[Co(NH3)5(H2O)](NO3)2 493 20283.97566 242.6503477
[CoCl(NH3)5]Cl2 740, 524, 360
KMnO4 Purple 636, 525, 458, 17385.44474 207.9761048
359
K2Cr2O7 Very light 636, 457,363 18802.55426 224.9284992
yellow
K3[Co(C2O4)3]·3.5H2O clear green 603, 426 8586.845736 102.7214866
K3[Fe(C2O4)3]·3H2O Turbid yellow 738, 406
K3[Cr(C2O4)3]·3H2O Blue green 562, 412
[Ti(H2O)6]2+ 503 19880.71571 237.8261743
cis-[CoCl2(en)2]+ 514, 590 3479.201109 41.83581582
trans-[CoCl2(en)2]+ 610, 514 3816.990179 45.66134264

In general, it can be observed from the table that the color of the solution that has large value of
Δo appears further to the right of the electromagnetic spectrum (colors with shorter wavelengths).
For example, the violet nickel(II) complex has a larger value of Δo than the green nickel(II)
complex. Another, the indigo copper(II) complex has a larger Δo than its blue complexes. Since
violet has a shorter wavelength than blue, it can absorb higher energy photons as indicated by the
equation,
E = ℎc⁄𝜆

In general, solutions that appear to have colors of shorter wavelength can be expected to have
larger values of Δo than those solutions that has colors of longer wavelength.
The observed trend to the values of the ligand field splitting energies can be explained by
the following factors: nature of ligands (π-acceptor or donor), charge/oxidation state of metal ion,
nature of the metal ion involved, and the number of ligands and geometry of the complex.
The nature of the ligands surrounding the metal ion greatly affects Δo. In general, there will
be a greater energy difference between the high and low energy d orbitals if the effect of the ligand
is stronger. The strength of the ligand can be explained by the ligand field theory (LFT). This
theory describes the characteristics of coordination complexes in terms of bonding and orbital
arrangement by applying the concept of molecular orbital theory. According to LFT, the σ-
symmetry orbitals of the ligands form bonding and anti-bonding combinations with the dz2 and
dx2-y2 orbitals (Miessler et al, 2014). The remaining three d-orbitals stay as non-bonding orbitals.
The total 6 bonding and 6 anti-bonding molecular orbitals are products of additional weak bonding
interactions with the s and p orbitals of the metal.
On the other hand, the π-bonding comes in two ways. First, the metal-to-ligand π bonding
(π backbonding) occurs when the lowest unoccupied anti-bonding π* orbitals of the ligand
combine with the three low-energy d-orbitals of the octahedral complexes to form bonding orbitals
(Miessler et al, 2014). The bond between the ligand and the metal becomes stronger and there is
an increase in Δo after the new π bonding orbitals are filled with electrons from the metal d-
orbitals, since the anti-bonding π* orbitals are higher in energy than the anti-bonding orbitals from
the σ bonding. As a final result, the corresponding π bond within the ligand weakens as the ligands
end up with electrons in their π* molecular orbital.
The second type of π bonding is the ligand-to-metal π bonding which occurs when π
orbitals donate electrons to the π-symmetry bonding orbital between them and the metal that is a
result of the combination of the three low-energy d-orbitals of the octahedral complex and the π
orbitals of the ligands (Miessler et al, 2014). However, the energy of the anti-bonding molecular
orbital from the σ bonding is higher than that of the anti-bonding molecular orbital from the ligand-
to metal bonding. As a result, Δo decreases and the anti-bonding molecular orbital from the ligand-
to-metal bonding is filled with electrons and becomes the highest occupied molecular orbital
(HOMO).
There are two types of ligands responsible for the trend in the ligand field splitting energy
and they are based to the nature of π-interaction between the ligand orbitals with the d-orbitals on
the central atom: the π-acceptor and the π-donor ligands. The π-acceptor ligands are ligands that
exhibit metal-to-ligand π bonding; thus, have a possibility to occur pi backbonding because of their
vacant π* and d orbitals of appropriate energy (Miessler et al, 2014). This additional bonding
increases the value of Δo. Examples of π-acceptor ligands are the strong field ligands.
On the other hand, π-donor ligands exhibits ligand-to-metal π bonding and has a stronger
metal-ligand interactions than π-acceptor ligands, resulting to a decrease in Δo. This type of ligands
tend to donate electrons from their occupied p orbitals to the metals along with the σ bonding
electrons. Weak field ligands are examples of π-donor ligands.
In the exercise, the effect of the nature of ligands on the value of Δo can be observed to the
complexes with the same metal ion but have different ligands. For example, the complex
[Ni(H2O)6](NO3)2 has a smaller computed value of Δo than [Ni(en)3](NO3)2. Since H2O is a weak
field ligand, it is considered as a π-donor ligand and therefore; expected to have a smaller value of
Δo. On the other hand, ethylenediamine is a π-acceptor ligand as it belongs to the strong field
ligands based on the spectrochemical series; hence, it will have an effective increase to the value
of Δo.
From the perspective of the metal ion center, its charge/oxidation state and its nature also
affect the value of Δo. Generally, the value of Δo increases as the oxidation number of the metal
ion increases, for a given set of ligands. At higher oxidation state, the radius of the metal ion
decreases, resulting for the metal and ligands to have a smaller distance between each other and
causing a stronger repulsion. The increase in the magnitude of the electron repulsion between the
metal and ligands results to a larger value of Δo. For example, cobalt in [Co(en)3](NO3)2 and
K3[Co(C2O4)3]·3.5H2O has an oxidation state/charge of +2 and +3, respectively. By looking at
their respective value of Δ in Table 2.3, it can be observed that [Co(en)3](NO3)2 has a smaller Δo
than K3[Co(C2O4)3]·3.5H2O.
Aside from the charge/oxidation state, the nature, specifically the principal quantum
number, of the metal ion also bring an increase or decrease to the value of Δo. As the principal
quantum number increases, the radius of the valence orbitals also increases; therefore, Δo also
increases given that the metal complexes to be compared has the same charge and same ligands.
In other words, the value of Δo increases down the group in the periodic table. Furthermore,
repulsive ligand-ligand interactions occur when the metal center is smaller. This will result to a
shorter distance between the metal and ligand, making their electronic interaction stronger.
Lastly, the number of ligands and geometry of the complex affect the value of Δo. All
complexes presented in the exercise follow the octahedral geometry, in which there are six ligands
or six electron pairs bonded to the metal center. In the octahedral complexes, the dxy, dxz and dyz
(nonbonding) orbitals will have a lower energy than the dz2 and dx2-y2 (bonding and antibonding),
because the latter orbitals are closer to the ligands and experience greater repulsion (Miessler et
al, 2014). The two d-orbitals with higher energy are collectively termed as eg while the other three
that have lower energy are referred to as t2g. The splitting of the d-orbitals into two sets has an
energy difference of Δoct. On the other hand, another common geometrical shape of coordination
complexes is the tetrahedral shape in which four ligands or four electron pairs form a tetrahedron
around the metal center. Similarly to octahedral complexes, the d-orbitals split into two groups in
the tetrahedral crystal field splitting, with an energy difference of Δtet. However, the grouping of
the d-orbitals in the splitting is different and completely opposite from that of the octahedral crystal
field. This time, the higher energy orbitals will be dxy, dxz and dyz (antibonding and bonding
orbitals) while the lower energy orbitals will be dz2 and dx2-y2 (nonbonding orbitals). This results
to a lower energy splitting (Δ) compared to that of the octahedral field because the ligand electrons
in this case are not oriented directly towards the d-orbitals; hence, there will be less repulsion.
Additionally, there were splitting or broadening of bands observed in the UV-vis
absorbance spectra of some of the complexes. This phenomenon is due to the reduced symmetry
of the ligand field or also known as the Jahn-Teller Effect. The Jahn-Teller Effect shows
geometrical distortion of nonlinear molecules and ions that have a spatially degenerate electronic
ground state. This distortion lowers the overall energy of the molecules, resulting for the
degeneracy to disappear. The Jahn-Teller Effect is common to transition metal complexes,
especially those with octahedral shape (Miessler et. al, 2014). For example, the d9 electronic
configuration of copper(II) in an octahedral copper complex, such as [Cu(H2O)6]2+, fills the two
degenerate eg orbitals with three electrons, resulting to a double degenerate electronic ground state.
The complex may have a distortion (elongation or compression) along the z-axis, however, the
Jahn-Teller Effect only describes the stability of the geometry but not the direction of the
distortion. The elongation distortion makes the energy of the complex lower because the
electrostatic repulsion between the electrons on the ligand and the electrons in the z-component d-
orbitals becomes lower as they have greater distance to each other. On the contrary, the
compression effect increases the energy of the complex as such repulsion becomes greater.
This distortion effect is strong when the eg orbitals are unevenly occupied (odd number of
electrons), such as high spin configuration of d4, low spin configuration of d7 and d9 configuration,
because the eg degenerate orbitals are pointing directly to the ligands (Miessler et al, 2014). In
contrary, the Jahn-Teller Effect is weak when t2g orbitals are unevenly occupied because these
orbitals do not point directly to the ligands; hence, there is smaller lowering of repulsion on taking
ligands further away from the t2g orbitals (Miessler et al, 2014). This weak distortion effect is
evident to d1, d2, low spin d4 and d5, high spin d6 and d7 configurations. On the other hand, no such
effect occurs in d3, high spin d5, low spin d6, d8 and d10 configurations because their orbitals are
equally filled with electrons. In the exercise, the complexes that experienced strong Jahn-Teller
distortion effect were [Cu(H2O)6](NO3)2, [Cu(en)2(H2O)2](NO3)2, [Co(en)3](NO3)2, and
[Co(NH3)5(H2O)](NO3)2.
The color of the metal complexes are due to either of the following mechanisms: the d-d
transitions or charge transfer reaction. A complex is said to exhibit d-d transitions when the
electron in a d orbital of the metal is excited by a photon to another d orbital of higher energy. A
lower d-electron is promoted to a higher excited state when a visible light of the correct wavelength
is absorbed, causing the appearance of the color of the solution. However, these colors are not so
intense. The excitation of the electron follows two selection rules:
The spin rule: ΔS=0
Laporte’s rule: Δl=±1
The selection rules indicate that the electron should not experience a change in spin upon
excitation. When the excited electron changes its spin, it is said to be spin-forbidden. On the other
hand, a d-d transition is said to be Laporte-forbidden of the complex has center of symmetry.
Therefore, for a forbidden transition to be allowed, the center of symmetry of the complex must
first be disrupted, for example, by Jahn-Teller Effect or asymmetric vibrations. The d-d transitions
do not occur in charge-transfer complexes and the absorption and colors in this mechanism is very
intense.
The charge transfer reaction is an electron-donor-acceptor reaction that is responsible for
the color and intensity of transition metal complexes (Atkins et al, 2010). The transfer of electronic
charge from the electron donor to the electron acceptor results to a relative energy balance that is
reflective to the color of the complexes. Moreover, since the charge transfer transitions are always
spin-allowed and Laporte-allowed (unlike d-d transitions that are always Laporte-forbidden), the
absorption bands of these complexes are very intense and can be observed in the ultraviolet or
visible portion of the spectrum. The charge transfer reaction has two mechanisms: excitation of an
electron from a metal orbital of a metal to an empty ligand orbital (metal-to-ligand charge transfer
of MLCT) or promotion of an electron in a ligand orbital into an empty metal orbital (ligand-to-
metal charge transfer or LMCT) (Atkins et al, 2010).
In LMCT reaction, the electron acceptor is the metal and the electron donor is the ligand.
This reaction arises if the complex has ligands with relatively high-energy lone pairs or if the metal
ion has low-lying empty orbitals (Atkins et al, 2010). These type of charge transfer reaction occurs
at metals with high oxidation state that makes their configuration very low (Atkins et al, 2010).
These configuration makes the electron acceptor level of the metal available and low in energy. In
this case, there are two transition of electrons possible: one to the eg level and the other one is to
the t2g level. Examples of LMCT complexes are MnO4- and Cr2O72-, in which both manganese(VII)
and chromium(VI) have a d configuration of d0. This makes the empty molecular orbitals of
manganese(IV) and chromium(IV) be easily filled with electron from the p orbitals of the ligands.
On the other hand, in MLCT reaction, the electron acceptor is now the ligand and the
electron donor is the metal ion. Ligands with low-lying π* orbitals initiate this type of charge
transfer reaction, as the electrons from the highly-filled d orbitals of the metal ion are promoted to
those low-lying π* orbitals (Atkins et al, 2010). Furthermore, the metal ion must also have a low
oxidation state and must be relatively high in energy for this reaction to occur in contrast to the
LMCT reaction (Atkins et al, 2010). Examples of such ligands participating in this reaction are
mostly aromatic compounds and other strong field ligands. In the exercise, the MLCT reaction
was exhibited by the complexes [Co(NH3)5(H2O)](NO3)2 and [CoCl(NH3)5]Cl2. The strong field
ligand ammonia and the low oxidation state of cobalt(II) are the basis for the reaction.

II. Conclusion

In the exercise, transition metal complexes were prepared and investigated by means of
spectroscopic analysis. Moreover, the peaks were located and the transitions corresponding to
them were labeled in each spectrum. From this, the ligand field splitting energy for each complex
were calculated using the Tanabe-Sugano diagram. It was observed that as the color of the complex
goes to the right of the electromagnetic spectrum, its ligand field splitting energy increases.
The factors affecting the value of the field splitting energy were also explained using the crystal
field and ligand field theory. Such factors include the nature of the ligands (π-acceptor and donor),
charge/oxidation state of the metal ion, nature of the ion involved, and number if ligands and
geometry of the complex. It was also observed that splitting and broadening of bands on the spectra
are caused by the reduced or distorted symmetry of the ligand field or the Jahn-Teller Effect.
The colors of the complexes are made by d-d transition reaction or the charge transfer reactions.
The d-d transition reaction follows the selection rules of spin and Laporte’s. The charge transfer
reactions shows the promotion of electron from either metal-to-ligand or ligand-to-metal transfer.
Metal-to-ligand charge transfer is favored if the metal ion has low oxidation state while ligand-to-
metal transfer occurs when the metal has high oxidation state.
It was also observed in the exercise that the value of the ligand field splitting energy of
complexes with the same type and number of ligands but of different central atom depends on the
oxidation state/charge of the metal center. In general, as the charge of the metal center increases,
the ligand field splitting energy also increases. For complexes with the same metal ion but different
ligands, the ligand field splitting energy is based on the nature of the ligand. Strong field or π-
acceptor ligands make a relative increase in the magnitude of the ligand field splitting energy. On
the other hand, weak field or π-donor ligands makes it relatively lower.

III. References
Atkins, P.W., Overton, T.L., Rourke, J.P. Weller, M.T. & Armstrong, E.A. (2010). Shriver and
Atkins’ Inorganic Chemistry (5th ed.). Oxford University Press.
Miessler, G.L., Fischer, P.J. & Tarr, D.A. (2014). Inorganic Chemistry. Pearson Education.

You might also like