Immanuel SnX Work-Final

You might also like

Download as docx, pdf, or txt
Download as docx, pdf, or txt
You are on page 1of 23

Organometallic strontium monochalcogenides [SnX (X = O, S, Se, and Te)] as alternative

ligands for carbonyls: A DFT Study

Selvaraj Immanuela,b, Selvaraj Nishalinic, S. Manivarmanb*, Francisxavier Paularokiadossa,


Thayalaraj Christopher Jeyakumarc*
a
PG and Research Department of Chemistry, St. Joseph’s College of Arts and Science
(Autonomous), Cuddalore, Tamil Nadu-607001, India.
b
PG and Research Department of Chemistry, Government Arts College, C-Mutlur,
Chidambaram, Tamil Nadu-608102, India.
c
PG and Research Department of Chemistry, The American College (Autonomous), Madurai,
Tamil Nadu-625002, India.

*Corresponding authors: S. Manivarman (dsmgac@gmail.com); Thayalaraj Christopher

Jeyakumar (chemistchris24@gmail.com)
Abstract

Quantum mechanical DFT calculations performed on the Fe(CO) 5 and for the axial and

equatorial isomers of [Fe(CO)4(SnX)] (X = O, S, Se, Te) complexes. The equatorial isomer of

[Fe(CO)4(SnX)] complexes has been shown to be more stable than the axial isomer by means of

energetic values. These complexes bonding nature is analysed by NPA and EDA results. The

bond index of the bond between Fe and SnX is provided by the WBI analysis. The FMO studies

these complexes shows lesser HOMO-LUMO energy gaps in the values of 4.00 to 4.76 eV than

that of [Fe(CO)5]. From the NBO analysis, the contribution to the bond formation of the Sn atom

to the Fe-Sn bond is greater than that of the Fe atom. Similar contributions are seen to the

carbonyl group (in the Fe-C bond), although the contribution of the carbon atom is larger than

that of the tin atom.

Keywords: EDA, Strontium chalcogenide, FMO, carbonyl, WBI


1. Introduction

For a very long time, experimental and theoretical chemists have been interested in

transition metal carbonyl compounds. [1-2] where the use of metal carbonyls has sparked

innovation across a range of sectors through uses in fields as diverse as materials research,

medical chemistry, and catalysis. [3-5] In organometallic chemistry, isoelectronic or isolable

connections play a crucial role in the hunt for novel ligand types. Hoffmann proposed “Two

fragments are isolable if the number, symmetry properties, approximate energy and shape of the

frontier orbitals and the number of electrons in them are similar—not identical, but similar” in

1981.[6] SnO, SnS, SnSe, and SnTe all have the same isoelectronic properties as the well-known

ligand carbon monoxide. Carbonyl metal complexes have been extensively studied theoretically

and empirically due to their versatility in catalytic processes, either as a spectator ligand or as a

reactant.[7-18] An essential function of iron carbonyl complexes is in organometallic chemistry.

Because of their distinct electronic properties and ability to interact with transition metals,

strontium chalcogenides offer a promising way to expand the chemistry of carbonyl complexes.

Tin monochalcogenides are compounds with the formula SnX, where X is any of the chalcogens

oxygen, sulfur, selenium, or terbium. The SnO atomic structure has a van der Waals gap caused

by a lone pair oriented toward the interlayer space. The other SnX compounds are more reactive

than SnS and SnTe, where it pushes away tin neighbors and causes just structural deformation.

[19] Density functional theory, or DFT, has proven to be an accurate and efficient method for

calculating transition metal compounds in recent years [20]-[22]. Therefore, the complex

[Fe(CO)4SnX] has been taken, and the interaction between these organometallic ligands has been

systematically examined using DFT, an efficient quantum mechanical approach. This method

offers an invaluable tool for probing the electronic structure and properties of complex molecules
and materials with remarkable accuracy, which gradually change with increasing atomic number.

[23-24]. Jeyakumar et al. compared the bonding patterns of [M(CO) 4] and [M(CO)3(AX)] with

SiX and GeX ligand substituted metal (M = Ni, Pd and Pt) complexes. [25- 26] Thirty-two

distinct [Fe(CO)4(AX)] complexes containing Group 13 monohalides have been studied

theoretically by Jeyakumar et al. as a substitute ligand for carbonyl in organometallics. The

reorganization of tin (Sn) chalcogenide complexes has been minimal in comparison to other

studies, despite the fact that several transition metals have been used in this study. The same

research group of people has used DFT to analyze the structure and bonding of carbonyl

complexes of Ni, Pd, and Pt with terminal group13 monohalide as a ligand in [TM(CO) 3AX]

where A is B, Al, Ga, and In and X is halogen[27]-[32] For this study, we compared the results

of a structural and bonding investigation of [Fe(CO) 4(SnX)] (where X might be O, S, Se, or Te)

to those obtained for [Fe(CO)5]. Eight different [Fe(CO)4(SnX)] complexes with the Sn-X link in

both axial and equatorial orientations have been analyzed by us utilizing NPA, NBO, EDA, and

WBI methods. Substitutions of the Sn-X bond in the axial and equatorial locations for the CO

ligand in [Fe(CO)5] were investigated, and their isolobal and isoelectronic character was

elucidated.

2. Computational details

At the Lee-Yang-Parr (B3LYP) level of theory, we used the Becke three-parameter

hybrid approach to do our computations for all the complexes. [33-34] Fe and Te were calculated

using the Los Alamos National Laboratory 2 double zeta (LANL2DZ) basis set, whereas the

remaining elements were calculated using the split valance 6-31G* basis set.[35] Using the same

theoretical setup we have used for computing geometry optimization, vibrations, Wiberg Bond
Indices (WBI), natural charge analysis, molecular orbital analysis, and Natural Bond Order

(NBO).[36-39] All of the computational work was performed in Gaussian09.[40-41] The Energy

Decomposition Analysis (EDA) was carried out using a triple zeta plus double polarization

(TZ2P) basis set in the Amsterdam Density Functional (ADF) software suite in conjunction with

the generalized gradient approximation: Perdew-Burke-Ernzerhof (GGA: PBE) level theory.[42-

43]” EDA analysis has been used to elucidate the peculiarities of M-SnX bonding.[44]

“Morokuma introduced it initially, and Ziegler and Rauk later developed it. Between two

fragments A and B, the bond dissociation energy, ΔE, is divided into multiple contributions that

may be identified as physically significant entities.[45] First, ΔE consists of two major

components ΔEprep and ΔEint.

ΔE = ΔEint + ΔEprep (1)

The interaction energy comprises three main components:

ΔEint = ΔEelstat + ΔEPauli + ΔEorb (2)

ΔEelstat refers to the electrostatic interaction ΔEPauli, refers to the repulsive interactions and ΔEorb

refers to the orbital interaction terms.

ΔEorb = ΔEσ +ΔEπ (3)

ΔEorb is the combined form of σ- and π-bonding donations (equation 3). The σ- and π-bonding

donations for the complexes showing C2v symmetry have been calculated according to equations

4 and 5. Similarly, for the complexes showing C 3v symmetry, equation 6 and 7 has been used to

calculate the σ- and π-bonding donations.

ΔEσ(C2v) = ΔE(a1) (4)

ΔEπ(C2v) = ΔE(b1) + ΔE(b2) (5)

ΔEσ(C3v) = ΔE(a1) (6)


ΔEπ(C3v) = ΔE(e) (7)

From the orbitals with a2 symmetry, the contributions were negligible (<0.7 kcal mol -1).

More information regarding the EDA is available in the literature.

3. Results and discussion

3.1. Structure and vibrations

The completely optimized structures of complexes 1 through 9 that were studied are shown in

Figure 1. You may get the optimized complexes' energy values in Table S1. The specified bond

lengths, bond frequencies, and structural characteristics are shown in Table 1. Each complex has

a trigonal bipyramidal (TBP) shape. “Those complexes in which the SnX ligand is in the

equatorial position display the C2V point group symmetry, whereas those in which it is in the

axial position display the C3V point group symmetry. [Fe]SnX complexes (where [Fe] = Fe(CO)4)

have dipole moments in the range of 2.0827–0.0928 Debye. Table 1 shows the general tendency

that the Fe-C bond in the axial position of the [Fe]SnX equatorial isomer is shorter than that in

the equatorial position. In the axial position of [Fe]SnX, the Fe-C bond is longer than in the

equatorial position. For each complex, the bond between the carbon and oxygen atoms shows

that the equatorial C-O has a greater bond distance than the axial C-O.” The bond distance

between Fe and Sn is shorter for Sn-X that is equatorially bound, and longer for Sn-X that is

axially bound. Bond angles for complexes in the TBP geometry are typically 180° for axial-Fe-

axial, 90° for equatorial-Fe-axial, and 120° for equatorial-Fe-equatorial. Changes in the

complexes that arise when a replacement is made at the equatorial location are tabulated in Table

1. When a substitution takes place in the axial position, however, the equatorial bond angle of the

complexes remains relatively unaffected.


Table 1. Structural parameters for the optimized geometries of complexes 1-9. Bond lengths are

in Å and angles are in deg.

∠ax- ∠eq- ∠eq-


Complex Fe-Cax Fe-Ceq C-Oax C-Oeq Fe-Sn Sn-X
Fe-ax Fe-eq Fe-ax

[Fe](CO) (1) 1.817 1.806 1.147 1.151 --- --- 180.0 120.0 90.0

[Fe](SnO)ax (2) 1.775 1.798 1.490 1.153 2.449 1.768 180.0 120.0 90.1

[Fe](SnO)eq (3) 1.813 1.789 1.150 1.151 2.432 1.774 179.1 120.7 90.2

[Fe](SnS)ax (4) 1.776 1.798 1.148 1.153 2.440 2.198 180.0 120.0 90.3

[Fe](SnS)eq (5) 1.811 1.789 1.150 1.151 2.424 2.207 177.5 118.7 90.6

[Fe](SnSe)ax (6) 1.775 1.797 1.149 1.153 2.444 2.302 180.0 120.0 90.4

[Fe](SnSe)eq (7) 1.810 1.788 1.150 1.151 2.429 2.311 176.8 118.3 90.8

[Fe](SnTe)ax (8) 1.776 1.796 1.149 1.153 2.434 2.555 180.0 120.0 90.6

[Fe](SnTe)eq (9) 1.810 1.789 1.150 1.151 2.423 2.565 175.5 117.1 91.2
Figure 1. Optimized geometries of the complexes 1-9. Bond lengths are in Å and angles are in

deg.

Since the stretching vibrations that are currently present are readily observable, the IR

spectra are primarily used in the structural analysis of the complexes. Two distinct functional

group sets are identified in the complexes: Fe-CO and Fe-SnX, for which stretching frequencies

are provided. In comparison to axial position-substituted complexes, equatorial position-

substituted complexes exhibit a higher M-CO bond stretching frequency. The M-CO and M-SnX

stretching frequency trends are at odds with one another. The stretching frequency decreases

with increasing bond distance, while the stretching frequency increases with decreasing bond

distance. The stretching frequencies of M-CO and M-SnX are compiled in Table 2. It
summarizes the stretching frequencies of M-CO and M-SnX. The ligands' stretching frequencies

decrease in the following order: CO > SnO > SnS > SnSe > SnTe.

Table 2. Vibrational frequency (cm-1), dipole moment (Debye), point group and total energy

(hatree) for the complexes 1-9.

Point Dipole
Complex Fe-CO Sn-X Total Energy
group moment

[Fe](CO) (1) 2189.78, 2121.49, 2120.96 (a1) --- D3h 0.0000 -690.042694

[Fe](SnO)ax (2) 2078.00 (e), 2103.88, 2151.51 (a1) 847.20 C3v 1.4690 -655.395359

[Fe](SnO)eq (3) 2101.08 (b1), 2106.90, 2157.47 (a1) 837.96 C2v 2.0827 -655.400230

[Fe](SnS)ax (4) 2077.99 (e), 2103.65, 2149.24 (a1) 478.88 C3v 1.0707 -978.406951

[Fe](SnS)eq (5) 2100.24 (b1), 2107.22, 2154.23 (a1) 472.99 C2v 1.8743 -978.411540

[Fe](SnSe)ax (6) 2076.22 (e), 2101.42, 2146.96 (a1) 340.70 C3v 0.0973 -2979.16228

[Fe](SnSe)eq (7) 2097.46 (b1), 2151.57, 2104.82 (a1) 337.96 C2v 0.9439 -2979.16654

[Fe](SnTe)ax (8) 2075.44 (e), 2100.94, 2145.18 (a1) 276.66 C3v 0.0928 -588.278191

[Fe](SnTe)eq (9) 2096.80 (b1), 2105.00, 2149.38 (a1) 277.57 C2v 0.9124 -588.282647

3.2 Bonding Analysis

In this section, the WBI, and NBO of Fe-C and the Fe-Sn bonds in complexes 1-9 are compared

and evaluated similarly atomic partial charges of all atoms were analysed. Tables 3-6 provide the

findings of the discussion of the Wiberg bond indices, atomic charges, and natural bond orbital

analysis of complexes 1-9.

Table 3. Atomic partial charges of the various SnX ligand substituted penta coordinate iron
carbonyl complexes.
Complex Fe Cax Ceq Oax Oeq Sn X
[Fe](CO) (1) -0.562 0.593 0.506 -0.418 -0.435 ---- ----
[Fe](SnO)ax (2) -2.052 0.715 0.822 -0.426 -0.409 1.775 -0.409
[Fe](SnO)eq (3) -2.070 0.768 0.750 -0.422 -0.409 1.709 -1.013
[Fe](SnS)ax (4) -2.012 0.723 0.820 -0.426 -0.409 1.432 -0.723
[Fe](SnS)eq (5) -2.022 0.774 0.753 -0.421 -0.409 1.374 -0.746
[Fe](SnSe)ax (6) -2.008 0.724 0.819 -0.428 -0.411 1.309 -0.596
[Fe](SnSe)eq (7) -2.017 0.774 0.753 -0.423 -0.414 1.252 -0.620
[Fe](SnTe)ax (8) -2.000 0.727 0.818 -0.428 -0.412 1.067 -0.372
[Fe](SnTe)eq (9) -2.001 0.777 0.753 -0.422 -0.411 1.010 -0.402

Through natural population analysis, we were able to find the partial charges of each atom in the

complexes 1-9. When compared to the axial isomer (2, 4, 6, and 8) and the equatorial isomer (3,

5, 7, and 9) of the SnX complexes, the metal atom Fe in each of these complexes carries more

negative charge. Similar results are seen when thinking about X atoms. When compared to the

axially substituted complexes, the positive charge on Sn is much bigger on all of the complexes.

This is especially true for the equatorial-position substituted complexes. All of the carbonyl

carbons in complexes 1–9 are found to be positive charge.

Bonding information, including atomic population, bond polarization, bond contribution (in

percentage), and total number of bonds, were obtained by NBO analysis. A tabulation of these

results may be seen in Table 4.

Table 4. Natural bond order analysis for the complexes 1-9.


Bond
Bond Bond Bond Polarization
Complex Population Contribution (%)
type order
Fe Sn Fe Sn
[Fe](CO) (1) Fe-C ax 1 1.90006 29.93 70.07 0.5470 0.8371
[Fe](SnO)ax (2) Fe-C eq 1 1.83968 19.45 80.55 0.4410 0.8975
[Fe](SnO)eq (3) Fe-Sn ax 1 1.88218 37.13 62.87 0.6094 0.7929
[Fe](SnS)ax (4) Fe-Sn eq 1 1.85788 31.56 68.44 0.5618 0.8273
[Fe](SnS)eq (5) Fe-Sn ax 1 1.87423 35.79 64.21 0.5983 0.8013
[Fe](SnSe)ax (6) Fe-Sn eq 1 1.70108 36.51 63.49 0.6042 0.7968
[Fe](SnSe)eq (7) Fe-Sn ax 1 1.72627 42.87 57.13 0.6547 0.7559
[Fe](SnTe)ax (8) Fe-Sn eq 1 1.85698 30.09 69.91 0.5485 0.8361
[Fe](SnTe)eq (9) Fe-Sn ax 1 1.87327 33.69 66.31 0.5804 0.813
[Fe](CO) (1) Fe-Sn eq 1 1.71273 35.69 64.31 0.5974 0.8019
Table 5. Calculated orbital contribution of Fe for the bonding in complexes 1-9.
Complex S px py pz p dxy dxz dyz dx2-y2 dz2 d
[Fe](CO) (1) 0.47 --- --- --- --- 1.70 1.76 1.76 1.70 1.17 8.09
[Fe](SnO)ax (2) 0.50 0.46 0.46 0.54 1.46 1.66 1.77 1.77 1.66 1.18 8.04
[Fe](SnO)eq (3) 0.52 0.39 0.49 0.54 1.94 1.74 1.75 1.68 1.25 1.69 8.11
[Fe](SnS)ax (4) 0.48 0.45 0.45 0.53 1.43 1.66 1.76 1.76 1.66 1.20 8.04
[Fe](SnS)eq (5) 0.50 0.38 0.48 0.53 1.89 1.74 1.75 1.67 1.26 1.70 8.12
[Fe](SnSe)ax (6) 0.48 0.45 0.45 0.53 1.43 1.66 1.76 1.76 1.66 1.20 8.04
[Fe](SnSe)eq (7) 0.50 0.38 0.48 0.53 1.89 1.74 1.74 1.64 1.27 1.74 8.13
[Fe](SnTe)ax (8) 0.47 0.45 0.45 0.53 1.43 1.65 1.76 1.76 1.65 1.22 8.04
[Fe](SnTe)eq (9) 0.49 0.38 0.48 0.53 1.88 1.74 1.74 1.63 1.27 1.70 8.08

It is evident from the data that complexes 2 through 9 lack a π bonding orbital in the Fe-Sn bond.

According to these findings, the Fe-Sn bond is always a σ bond. The metal Fe is contributing less

to the formation of Fe-Sn bond when compared to Sn. It means that, compared to the Sn atom is

more polarized than Fe. The bond population of the axial isomer of the Fe-SnX complexes (2, 4,

6, and 8) are higher than that of the equatorial isomers. In these Fe-SnX complexes, the axial

isomer has a greater Fe bond contribution for the Fe-Sn bond than the equatorial isomer.

For complexes 2–9, the Fe orbital contribution to the Fe–Sn bond is calculated. The d orbital of

the Fe atom is shown to contribute more than the s and p orbitals in all the complexes. When (d)

> (p) > (s), the orbital contribution is less. All three orbital of s, p, and d contributions are higher

in the equatorial isomer of [Fe]SnX when compared to the axial isomer of the same.

Table 6. Wiberg bond indices analysis for the complexes 1-9.


Complex Fe-Cax Fe-Ceq C-Oax C-Oeq Fe-Sn Sn-X
[Fe](CO) (1) 0.6807 0.5959 2.1663 2.1521 ----- -----
[Fe](SnO)ax (2) 1.0645 1.0959 2.1511 2.1765 0.811 1.4124
[Fe](SnO)eq (3) 1.0538 1.0566 2.1709 2.1685 0.8724 0.1388
[Fe](SnS)ax (4) 1.0488 1.0902 2.1508 2.1768 0.9053 0.1840
[Fe](SnS)eq (5) 0.2821 1.3390 2.4757 1.2841 1.001 0.6942
[Fe](SnSe)ax (6) 1.0494 1.0919 2.1474 2.1739 0.8886 0.1975
[Fe](SnSe)eq (7) 0.2770 1.3415 2.4757 1.2650 0.9845 0.7437
[Fe](SnTe)ax (8) 1.0429 0.2234 2.1465 0.0243 0.8941 0.0013
[Fe](SnTe)eq (9) 1.0396 1.0380 2.1736 2.1611 0.9058 1.9176

The WBIs for the Fe-C and Fe-Sn bonds in complexes 1-9 calculated and Fe-Sn bond fall in the

range of 0.81-1.00, showing a significant covalent contribution in the Fe-Sn bond. When

comparing the complexes, the CO bonds are stronger than the SnX bond because they are greater

in energy than the Fe-C bonds (2-9). According to WBI research, the equatorial Fe-Sn bond in

SnX complexes is stronger than the axial isomer. This example shows that the axial isomer of

SnX complexes is strongly bound by the equatorial bonds. As one progresses down the

chalcogenide group from O, S, Se, and Te, the Fe-Sn bond connection gets stronger.

3.3 Molecular orbital analysis

Figure 2. Molecular orbital analysis for the complexes 1-9. Black colour flat lines shows
HOMO, Red colour flat lines shows LUMO and values are energy gap between HOMO and
LUMO. All the values are in eV.
Each complex has had its molecular orbitals calculated. Higher occupied molecular orbital

(HOMO), lower unoccupied molecular orbital (LUMO), and energy gap (Eg) are tabulated in

Table 7 based on the calculated findings. It is found that the HOMO and LUMO energy

differences are less because the metal center has a larger electron density. The stability of the

complexes is compromised by this narrower energy gap. The Axial isomer of [Fe(CO) 4(SnO)]

complex has a larger energy gap than the other complexes, suggesting it is more stable. If the

energy gap of the SnTe ligand is reduced, the stability of these complexes may decrease.

Table 7. Molecular orbital analysis for the complexes 1-9 (eV).


Complex HOMO LUMO Energy gap
[Fe](CO) (1) -6.90 -1.38 5.52
[Fe](SnO)ax (2) -6.50 -1.77 4.73
[Fe](SnO)eq (3) -6.80 -2.45 4.35
[Fe](SnS)ax (4) -6.42 -1.66 4.76
[Fe](SnS)eq (5) -6.25 -1.90 4.35
[Fe](SnSe)ax (6) -6.86 -2.86 4.00
[Fe](SnSe)eq (7) -6.25 -1.90 4.35
[Fe](SnTe)ax (8) -6.28 -1.77 4.51
[Fe](SnTe)eq (9) -5.44 -1.36 4.08

3.4 Energy decomposition analysis

Table 8 displays the energy decomposition analysis (EDA) for the Fe-CO bond in [Fe(CO) 5] 1

and the Fe-SnX bond in the axial and equatorial isomers of [Fe(CO) 4(SnX)] 2-9. Regardless of

the ligand, the behaviour of Eelstat and Eorb is very consistent. The ionic bonding is controlled by

Eelstat, and the strength of the bonds varies somewhat amongst ligands. Covalent bonding, which

occurs between Eorb and ligands, is very consistent. Therefore, the bonds between [Fe]-L, where

L = CO, and SnX are neither purely ionic nor purely covalent. The orbital effects of the

symmetry have been analyzed. We have extrapolated the contributions from - and -bonding from

these data. It was found that a2 symmetric orbitals had a negligible role (0.7 kcal mol -1) in bond
formation. The orbital interaction term is the sum of these two quantities, E and E. SnX

complexes have an axial isomer with a smaller -bonding contribution and an equatorial isomer

with a smaller -bonding contribution. This results in excellent stability for the equatorial isomers.

In organometallics, researchers examine the effects of ligand location on the binding strength

between metal ions in a complex.

Table 8. EDA for [Fe(CO)5] → [Fe(CO)4] and CO, [Fe(CO)4(SnX)] → [Fe(CO)4] and SnX.
Values are given in kcal/mol. bValues in parentheses give the % contribution to the total
attractive interactions ΔEelstat + ∆Eorb. c Values in parentheses give the contributions to the total
orbital interaction ∆Eorb.
b b c c
Complex ∆Epauli ∆Eelstat ∆Eorb ΔEσ ΔEπ ∆Eint
-96.32 -86.94 -43.91 -43.02
[Fe](CO)ax (1) 130.47 -52.80
(52.56) (47.44) (50.51) (49.49)
-112.32 -88.58 -41.48 -47.14
[Fe](CO)eq (1) 152.70 -48.20
(55.91) (44.09) (46.81) (53.19)
-58.87 -62.62 -48.08 -14.49
[Fe](SnO)ax (2) 85.12 -36.37
(48.46) (51.54) (76.85) (23.15)
-69.61 -52.59 -36.53 -15.76
[Fe](SnO)eq (3) 91.13 -31.07
(56.97) (43.03) (69.86) (30.14)
-59.05 -67.22 -50.93 -16.23
[Fe](SnS)ax (4) 87.37 -38.90
(46.77) (53.23) (75.83) (24.17)
-70.08 -57.37 -39.24 -17.84
[Fe](SnS)eq (5) 93.52 -33.93
(54.99) (45.01) (68.75) (31.25)
-58.86 -68.31 -52.21 -16.05
[Fe](SnSe)ax (6) 87.66 -39.52
(46.29) (53.71) (76.49) (23.51)
-70.06 -58.55 -40.51 -17.75
[Fe](SnSe)eq (7) 94.12 -34.49
(54.48) (45.52) (69.53) (30.47)
-60.10 -71.73 -55.00 -16.67
[Fe](SnTe)ax (8) 91.55 -40.28
(45.59) (54.41) (76.75) (23.25)
-71.49 -61.91 -42.99 -18.63
[Fe](SnTe)eq (9) 97.88 -35.52
(53.59) (46.41) (69.77) (30.23)

Table 9. EDA for [Fe(CO)4(SnX)] → [Fe(CO)3(SnX)] and CO. (All the values are in kcal/mol)
b
Values in parentheses give the percentage contribution to the total attractive interactions ΔEelstat
+ ∆Eorb. c Values in parentheses give the contributions to the total orbital interaction ∆Eorb.
b b c c
Complex ∆Epauli ∆Eelstat ∆Eorb ΔEσ ΔEπ ∆Eint
[Fe](CO)ax (1)
-96.32 -86.94 -43.91 -43.02
130.47 -52.80
(52.56) (47.44) (50.51) (49.49)
[Fe](SnO)ax (2)
-111.47 -101.19 -51.27 -49.91
155.07 -57.59
(52.42) (47.58) (50.67) (49.33)
[Fe](SnS)ax (4)
-111.61 -100.35 -50.53 -49.81
155.24 -56.52
(52.66) (47.34) (50.36) (49.64)
[Fe](SnSe)ax (6) -111.88 -100.34 -50.28 -50.04
155.65 -56.57
(52.72) (47.28) (50.12) (49.88)
[Fe](SnTe)ax (8) -112.43 -99.99 -49.99 -49.98
156.73 -55.69
(52.93) (47.07) (50.01) (49.99)

In the current investigation, all axial isomers of [Fe(CO) 4(SnX)] have the carbonyl trans

to SnX. Using the EDA, we also looked at how the Fe-COax interactions change when SnX

replaces the CO ligand in the trans position. Table 9 displays the results of our research. In

comparison to the axial CO in [Fe(CO) 5], Table 9 shows that the axial CO bonds in

[Fe(CO)4(SnX)], which is trans to SnX, are both stronger and shorter. This is because, in

[Fe(CO)4(SnX)], “all the attractive contributions involved in the Fe-COax bond bonding are

stronger than in the well-known [Fe(CO)5].” The ligand SnX's effect on the Fe-COtrans bond is

independent of the ΔEσ/ΔEπ ratio. The [Fe]-CO trans bond is almost affected in the same way as

the [Fe]-SnX bond. Due to their stronger weaker bonds than CO (Table 8), the latter ligands form

stronger Fe-COtrans bonds in [Fe(CO)4(SnX)]. ΔEπ rises from ligands CO to SnX, as Table 8

demonstrates. The ΔEπ bonding contribution increases and the ΔEσ contribution decreases as X

changes from O to Te. The π-back-donation can be used to explain the longer C-O trans bond and

the shorter Fe-COtrans bond observed in the axial isomers of [Fe(CO) 4(SnX)]. It can also explain

the stronger Fe-COtrans bonds in these isomers.


4 Conclusion

The computed outcomes of this study are summed up as follows: The equatorial isomers

of [Fe(CO)4(SnX)] 2, 4, 6, and 8, are more stable than their same in axial counterparts. The

difference between these isomers are 6.0 kcal/mol only. The Fe-C bond in the axial position of

an equatorial isomer is shorter than in the equatorial bond. When the X atom is changed from O

to Te in the axial isomer, the bond is much longer than its equatorial position. The Fe-Sn bond

distance is also larger in the equatorial isomer than that of the axial isomer. In all of the

complexes of [Fe(CO)4(SnX)] 2-9, the Sn atom exhibits a positive charge, and the Fe atom has a

larger atomic charge, which is negative. The bond contribution from the Sn atom is greater than

that of the Fe atom, according to the NBO results for the Fe-Sn bond. These findings suggest the

polarization of Sn and Fe. The energy gaps of all the isomers are greater than 4.0 eV, according

to the FMO data. The SnS axial isomer exhibits the largest energy gap among these substituted

complexes, indicating higher stability and lower reactivity than the other complexes of the same.

The EDA results indicate that the σ contribution in the [Fe]-SnX bonding interaction is smaller

than σ. This results suggests that the σ-donation more strengthens the [Fe]-SnX bond more than

that of the σ-donation in [Fe(CO)5], where both σ-donation as well as π-back-donation. The

covalent bonding and electrostatic attraction ratios of the equatorial isomer of [Fe(CO) 4(SnX)]

are closer to those of [Fe(CO) 5]. Because of the π-back-donation, the SnX ligand's trans-ligand

effect strengthens the axial Fe-CO bond in the axial isomer.

DECLARATION OF COMPETING INTEREST


The authors declare that they have no competing interests.
5. References

1. Werner, H. (1990). Complexes of Carbon Monoxide and Its Relatives: An

Organometallic Family Celebrates Its Birthday. Angewandte Chemie International


Edition in English, 29(10), 1077–1089. https://doi.org/10.1002/anie.199010773

2. Davidson, E. R., Kunze, K. L., Machado, F. B. C., & Chakravorty, S. J. (1993). The

transition metal-carbonyl bond. Accounts of Chemical Research, 26(12), 628–635.

https://doi.org/10.1021/ar00036a004

3. Li, J., Schreckenbach, G., & Ziegler, T. (1994). First Bond Dissociation Energy of

M(CO)6 (M = Cr, Mo, W ) Revisited: The Performance of Density Functional Theory

and the Influence of Relativistic Effects. The Journal of Physical Chemistry, 98(18),

4838–4841. https://doi.org/10.1021/j100069a011

4. Li, J., Schreckenbach, G., & Ziegler, T. (1995). A Reassessment of the First Metal-

Carbonyl Dissociation Energy in M(CO)4 (M = Ni, Pd, Pt), M(CO)5 (M = Fe, Ru, Os),

and M(CO)6 (M = Cr, Mo, W) by a Quasirelativistic Density Functional Method. Journal

of the American Chemical Society, 117(1), 486–494.

https://doi.org/10.1021/ja00106a056

5. Jonas, V., & Thiel, W. (1995). Theoretical study of the vibrational spectra of the

transition metal carbonyls M(CO)6 [M=Cr, Mo, W], M(CO)5 [M=Fe, Ru, Os], and

M(CO)4 [M=Ni, Pd, Pt]. The Journal of Chemical Physics, 102(21), 8474–8484.

https://doi.org/10.1063/1.468839

6. Hoffmann, R. (1982). Building Bridges Between Inorganic and Organic Chemistry

(Nobel Lecture). Angewandte Chemie International Edition in English, 21(10), 711–724.

https://doi.org/10.1002/anie.198207113

7. Johnson, B. G., & Fisch, M. J. (1994). An implementation of analytic second derivatives

of the gradient-corrected density functional energy. The Journal of Chemical Physics,

100(10), 7429–7442. https://doi.org/10.1063/1.466887


8. Cable, J. W., & Sheline, R. K. (1956). Bond Hybridization And Structure In The Metal

Carbonyls. Chemical Reviews, 56(1), 1–26. https://doi.org/10.1021/cr50007a001

9. Elian, M., & Hoffmann, R. (1975). Bonding capabilities of transition metal carbonyl

fragments. Inorganic Chemistry, 14(5), 1058–1076. https://doi.org/10.1021/ic50147a021

10. Lewis, K. E., Golden, D. M., & Smith, G. P. (1984). Organometallic bond dissociation

energies: laser pyrolysis of iron pentacarbonyl, chromium hexacarbonyl, molybdenum

hexacarbonyl, and tungsten hexacarbonyl. Journal of the American Chemical Society,

106(14), 3905–3912. https://doi.org/10.1021/ja00326a004

11. F.A. Cotton, G. Wilkinson, C.A. Murillo, M. Bochmann, Advanced inorganic chemistry,

John Wiley & Sons; (1999).

12. Chi, K. Ming., Frerichs, S. R., Philson, S. B., & Ellis, J. E. (1988). Highly reduced

organometallics. 23. Synthesis, isolation, and characterization of

hexacarbonyltitanate(2-), (Ti(CO)62-). Titanium NMR spectra of carbonyltitanates.

Journal of the American Chemical Society, 110(1), 303–304.

https://doi.org/10.1021/ja00209a056

13. Persson, B. J., Roos, B. O., & Pierloot, K. (1994). A theoretical study of the chemical

bonding in M(CO) x (M=Cr, Fe, and Ni). The Journal of Chemical Physics, 101(8),

6810–6821. https://doi.org/10.1063/1.468309

14. Lefebvre, I., Szymanski, M. A., Olivier-Fourcade, J., & Jumas, J. C. (1998). Electronic

structure of tin monochalcogenides from SnO to SnTe. Physical Review B, 58(4), 1896–

1906. https://doi.org/10.1103/PhysRevB.58.1896

15. Jonas, V., & Thiel, W. (1999). Symmetry Force Fields for Neutral and Ionic Transition

Metal Carbonyl Complexes from Density Functional Theory. The Journal of Physical
Chemistry A, 103(10), 1381–1393. https://doi.org/10.1021/jp983600h

16. Parrish, S. H., R. J. Van Zee, and, & Weltner, W. (1999). Vanadium and Niobium

Hexadinitrogen and Hexacarbonyl Complexes: Electron-Spin-Resonance Spectra at 4 K.

The Journal of Physical Chemistry A, 103(8), 1025–1028.

https://doi.org/10.1021/jp983952o

17. Frenking, G., & Fröhlich, N. (2000). The Nature of the Bonding in Transition-Metal

Compounds. Chemical Reviews, 100(2), 717–774. https://doi.org/10.1021/cr980401l

18. Ellis, J. E. (2003). Metal Carbonyl Anions: from [Fe (CO) 4 ] 2 - to [Hf(CO) 6 ] 2 - and

Beyond. Organometallics, 22(17), 3322–3338. https://doi.org/10.1021/om030105l

19. Frenking, G., Wichmann, K., Fröhlich, N., Loschen, C., Lein, M., Frunzke, J., Rayón,

V.M., (2003) Towards a rigorously defined quantum chemical analysis of the chemical

bond in donor–acceptor complexes, Coord. Chem. Rev. 238,55-82.

https://doi.org/10.1016/S0010-8545(02)00285-0.

20. Pople, J. A., Gill, P. M. W., & Johnson, B. G. (1992). Kohn—Sham density-functional

theory within a finite basis set. Chemical Physics Letters, 199(6), 557–560.

https://doi.org/10.1016/0009-2614(92)85009-Y

21. Kohn, W., & Sham, L. J. (1965). Self-Consistent Equations Including Exchange and

Correlation Effects. Physical Review, 140(4A), A1133–A1138.

https://doi.org/10.1103/PhysRev.140.A1133

22. Wu, Q., & Yang, W. (2002). Empirical correction to density functional theory for van der

Waals interactions. The Journal of Chemical Physics, 116(2), 515–524.

https://doi.org/10.1063/1.1424928

23. Delley, B., Wrinn, M., & Lüthi, H. P. (1994). Binding energies, molecular structures, and
vibrational frequencies of transition metal carbonyls using density functional theory with

gradient corrections. The Journal of Chemical Physics, 100(8), 5785–5791.

https://doi.org/10.1063/1.467142

24. Delley, B., in Density Functional Theory: A Tool for Chemistry, edited by

J. M. Seminario and P. Politzer ~Elsevier, Amsterdam, 1994.

25. Vetri, P., Paularokiadoss, F., Celaya, C. A.,. Mary Novena, L., Thomas, J.M.,

Christopher Jeyakumar, T., (2023) A DFT study on structural and bonding analysis of

transition-metal carbonyls [M(CO)4] with terminal silicon chalcogenides complexes [M

(CO)3SiX] (M = Ni, Pd, and Pt; X = O, S, Se, and Te) Computational and Theoretical

Chemistry 1226, 114214. https://doi.org/10.1016/j.comptc.2023.114214

26. Priyadharsan, R. R., A. Timothy, R. A., Thomas, J.M., Christopher Jeyakumar, T.,

Rajaram, R., Louis, H., (2023) Investigating the structure, bonding, and energy

decomposition analysis of group 10 transition metal carbonyls with substituted terminal

germanium chalcogenides [M(CO)3GeX] (M=Ni, Pd, and Pt; X=O, S, Se, and Te)

complexes: insight from frst-principles calculations Journal of Molecular Modeling

29:344 https://doi.org/10.1007/s00894-023-05745-8

27. Christopher Jeyakumar, T., Mary Thomas, J., Sivan, A. K., & Sivasankar, C. (n.d.).

Molecular and electronic structure analysis of [Fe(CO)4(SiX)] (X = O, S, Se and Te): a

DFT study. https://doi.org/10.1007/s12039-022

28. Paularokiadoss, F., Christopher Jeyakumar, T., Thomas, R., Sekar, A., & Bhakiaraj, D.

(2022). Group 13 monohalides [AX (A = B, Al, Ga and In; X = Halogens)] as alternative

ligands for carbonyl in organometallics: Electronic structure and bonding analysis.

Computational and Theoretical Chemistry, 1209, 113587.


https://doi.org/10.1016/j.comptc.2021.113587

29. Paularokiadoss, F., Sekar, A., & Christopher Jeyakumar, T. (2020). A DFT study on

structural and bonding analysis of transition-metal carbonyls with terminal haloborylene

ligands [M(CO)3(BX)] (M = Ni, Pd, and Pt; X = F, Cl, Br, and I). Computational and

Theoretical Chemistry, 1177, 112750. https://doi.org/10.1016/j.comptc.2020.112750

30. Paularokiadoss, F., Alagan, S., & Christopher Jeyakumar, T. (2021). Coordination of

indium monohalide with group-10 metal carbonyls [TM(CO)3(InX)]: a DFT study.

Chemical Papers, 75(1), 311–324. https://doi.org/10.1007/s11696-020-01297-w

31. Paularokiadoss, F., Antony Sandosh, T., Sekar, A., & Christopher Jeyakumar, T. (2021).

Theoretical studies of group 10 metal gallylene complexes [TM(CO)3(GaX)].

Computational and Theoretical Chemistry, 1197, 113139.

https://doi.org/10.1016/j.comptc.2020.113139

32. Sekar, A., Paularokiadoss, F., Immanuel, S., & Christopher Jeyakumar, T. (2021).

Chemistry of group-10 metals monohaloalumylene complexes [TM(CO)3AlX]: a DFT

study. Theoretical Chemistry Accounts, 140(8), 101.

https://doi.org/10.1007/s00214-021-02801-5

33. Lee, C., Yang, W., & Parr, R. G. (1988). Development of the Colle-Salvetti correlation-

energy formula into a functional of the electron density. Physical Review B, 37(2),

785-789. https://doi.org/10.1103/PhysRevB.37.785

34. Oueslati, Y., Kansız, S., Dege, N., de la Torre Paredes, C., Llopis-Lorente, A., Martínez-

Máñez, R., & Sta, W. S. (2022). Growth, crystal structure, Hirshfeld surface analysis,

DFT studies, physicochemical characterization, and cytotoxicity assays of novel organic


triphosphate. Journal of Molecular Modeling, 28(3), 65. https://doi.org/10.1007/s00894-

022-05047-5.

35. Chiodo, S., Russo, N., & Sicilia, E. (2006). LANL2DZ basis sets recontracted in the

framework of density functional theory. The Journal of Chemical Physics, 125(10).

https://doi.org/10.1063/1.2345197

36. Hay, P. J., & Wadt, W. R. (1985). Ab initio effective core potentials for molecular

calculations. Potentials for the transition metal atoms Sc to Hg. The Journal of Chemical

Physics, 82(1), 270–283. https://doi.org/10.1063/1.448799

37. Wadt, W. R., & Hay, P. J. (1985). Ab initio effective core potentials for molecular

calculations. Potentials for main group elements Na to Bi. The Journal of Chemical

Physics, 82(1), 284–298. https://doi.org/10.1063/1.448800

38. Reed, A. E., Curtiss, L. A., & Weinhold, F. (1988). Intermolecular interactions from a

natural bond orbital, donor-acceptor viewpoint. Chemical Reviews, 88(6), 899–926.

https://doi.org/10.1021/cr00088a005

39. Van Lenthe, E., Ehlers, A., & Baerends, E.-J. (1999). Geometry optimizations in the

zero-order regular approximation for relativistic effects. The Journal of Chemical

Physics, 110(18), 8943–8953. https://doi.org/10.1063/1.478813

40. Dennington, R., Keith, T.A. and Millam, J.M. (2016) GaussView 6.0. 16. Semichem Inc.,

41. Richard, R. M., & Ball, D. W. (2008). G2, G3, and complete basis set calculations on the

thermodynamic properties of triazane. Journal of Molecular Modeling, 14(1), 29–37.

https://doi.org/10.1007/s00894-007-0247-y
42. Henderson, T. M., Janesko, B. G., & Scuseria, G. E. (2008). Generalized gradient

approximation model exchange holes for range-separated hybrids. The Journal of

Chemical Physics, 128(19). https://doi.org/10.1063/1.2921797

43. Maximoff, S. N., Ernzerhof, M., & Scuseria, G. E. (2004). Current-dependent extension

of the Perdew–Burke–Ernzerhof exchange-correlation functional. The Journal of

Chemical Physics, 120(5), 2105–2109. https://doi.org/10.1063/1.1634553

44. Frenking, G., & Matthias Bickelhaupt, F. (2014). The EDA Perspective of Chemical

Bonding. In The Chemical Bond (pp. 121–157). Wiley.

https://doi.org/10.1002/9783527664696.ch4

45. Vyboishchikov, S. F., Krapp, A., & Frenking, G. (2008). Two complementary molecular

energy decomposition schemes: The Mayer and Ziegler–Rauk methods in comparison.

The Journal of Chemical Physics, 129 (14). https://doi.org/10.1063/1.2989805

You might also like