Jump to content

Rep:Mod:1CSam11

From ChemWiki
                                                       ==Computational Lab Experiment 1C==



[1]

Part 1

Thermodynamic Analysis of Cyclopentadiene Dimerisation and Hydrogenation Products

  • Molecule 1

molecule 1 minimised using MMFF94s force field

Type of Energy Contribution Energy (kcal/mol) Energy (kJ/mol)
Total Electrostatic Energy 13.01365
Total Van der Waal's Energy 3.54286
Total Out-of-Plane Bending Energy 30.77264
Total Torsional Energy 2.04136
Total Stretch Bending Energy 2.04136
Total Angle Bending Energy 0.01486
Total Bond Stretching Energy 12.80121
Total Energy 55.37344 231.838

Molecule 2

molecule 2 minimised using MMFF94s force field

Type of Energy Contribution Energy (kcal/mol) Energy (kJ/mol)
Total Electrostatic Energy 58.1907 243.633
Total Van der Waal's Energy 3.46742
Total Out-of-Plane Bending Energy 33.19047
Total Torsional Energy -2.08217
Total Stretch Bending Energy -2.94986
Total Angle Bending Energy 0.02197
Total Bond Stretching Energy 12.35798
Total Energy 14.18488

Molecules 1 and 2 have no partial charges. Of the reactant molecules, number 2, the endo Diels-Alder product, is the most unstable. The Van der Waal’s energy of the endo product is lower than number 1, the exo product, due to two favourable interactions available in the endo product, allowed by virtue of the reduced distance of the overhanging methylene bridging proton two protons of the cyclo-pentene ring. This is illustrated in the figure below:

  1. Molecule 1 (Exo Product)
Exo-product
Exo-product
Van der Waal's Interaction Interaction Length (Å)
H1-H5 2.632
H2-H5 3.448
H3-H5 2.667
H4-H5 2.668
H3-H6 2.622
H4-H6 2.624
  1. Molecule 2 (Endo-product)
Endo-product
Endo-product
Van der Waal's Interaction Interaction Length (Å)
H1-H5 2.541
H2-H5 2.667
H3-H5 2.574
H4-H5 2.667
H4-H6 2.624
H2-H6 2.623

Torsional and stretch bending energies of the endo product are also lower than for the exo product. Torsional strain is reduced in the endo product due to the gauche-antiperiplanar arrangement of alkyl substituents on bonds C1-C2 and C3-C4, which allows for increased σ-hyperconjugation and reduced steric clashing c.f the gauche-gauche conformation in the exo product. All other energy sources of the endo product are higher. Thus the dimerization is kinetically controlled, as molecule 2, the endo product is favoured over the lower total energy, more stable exo product. This agrees with literature computational modelling reports[2] and experimental observations.[3]

  • Molecule 3

molecule 3 minimised using MMFF94s force field

Type of Energy Contribution Energy (kcal/mol) Energy (kJ/mol)
Total Electrostatic Energy 50.72283 212.366
Total Van der Waal's Energy 3.3081
Total Out-of-Plane Bending Energy 30.85901
Total Torsional Energy -1.92668
Total Stretch Bending Energy 0.06933
Total Angle Bending Energy 0.01531
Total Bond Stretching Energy 13.27678
Total Energy 5.12098

Important electrostatic interaction lengths between the nucleophilic allylic carbons and electrophilic overhanging hydrogens are noted below, alongside important H-H Van der Waal's interaction lengths. Numbering of interaction lengths is made with respect to the figure below.

Molecule 3 Carbon and Hydrogen Interaction Lengths
Molecule 3 Carbon and Hydrogen Interaction Lengths
C-H Electrostatic Interaction Interaction Lengths (Å)
C6-H9 2.589
C7-H9 2.589
H6-H11 2.803
H7-H11 2.803
H-H Van der Waal's Interaction Interaction Lengths (Å)
H1-H2 2.29
H3-H4 2.29
H1-H8 2.502
H4-H5 2.502
H4-H10 2.666
H1-H10 2.542
H7-H9 3.423
H6-H9 3.423
  • Molecule 4

molecule 4 minimised using MMFF94s force field

Type of Energy Contribution Energy (kcal/mol) Energy (kJ/mol)
Total Electrostatic Energy 41.25749 172.737
Total Van der Waal's Energy 2.82311
Total Out-of-Plane Bending Energy 24.68536
Total Torsional Energy -1.65719
Total Stretch Bending Energy -0.37839
Total Angle Bending Energy 0.00028
Total Bond Stretching Energy 10.6373
Total Energy 5.14702

Important electrostatic interaction lengths between the nucleophilic allylic carbons and electrophilic overhanging hydrogens are noted below, alongside important H-H Van der Waal's interaction lengths. Numbering of interaction lengths is made with respect to the figure below.

Molecule 4 Carbon and Hydrogen Interaction Lengths
Molecule 4 Carbon and Hydrogen Interaction Lengths
Electrostatic Interaction Interaction Distance (Å)
C1-H4 2.666
C2-H3 3.206
C2-H4 3.159
C1-H3 3.471
H-H Van der Waal's Interaction Interaction Lengths (Å)
H3/4-H5 2.285
H1-H6 2.674
H5-H14 2.459
H5-H13 2.518
H6-H13 2.632
H8-H12 2.609
H10-H12 2.572


The thermodynamic product of hydrogenation is dimer 4, which has a lower total energy. Van der Waal’s energy for product 4 is lower as only one H-H steric interaction lies within the repulsive region (<2.4 A, [2] and experimental observations.[2]), c.f two steric clashes for product 3. Bond stretch, angle bending, out of plane bending and torsional energies are lower for product 4. Bend stretching energy is lower for product 3. Torsional energy is higher for dimer 3 as an eclipsed arrangement is forced upon the allyl group, with both allylic hydrogens arranged syn-periplanar with the X and Y hydrogens. In molecule 4, the hydrogens of the allylic group are in a staggered conformation with the bonded carbon hydrogens, which reduces steric clashing and increases σ-hyperconjugation, so lowering strain. These effects lower the bond stretch, angle bending and out of plane bending energies as less steric clashing is exhibited. Electrostatic energy of molecule 3 is lower than for molecule 4 as the distance between the partially negative, nucleophilic allylic carbons and the electrophilic overhanging hydrogens is reduced, so allowing for an increased interaction in molecule 3.

Conformations of an Intermediate in Taxol Synthesis

For both the carbonyl-up and the carbonyl-down atropisomer, four conformational isomers exist, which can be interconverted through a series of bond rotations. These conformational isomers occur due to the stereochemistry of the C ring, which has a cyclohexane structure and four energy minima: chair 1, twist-boat 1, twist-boat 2, chair 2 which are allowed by distortions of the cycloheptane B ring. The cycloheptane B ring has greater strain than cyclohexane C ring, partly due to transannular crowding (steric hindrance by groups on opposite sides of the ring). Thus to find the lowest energy conformation, the cyclohexane ring geometry need only be changed.
Considering only the cyclohexane ring, the planar structure was not trialled as it is well reported that with bond angles of 120º, (10.5º larger than the ideal angle of tetrahedral geometry) angle strain is too high. Additionally, every C-C bond in this structure would be eclipsed, resulting in high eclipsing strains and a highly destabilised structure. Of the reported energetically allowed confromers of the cyclohexane ring: chair; twist-chair; twist-boat and boat structures were trialled.
The highest energy conformer is reported to be the twist-chair form, which could not be modelled for either the atropisomers due to the global energy maximum lying at this geometry, between the two more favourable of the chair (a global minimum) and the twist-boat (local minimum) conformers.
The cyclohexane boat structure has two eclipsed bonds, with severed steric crowding of the "bow" and "stern" hydrogen atoms. As a result of these destabilising effects, which result in the bot form lying at the top of a local energy maxima between the two twist-boat froms, the cyclohexane C ring boat structure could not be found in either the carbonyl-up or carbonyl-down atropisomers.
By twisting the boat conformation, the steric hindrance of the boat is somewhat relieved. This allowed the two twist-boat conformers to be modelled for both atropisomers.
Lifting one carbon atom above the ring plane, and moving the other below ring planereduces the conformer energy further. This staggered structure conformer, in chair conformation, exists with the lowest energy, and is well-reported to exist as the dominant structure adopted by cyclohexane molecules. (REFERENCES: ORGANIC CHEM)


Carbonyl-down Atropisomers

Conformational Isomer Conformer 1 Conformer 2 Conformer 3 Conformer 4 Conformer 5
Structure Energy Contribution (kcal/mol) Structure Energy Contribution (kcal/mol) Structure Energy Contribution (kcal/mol) Structure Energy Contribution (kcal/mol) Structure Energy Contribution (kcal/mol)
B ring eclipsed, C ring chair 1
B ring eclipsed, C ring chair 60.55459 kcal/mol
B ring eclipsed, C ring chair 60.55459 kcal/mol

B ring eclipsed, C ring chair 60.55459 kcal/mol

Total Electrostatic Energy = -0.05161

Total Van der Waals = 33.22746
Total Out-of-Plane Bending Energy = 0.85124
Total Torsional Energy = 0.26085
Total Stretch Bending Energy = -0.14424
Total Angle Bending Energy = 18.81474
Total Bond Stetching Energy = 7.59616
Total Energy = 60.55459

B ring staggered, C ring chair 2 B ring staggered, C ring chair 74.97578 kcal/mol Total Electrostatic Energy = 0.47721

Total Van der Waals = 35.62172
Total Out-of-Plane Bending Energy = 1.59809
Total Torsional Energy = 7.92190
Total Stretch Bending Energy = -0.21690
Total Angle Bending Energy = 21.18430
Total Bond Stetching Energy = 8.38944
Total Energy = 74.97578

B ring staggered, C ring chair 74.89187 kcal/mol Total Electrostatic Energy = 0.45760

Total Van der Waals = 35.71917
Total Out-of-Plane Bending Energy = 1.61037
Total Torsional Energy = 7.60033
Total Stretch Bending Energy = -0.21462
Total Angle Bending Energy = 21.25908
Total Bond Stetching Energy = 8.45994
Total Energy = 74.89187

B ring staggered, C ring chair 70.70427 kcal/mol Total Electrostatic Energy = 0.23104

Total Van der Waals = 35.22782
Total Out-of-Plane Bending Energy = 0.92455
Total Torsional Energy = -0.85820
Total Stretch Bending Energy = -0.15877
Total Angle Bending Energy = 27.10501
Total Bond Stretching Energy = 8.23281
Total Energy = 70.70427

B ring staggered, C ring chair 70.73128 kcal/mol Total Electrostatic Energy = 0.22888

Total Van der Waals = 35.25229
Total Out-of-Plane Bending Energy = 0.92546
Total Torsional Energy = -0.92439
Total Stretch Bending Energy = -0.15213
Total Angle Bending Energy = 27.17449
Total Bond Stretching Energy = 8.22669
Total Energy = 70.73128

B ring staggered, C ring chair 70.67903 kcal/mol Total Electrostatic Energy = 0.22795

Total Van der Waals = 35.27567
Total Out-of-Plane Bending Energy = 0.95955
Total Torsional Energy = -0.75685
Total Stretch Bending Energy = -0.16890
Total Angle Bending Energy = 26.99255
Total Bond Stretching Energy = 8.14905
Total Energy = 70.67903

B ring staggered, C ring twist-boat 1 B ring staggered, C ring twist boat 138.37778 kcal/mol Total Electrostatic Energy = 1.91522

Total Van der Waals = 52.42341
Total Out-of-Plane Bending Energy = 2.76891
Total Torsional Energy = 7.67376
Total Stretch Bending Energy = -0.51417
Total Angle Bending Energy = 60.24160
Total Bond Stretching Energy = 13.86905
Total Energy = 138.37778

B ring staggered, C ring twist boat 138.43313 kcal/mol Total Electrostatic Energy = 1.91704

Total Van der Waals = 52.14369
Total Out-of-Plane Bending Energy = 2.78293
Total Torsional Energy = 7.76965
Total Stretch Bending Energy = -0.49889
Total Angle Bending Energy = 60.52090
Total Bond Stretching Energy = 13.79780
Total Energy = 138.43313

B ring staggered, C ring twist boat 138.43466 kcal/mol Total Electrostatic Energy = 1.91217

Total Van der Waals = 52.10441
Total Out-of-Plane Bending Energy = 2.73095
Total Torsional Energy = 7.59818
Total Stretch Bending Energy = -0.50455
Total Angle Bending Energy = 60.85561
Total Bond Stretching Energy = 13.73789
Total Energy = 138.43466

B ring staggered, C ring twist boat 65.82924 kcal/mol Total Electrostatic Energy = 0.09374

Total Van der Waals = 35.38455
Total Out-of-Plane Bending Energy = 0.70011
Total Torsional Energy = 4.45351
Total Stretch Bending Energy = -0.02810
Total Angle Bending Energy = 17.43881
Total Bond Stretching Energy = 7.78661
Total Energy = 65.82924

B ring eclipsed, C ring twist-boat 2 B ring eclipsed, C ring higher energy twist-boat 75.03719 kcal/mol Total Electrostatic Energy = 0.40144

Total Van der Waals = 35.79491
Total Out-of-Plane Bending Energy = 1.55287
Total Torsional Energy = 7.44715
Total Stretch Bending Energy = -0.20440
Total Angle Bending Energy = 21.52222
Total Bond Stretching Energy = 8.52300
Total Energy = 75.03719

B ring eclipsed, C ring higher energy twist-boat 68.88626 kcal/mol Total Electrostatic Energy = -0.04523

Total Van der Waals = 34.66701
Total Out-of-Plane Bending Energy = 0.95110
Total Torsional Energy = 4.45092
Total Stretch Bending Energy = -0.06447
Total Angle Bending Energy = 21.08743
Total Bond Stretching Energy = 7.83950
Total Energy = 68.88626

B ring eclipsed, C ring higher energy twist-boat 66.32321 kcal/mol Total Electrostatic Energy = -0.04537

Total Van der Waals = 35.08423
Total Out-of-Plane Bending Energy = 0.84920
Total Torsional Energy = 3.17749
Total Stretch Bending Energy = -0.05922
Total Angle Bending Energy = 19.57742
Total Bond Stretching Energy = 7.73947
Total Energy = 66.32321

The conformers, in order of increasing energy and decreasing stability are: chair 1, twist-boat 1, twist-boat 2, chair 2. The lowest energy conformer of the carbonyl-down atropisomer exists as the B ring eclipsed, C ring chair conformation with total energy of 60.55459 kcal/mol. Differences in energy between the optimised structures for each specified conformational isomer is mostly due to the cycloheptane B ring geometry, which can change dramatically with even minute geometry differences of a specified cyclohexane C ring conformer. This is due to increased complexity of the cycloheptane ring, allowing a wide array of geometries lying between two extremes of all-staggered C-C bonds and all-eclipsed C-C bonds. For simplicity here, all structures that did not appear with all-eclipsed bonds were tabulated with B ring staggered bonds.
The lowest energy carbonyl-down confomer exists with a cyclohexane C ring chair geometry, agreeing with literature. (REFERENCE ORGANIC CHEMISTRY) Of the two chair forms, termed chair 1 and chair 2, the lowest energy conformer, chair 1, has an eclipsed B ring structure, c.f. the staggered B ring of the higher energy conformer. This indicates that eclipsing interactions are not the only factor that needs to be considered, as an increase in eclipsing interactions should result in a higher energy conformer. All conformers have a strained gauche arrangement around the cyclohexane ring bridging C-C bond. However, the chair 1 form has a more regular gauche arrangement, as can be seen through dihedral angle analysis. The selected dihedral angles here are those involving the two ring bridging hydrogens, and are as summarised below.

Conformational Isomer Torsion Angle H1-C-C-H2 (°) Torsion Angle H1-C-C-C1 (°) Torsion Angle H1-C-C-C2 (°) Torsion Angle H2-C-C-C3 (°) Torsion Angle H2-C-C-C4 (°)
B ring eclipsed, C ring chair 1 48.5 67 167 68.1 163.5
B ring staggered, C ring chair 2 52.9 62.4 168.1 59.2 174.1
B ring eclipsed, C ring twist-boat 1 33.2 88.3 153.5 90.5 35
B ring staggered, C ring twist-boat 2 28.9 86.6 147.4 86.3 142

The two chair forms give a more regular gauche arrangement than the twist boat forms, allowing for increasing overlap of σ-σ* orbitals in stabilising hyperconjugation. This effect thus contributes to the increased stabilisation of chair form 1.
Hyperconjugation interactions are summarised in the table below.

Conformational Isomer (C-C-(R) σ -> H-C σ*) (C-H σ -> C-C-(R) σ*) (C-H σ -> C-H σ*) (C-C σ -> C-C σ*) (C-C-(OR) σ -> C-H σ*) (H-C σ -> C-C-(OR) σ*) (C-C-(OR) σ -> C-C σ*) (C-C σ -> C-C-(OR) σ*)
B ring eclipsed, C ring chair 1 1 1 0 2 1 1 0 O
B ring staggered, C ring chair 2 2 2 0 0 0 0 1 1
B ring eclipsed, C ring twist-boat 1 0 0 2 2 0 0 1 1
B ring staggered, C ring twist-boat 2 1 1 0 2 1 1 0 0

As C is more electronegative than H (χP 2.5 c.f 2.1 respectively), so the C–C σ* orbital is lower in energy than the C–H σ* orbital, giving a slightly more favourable hyperconjugative interaction between the C–H σ bond and the C–C σ* orbital compared to that between the C–H σ bond and the C–H σ* orbital (1.5 kcal/mol vs. 1.0 kcal/mol). [1] This effect is increased for the C-C-(R)-> C-C-(OR) σ* interaction, and even more so for the C-C-(OR) σ* -> C-C-(H) σ conjugation due to the increased χP of oxygen relative to carbon and hydrogen. The chair 1 form benefits from both additional stabilisation from the C-(H)σ -> C-C-(OR) σ* and C-H σ -> C-C σ* hyperconjugations, which contribute to the stabilisation of this conformer. The chair 2 conformer benefits from C-H σ -> C-C-(R) σ* and C-C σ -> C-C-(OR) σ* interactions, whilst the twist-boat 1 conformer benefits from C-C σ -> C-C-(OR) σ* conjugations and twist-boat 2 benefits from H-C σ -> C-C-(OR) σ* and C-H σ -> C-C-(R) σ* hyperconjugations. Considering only hyperconjugation effects, the twist-boat 1 would be expected to be less stable than chair 2 and twist-boat 2, with chair-2 being less stable than twist-boat 2. Put simply, using the hyperconjugation analysis, the conformers in theoretical order of increasing energy are: chair 1, twist-boat 2, chair 2, twist-boat 1.
Thus, as the observed energy order differs from that obtained through analysis of equatorial effects (hyperconjugation interactions) it is evident that significant steric hindrance occurs between the B ring carbonyl group and the downward projecting carbon atom of the chair 2 and twist-boat 2 conformers. A 2.757 A close contact interaction exists between the carbonyl-down oxygen and the downward projecting carbon of the chair 2 cylohexane ring. This steric interaction is not present for the chair 1, B eclipsed ring form, as the cyclohexane chair form present here is the mirror image of that present in the B staggered ring isomer, and so the equivalent cyclohexane carbon atom projects upwards. The favourable reduction in the carbonyl-cyclohexane chair steric interaction outweighs the destabilising eclipse strain present in the B-ring, and this is reinforced through analysis of the chair 2 and twist-boat conformers. Both twist-boat conformers would be expected to give a higher energy than both chair forms, considering cyclohexane effects only. However, as is tabulated above, both twist-boats give a lower, more stabilised energy than the chair 2 isomer. This indicates that the carbonyl-cyclohexane repulsive steric clash is the dominant repulsive interaction in the molecule, and is alleviated in the twist-boat forms due to increases in the oxygen-carbon distance by the rotation of the downward projecting cyclohexane carbon atom. The reduction in steric hindrance is greater than the increase in energy obtained from conformation conversion of the C ring from chair to twist-boat.
Twist-boat 2 is higher in energy than twist-boat 1 as, though the principal carbonyl-carbon steric repulsion of chair 2 form is reduced in both twist-boat isomers, twist-boat 2 results in a downward projecting carbon atom, at a distance of 3.128 A from the carbonyl oxygen. In twist boat 1, this carbon projects upwards, and the interaction is reduced due to the longer 4.232 A distance.
The B ring eclipsed, chair 1 form will be used in the following dithiol experimentation as this conformation is the most stable, and therefore will be observed as the predominant conformation. Additionally, the total taxol geometry of the lowest energy conformer is reinforced through comparison with literature, as this conformation was observed as the lowest energy conformation for a dithiol, methylated taxol derivative. [4]

ChemBio3D Optimisations

As a comparative aside, the twist-boat chair structures were trialled in ChemBio3D using the MMFF94 force field, a maximum of 500 iterations and an RMS gradient of 0.00000001. Results are as tabulated below.

Conformational Isomer Conformer 1 Conformer 2
Structure Energy Contribution (kcal/mol) Structure Energy Contribution (kcal/mol)
B ring eclipsed, C ring chair 1 B ring eclipsed, C ring chair 1 61.0278 kcal/mol Total Energy = 61.0278 B ring eclipsed, C ring chair 1 60.5445 kcal/mol Total Energy = 60.5445
B ring staggered, C ring chair 2 B ring staggered, C ring chair 2 70.6275 kcal/mol Total Energy = 70.6275
B ring eclipsed, C ring twist-boat 2 B ring eclisped, C ring twist-boat 2 66.3100 kcal/mol Total Energy = 66.3100

As for the minimisations performed in Avogadro, the B ring eclipsed, C ring chair 1 returns the lowest total energy, with a 0.00009 kcal/mol difference between the two programs and MMFF94s and MMFF94 force fields. The fit between the twist-boat 2 forms and the chair 2 forms using MMFF94 and MMF94s is similar, with differences of 0.01321 and 0.05153 kcal/mol respectively. All forms return lower energy using the MMFF94 force field. However, the energy distribution between the conformers is the same, with the chair 2 conformer is returned with higher energy than the twist-boat conformer with the MMFF94s force field, reinforcing the results obtained with the MMFF94s force field using Avogadro. This reinforces that the steric interaction between the carbonyl oxygen and the cyclohexane downward projecting carbon in chair form 2 is more repulsive than eclipsing interactions in the B ring obtained through alleviation of steric hindrance by upward rotation of the carbon atom to chair form, and to a reduced effect in the twist-boat 2 form.

Carbonyl-up Atropisomers

Carbonyl-up atropisomer conformations were trialed in the same method as the carbonyl-down atropisomer. The Taxol carbonyl-up atropisomer was modelled with each of the four cyclohexane ring conformations and the total energy minimised using the MMFF94s force field with conjugate gradients in Avogadro. To decrease the energy of each conformer further, atomic geometries were changed slightly prior to further minimisations using the MMF94s force field. The table below summarises a selection of the conformations trialed for each conformational isomer, with the lowest energy geometry for each isomer tabulated.

Conformational Isomer Conformer 1 Conformer 2 Conformer 3 Conformer 4 Conformer 5 Conformer 6
Structure Energy Contribution (kcal/mol) Structure Energy Contribution (kcal/mol) Structure Energy Contribution (kcal/mol) Structure Energy Contribution (kcal/mol) Structure Energy Contribution (kcal/mol)
B ring staggered, C ring chair 2 B ring staggered, C ring chair 131.2788231 kcal/mol Total Electrostatic Energy = 1.62706

Total Van der Waals = 47.53841
Total Out-of-Plane Bending Energy = 2.40406
Total Torsional Energy = 6.67894
Total Stretch Bending Energy = -0.53556
Total Angle Bending Energy = 61.26476
Total Bond Stetching Energy = 12.30116
Total Energy = 131.2788231

B ring staggered, C ring chair 131.25834 kcal/mol Total Electrostatic Energy = 1.59637

Total Van der Waals = 47.29274
Total Out-of-Plane Bending Energy = 2.38051
Total Torsional Energy = 6.46826
Total Stretch Bending Energy = -0.55717
Total Angle Bending Energy = 61.77266
Total Bond Stetching Energy = 12.30497
Total Energy = 131.25834

B ring staggered, C ring chair 131.26806 kcal/mol Total Electrostatic Energy = 1.61176

Total Van der Waals = 47.44958
Total Out-of-Plane Bending Energy = 2.38811
Total Torsional Energy = 6.56152
Total Stretch Bending Energy = -0.54461
Total Angle Bending Energy = 61.50072
Total Bond Stetching Energy = 12.30099
Total Energy = 131.26806

B ring staggered, C ring chair 2 131.07731 kcal/mol Total Electrostatic Energy = 1.50782

Total Van der Waals =46.74049
Total Out-of-Plane Bending Energy = 2.24378
Total Torsional Energy = 5.90459
Total Stretch Bending Energy = -0.56858
Total Angle Bending Energy = 62.89435
Total Bond Stetching Energy = 12.35487
Total Energy = 131.07731

B ring eclipsed, C ring chair 1 B ring eclipsed, C ring chair 1 123.41854 kcal/mol Total Electrostatic Energy = 1.63300

Total Van der Waals = 44.46712
Total Out-of-Plane Bending Energy = 2.52423
Total Torsional Energy = 4.62418
Total Stretch Bending Energy = -0.16983
Total Angle Bending Energy = 57.91489
Total Bond Stetching Energy = 12.42495
Total Energy = 123.41854

B ring eclipsed, C ring chair 1 123.38535 kcal/mol Total Electrostatic Energy = 1.63578

Total Van der Waals = 44.47899
Total Out-of-Plane Bending Energy = 2.53372
Total Torsional Energy = 4.70558
Total Stretch Bending Energy = -0.18453
Total Angle Bending Energy = 57.75445
Total Bond Stetching Energy = 12.46136
Total Energy = 123.38535

B ring eclipsed, C ring chair 1 123.37353 kcal/mol Total Electrostatic Energy = 1.63629

Total Van der Waals = 44.50357
Total Out-of-Plane Bending Energy = 2.53751
Total Torsional Energy = 4.70787
Total Stretch Bending Energy = -0.17295
Total Angle Bending Energy = 57.69574
Total Bond Stetching Energy = 12.46551
Total Energy = 123.37353

B ring eclipsed, C ring chair 1 123.32868 kcal/mol Total Electrostatic Energy = 1.63859

Total Van der Waals = 44.54910
Total Out-of-Plane Bending Energy = 2.56324
Total Torsional Energy = 4.95789
Total Stretch Bending Energy = -0.19997
Total Angle Bending Energy = 57.32752
Total Bond Stetching Energy = 12.52280
Total Energy = 123.32868

B ring eclipsed, C ring chair 1 123.32349 kcal/mol Total Electrostatic Energy = 1.63955

Total Van der Waals = 44.57888
Total Out-of-Plane Bending Energy = 2.56823
Total Torsional Energy = 5.00947
Total Stretch Bending Energy = -0.20268
Total Angle Bending Energy = 57.23473
Total Bond Stetching Energy = 12.49530
Total Energy = 123.32349

B ring eclipsed, C ring chair 1 123.32320 kcal/mol Total Electrostatic Energy = 1.63764

Total Van der Waals = 44.58146
Total Out-of-Plane Bending Energy = 2.56702
Total Torsional Energy = 4.99945
Total Stretch Bending Energy = -0.19687
Total Angle Bending Energy = 57.22484
Total Bond Stetching Energy = 12.50967
Total Energy = 123.32320

B ring eclipsed, C ring twist-boat 1 B ring eclipsed, C ring twist-boat 1 128.48151 kcal/mol Total Electrostatic Energy = 1.66793

Total Van der Waals = 45.98552
Total Out-of-Plane Bending Energy = 2.53493
Total Torsional Energy = 7.45904
Total Stretch Bending Energy = -0.21728
Total Angle Bending Energy = 58.39814
Total Bond Stetching Energy = 12.65323
Total Energy = 128.48151

B ring eclipsed, C ring twist-boat 1 128.47502 kcal/mol Total Electrostatic Energy = 1.65517

Total Van der Waals = 46.00742
Total Out-of-Plane Bending Energy = 2.53554
Total Torsional Energy = 7.39507
Total Stretch Bending Energy = -0.53549
Total Angle Bending Energy = 58.49646
Total Bond Stetching Energy = 12.58139
Total Energy = 128.47502

B ring eclipsed, C ring twist-boat 1 128.43682 kcal/mol Total Electrostatic Energy =1.65472

Total Van der Waals = 46.03398
Total Out-of-Plane Bending Energy = 2.55088
Total Torsional Energy = 7.48060
Total Stretch Bending Energy = -0.19954
Total Angle Bending Energy = 58.31709
Total Bond Stetching Energy = 12.59909
Total Energy = 128.43682

B ring eclipsed, C ring twist-boat 1 128.43442 kcal/mol Total Electrostatic Energy = 1.65733

Total Van der Waals = 46.03176
Total Out-of-Plane Bending Energy = 2.55578
Total Torsional Energy = 7.52936
Total Stretch Bending Energy = -0.20264
Total Angle Bending Energy = 58.20822
Total Bond Stetching Energy = 12.65461
Total Energy = 128.43442

B ring eclipsed, C ring twist-boat 1 128.40590 kcal/mol Total Electrostatic Energy = 1.65531

Total Van der Waals = 46.02818
Total Out-of-Plane Bending Energy = 2.59326
Total Torsional Energy = 7.92157
Total Stretch Bending Energy = -0.23094
Total Angle Bending Energy = 57.76893
Total Bond Stetching Energy = 12.67129
Total Energy = 128.40590

B ring staggered, C ring twist boat 2 B ring staggered, C ring twist-boat 135.38221 kcal/mol Total Electrostatic Energy = 1.61088

Total Van der Waals = 49.00052
Total Out-of-Plane Bending Energy = 2.28097
Total Torsional Energy = 9.14482
Total Stretch Bending Energy = -0.39447
Total Angle Bending Energy = 61.50182
Total Bond Stetching Energy = 12.23766
Total Energy = 135.38221

B ring staggered, C ring twist-boat 135.36156 kcal/mol Total Electrostatic Energy = 1.60889

Total Van der Waals = 49.13997
Total Out-of-Plane Bending Energy = 2.24812
Total Torsional Energy = 9.04908
Total Stretch Bending Energy = -0.37416
Total Angle Bending Energy = 61.42617
Total Bond Stetching Energy = 12.26349
Total Energy = 135.36156

B ring staggered, C ring twist-boat 2 131.26806 kcal/mol Total Electrostatic Energy = 1.61029

Total Van der Waals = 47.53394
Total Out-of-Plane Bending Energy = 2.36877
Total Torsional Energy = 6.51641
Total Stretch Bending Energy = -0.53549
Total Angle Bending Energy = 61.45503
Total Bond Stetching Energy = 12.32172
Total Energy = 131.27067

B ring staggered, C ring twist-boat 2 131.27409 kcal/mol Total Electrostatic Energy = 1.62127

Total Van der Waals = 47.46346
Total Out-of-Plane Bending Energy = 2.40461
Total Torsional Energy = 6.64622
Total Stretch Bending Energy = -0.54107
Total Angle Bending Energy = 61.37870
Total Bond Stetching Energy = 12.30090
Total Energy = 131.27409

B ring staggered, C ring twist-boat 2 131.27882 kcal/mol Total Electrostatic Energy = 1.62706

Total Van der Waals = 47.53841
Total Out-of-Plane Bending Energy = 2.40406
Total Torsional Energy = 6.67894
Total Stretch Bending Energy = -0.53556
Total Angle Bending Energy = 61.26476
Total Bond Stetching Energy = 12.30116
Total Energy = 131.27882

B ring staggered, C ring twist-boat 2 131.27441 kcal/mol Total Electrostatic Energy = 1.62048

Total Van der Waals = 47.42292
Total Out-of-Plane Bending Energy = 2.40991
Total Torsional Energy = 6.65155
Total Stretch Bending Energy = -0.54446
Total Angle Bending Energy = 61.41411
Total Bond Stetching Energy = 12.29989
Total Energy = 131.27441

As indicated above, the conformations, in order of increasing energy are as follows:chair 1, twist-boat 1, chair 2, twist-boat 2. This complements the result of the carbonyl-up atropisomer, as for both configurations, the B ring eclipsed, C ring chair conformational isomer exists with lowest energy. As for the carbonyl-up atropisomer, this occurs through the combination of numerous effects, whihc include hyperconjugation and steric hindrance of the electron dense carbonyl oxygen. In order to rationalise electronic stabilisation by hyperconjugation, the summary table is shown below.

Conformational Isomer (C-C-(R) σ -> H-C σ*) (C-H σ -> C-C-(R) σ*) (C-H σ -> C-H σ*) (C-C σ -> C-C σ*) (C-C-(OR) σ -> C-H σ*) (H-C σ -> C-C-(OR) σ*) (C-C-(OR) σ -> C-C σ*) (C-C σ -> C-C-(OR) σ*)
B ring eclipsed, C ring chair 1 1 1 0 2 1 1 0 0
B ring staggered, C ring chair 2 2 2 0 0 0 0 1 1
B ring eclipsed, C ring twist-boat 1 1 1 0 2 1 1 0 0
B ring staggered, C ring twist-boat 2 2 2 0 0 0 0 1 1

Considering hyperconjugation effects only, the chair 1 and twist-boat 1 conformers would be expected to lie at lower energy and be mores stable than the chair 2 and twist-boat 2 conformers. This is due to the chair and twist-boat 1 conformers obtaining the highly stabilising H-C σ -> C-C-(OR) σ* interaction, compared to the C-C σ -> C-C-(OR) σ* hyperconjugation present in the chair and twist-boat 2 conformers, which exerts a reduced stabilision due to the higher energy C-C σ orbital c.f. the C-H σ orbital, giving reduced overlap, and so weaker orbital interaction and reduced splitting energy with the low energy C-C-(OR) σ* orbital.
Further disparity between the conformers is partially due to steric interations with the bulky, electron dense carbonyl group of the B ring with the C ring carbons. These carbons are bulkier and more electronically dense than the ring hydrogens, which thus do not exert steric repulsions with the carbonyl electron cloud with as great a magnitude as do the carbon atoms. Steric clashing is reduced in the chair 1 isomer, as the eclipsed B ring geometry allows the cyclohexane ring to lie below the plane of the carbonyl, and so the upward projecting cyclohexane carbon, at 3.075 A, is too far away from the carbonyl oxygen to provide any substantial repulsion, whilst the upward projecting B ring carbon, at 4.556 A also does not provide significant steric hindrance. The twist boat 1 form exerts a roughly equivalent steric repulsion through the upward projecting B ring and C ring atoms lying at 4.551 A and 3.056 A from the carbonyl oxygen respectively. The carbonyl steric repulsions present in chair 2 form are increased compared to those present in the chair 1 and twist-boat 1 forms, due to an increase in the number of upward projecting carbon atoms in the B and C rings at close proximity to the carbonyl oxygen. These interactions are roughly equivalent to those present in twist-boat 2 form. In chair 2 form, the C and B ring upward projecting carbons lie at 4.522 and 3.323 A from the carbonyl oxygen. c.f the shorter 4.419 and 3.421 A distances between oxygen and the C and B ring carbons, present in twist-boat 1.
Thus, as both steric effects and hyperconjugation interactions give a roughly equivalent energy for the chair 1, twist-boat 1 and chair 2, twist-boat 2 pairs, the increased stabilisation of each chair form relative to the twist-boat conformers results from the inherent stability of the chair conformation, by virtue of reduced eclipsing interactions and increased σ-σ* overlap, allowing for stronger hyperconjugation interactions.

Comparison of Carbonyl-Up and Carbonyl-Down Atropisomers

The lower energy conformation is the carbonyl-down atropisomer, which at 60.55459 kcal/mol is almost twice as stable as the lowest energy 123.3232 kcal/mol conformation of the carbonyl up atropisomer. Both structures feature a chair C ring, which is energetically favourable and locked into place by the centrally positioned bridge E-cyclononenone , which features heightened non-bonded trans-annular interactions than for the staggered B ring conformers, as has been reported. [5] The lower energy of the carbonyl-down isomer arises from the vast reduction in carbonyl oxygen cyclohexane carbon steric hindrance, which cannot be achieved in the carbonyl-up configuration. As thermodynamically, the final molecule in the oxy-cope equilibrium (carbonyl-up atropisomer) is higher and more unstable than the reagent and first carbonyl-down atropisomer product, this allows for the establishment of an observable equilibrium by the principle of microscopic reversibility.[5] The carbonyl-down atropisomer is an example of a hyperstable olefin, a set of olefins which are stabilised rather than destabilised because of their bridgehead location and are unreactive. Poly-cyclic rings magnify the effects of hyperstability due to the stability gained by the olefin cage structure and the parent poly-cycloalkane being more strained.[6] A selection of successive energy mimisations, achieved using the MMFF94s force field and conjugate gradient algorithm in Avogadro, of the cage structure in question, a bicyclo [6.2.1]undec-1-ene, is tabulated below, and reveals that Oxy-Cope Rearrangement stabilises the molecule from 85.54366 kcal/mol to 60.53455 kcal/mol at the MMFF94s level of theory. The lowest obtained pre-Oxy-Cope geometry and the stabilisation gained upon Oxy-Cope Rearrangement are in accordance with results documented by Paquette et al.[5]

Taxol Pre-oxy Cope Rearrangement
Conformer 1 Conformer 2 Conformer 3
Structure Energy Contribution (kcal/mol) Structure Energy Contribution (kcal/mol) Structure Energy Contribution (kcal/mol)
Taxol pre-Oxy Cope Rearrangement minimised to 94.80472 kcal/mol Total Electrostatic Energy = 19.29532

Total Van der Waals = 38.64676
Total Out-of-Plane Bending Energy = 0.04770
Total Torsional Energy = -3.11166
Total Stretch Bending Energy = -1.45139
Total Angle Bending Energy = 29.29113
Total Bond Stetching Energy = 12.08687
Total Energy = 94.80472

Taxol pre-Oxy Cope Rearrangement minimised to 85.54372 kcal/mol Total Electrostatic Energy = 16.34218

Total Van der Waals = 37.35729
Total Out-of-Plane Bending Energy = 0.03619
Total Torsional Energy = -3.77616
Total Stretch Bending Energy = -1.47038
Total Angle Bending Energy = 26.31645
Total Bond Stetching Energy = 10.73814
Total Energy = 85.54372

Taxol pre-Oxy Cope Rearrangement minimised to 85.54366 kcal/mol Total Electrostatic Energy = 16.34231

Total Van der Waals = 37.35759
Total Out-of-Plane Bending Energy = 0.03606
Total Torsional Energy = -3.77637
Total Stretch Bending Energy = -1.46992
Total Angle Bending Energy = 26.31525
Total Bond Stretching Energy = 10.73874
Total Energy = 85.54366

As is the requirement for hyperstable olefins, the parent alkane of the carbonyl-down atropisomer, at 69.5350 kcal/mol using the MMFF94s force field, is higher in total energy than the carbonyl-down olefin, at 60.53445 kcal/mol using the same level of theory and force field.[6]

Taxol Hydrogenation Product
Conformer 1 Conformer 2 Conformer 3
Structure Energy Contribution (kcal/mol) Structure Energy Contribution (kcal/mol) Structure Energy Contribution (kcal/mol)
Taxol Hydrogenation Product minimised to 69.54399 kcal/mol Total Electrostatic Energy = 0.00000

Total Van der Waals = 31.24484
Total Out-of-Plane Bending Energy = 0.03592
Total Torsional Energy = 9.27390
Total Stretch Bending Energy = 0.29445
Total Angle Bending Energy = 22.29051
Total Bond Stetching Energy = 6.40437
Total Energy = 69.54399

Taxol Hydrogenation Product minimised to 69.54357 kcal/mol Total Electrostatic Energy = 0.00000

Total Van der Waals = 31.21907
Total Out-of-Plane Bending Energy = 0.03469
Total Torsional Energy = 9.18404
Total Stretch Bending Energy = 0.29428
Total Angle Bending Energy = 22.40441
Total Bond Stretching Energy = 6.40707
Total Energy = 69.54357

Taxol Hydrogenation Product minimised to 69.53530 kcal/mol Total Electrostatic Energy = 0.00000

Total Van der Waals = 31.28088
Total Out-of-Plane Bending Energy = 0.03615
Total Torsional Energy = 9.20654
Total Stretch Bending Energy = 0.29431
Total Angle Bending Energy = 22.29783
Total Bond Stetching Energy = 6.41958
Total Energy = 69.53530

This reinforces the stability of the olefin and the elongated time required for the oxy-cope rearrangement, the product of which is higher in energy than the parent carbonyl-down olefin, as documented above. [5]

Dithiol, Methylated Derivative of Taxol

The lowest energy conformer for taxol, the carbonyl-down, B=eclipsed C=chair isomer, was used as a "starting block" for the dithiol, methylated taxol derivative, drawn using the appropriate tools in Avogadro and minimised using the MMFF94s force field with conjugate gradients in Avogadro. This gave two conformers based on the slope of the dithiol ethylene group. Results are noted below.

Conformational Isomer Conformer 1 Conformer 2 Conformer 3
Structure Energy Contribution (kcal/mol) Structure Energy Contribution (kcal/mol) Structure Energy Contribution (kcal/mol)
B ring eclipsed, C ring chair, dithiol slope right to left taxol dithiol derivative, carbonyl down, eclipsed, chair, higher energy dithiol conformer Total Electrostatic Energy = -6.01552

Total Van der Waals = 49.26889
Total Out-of-Plane Bending Energy = 0.89120
Total Torsional Energy = 9.65702
Total Stretch Bending Energy = 0.65716
Total Angle Bending Energy = 31.08724
Total Bond Stetching Energy = 14.96419
Total Energy = 100.51019

taxol dithiol derivative, carbonyl down, eclipsed, chair, lower energy dithiol conformer 100.49638 kcal/mol Total Electrostatic Energy = -6.04364

Total Van der Waals = 49.35210
Total Out-of-Plane Bending Energy = 0.89469
Total Torsional Energy = 9.65981
Total Stretch Bending Energy = 0.65065
Total Angle Bending Energy = 30.98398
Total Bond Stetching Energy = 14.99878
Total Energy = 100.49638

Total Electronic Energy = 5.07403

Total Van der Waals = 19.97645
Total Out-of-Plane Bending Energy = 0.02967
Total Torsional Energy = 1.01202
Total Stretch Bending Energy = -0.80469
Total Angle Bending Energy = 2.61375
Total Bond Stetching Energy = 2.78242
Total Energy = 30.68364

B ring eclipsed, C ring chair, dithiol slope left to right Dithiol taxol derivative, B ring eclipsed, C ring chair, dithiol slope left to right 100.45058 kcal/mol Total Electronic Energy = -6.01409

Total Van der Waals = 49.27260
Total Out-of-Plane Bending Energy = 0.88841
Total Torsional Energy = 9.65287
Total Stretch Bending Energy = 0.63018
Total Angle Bending Energy = 30.94874
Total Bond Stretching Energy = 15.07185
Total Energy = 100.45058

Total Electronic Energy = 5.15300

Total Van der Waals = 19.37922
Total Out-of-Plane Bending Energy = 0.00939
Total Torsional Energy = 1.67403
Total Stretch Bending Energy = -0.86951
Total Angle Bending Energy = 2.18029
Total Bond Stretching Energy = 2.69798
Total Energy = 30.22440


This was optimised further to give a lower energy conformation of this isomer. taxol dithiol derivative, carbonyl down, eclipsed, chair, higher energy dithiol conformer

Type of Energy Contribution Energy (kcal/mol)
Total Electrostatic Energy -6.00757
Total Van der Waal's Energy 38.80345
Total Out-of-Plane Bending Energy 0.88053
Total Torsional Energy 9.67023
Total Stretch Bending Energy 0.67389
Total Angle Bending Energy 30.97644
Total Bond Stretching Energy 14.98488
Total Energy

  • Second Conformer (lower energy)

The lower energy conformer has a geometry in agreement with literature. [1]


NEED DATA FOR THIS

This structure was optimised to a lower energy conformer.

taxol dithiol derivative, carbonyl down, eclipsed, chair, lower energy dithiol conformer

Type of Energy Contribution Energy (kcal/mol)
Total Electrostatic Energy -6.01409
Total Van der Waal's Energy 38.75358
Total Out-of-Plane Bending Energy 0.88842
Total Torsional Energy 9.65288
Total Stretch Bending Energy 0.63018
Total Angle Bending Energy 30.9487
Total Bond Stretching Energy 15.07189
Total Energy 89.93156

To check that the B=envelope, C=chair conformer was a minimum for the dithiol, methylated taxol derivative, a number of other isomers were tested, and documented below.

Conformational Isomer Conformer 1 Conformer 2 Conformer 3
Structure Energy Contribution (kcal/mol) Structure Energy Contribution (kcal/mol) Structure Energy Contribution (kcal/mol)
B ring staggered, C ring chair, dithiol slope right to left taxol dithiol derivative, carbonyl down, eclipsed, twist-boat, dithiol conformer 102.35566 kcal/mol Total Electrostatic Energy = -7.63857

Total Van der Waals = 51.72894
Total Out-of-Plane Bending Energy = 1.23175
Total Torsional Energy = 10.70883
Total Stretch Bending Energy = 0.01332
Total Angle Bending Energy = 31.07099
Total Bond Stetching Energy = 15.24039
Total Energy = 102.35566

B ring eclipsed, C ring boat, dithiol slope left to right taxol dithiol derivative, B ring eclipsed, C ring boat, dithiol left to right 102.28425 kcal/mol Total Electrostatic Energy = -6.24643

Total Van der Waals = 50.49824
Total Out-of-Plane Bending Energy = 1.02534
Total Torsional Energy = 13.59904
Total Stretch Bending Energy = 0.48338
Total Angle Bending Energy = 28.52666
Total Bond Stetching Energy = 14.39801
Total Energy = 102.28425

B ring eclipsed, C ring twist-boat, dithiol slope left to right B ring eclipsed, C ring twist-boat, dithiol left to right 102.58219 kcal/mol Total Electrostatic Energy = -8.72989

Total Van der Waals = 52.55376
Total Out-of-Plane Bending Energy = 0.98243
Total Torsional Energy = 13.09340
Total Stretch Bending Energy = 0.31291
Total Angle Bending Energy = 28.88162
Total Bond Stetching Energy = 15.48796
Total Energy = 102.58219

B ring eclipsed, C ring twist-boat, dithiol slope left to right B ring eclipsed, C ring twist-boat, dithiol left to right 102.31850 kcal/mol Total Electrostatic Energy = -8.72989

Total Van der Waals = 52.55376
Total Out-of-Plane Bending Energy = 0.98243
Total Torsional Energy = 13.09340
Total Stretch Bending Energy = 0.31291
Total Angle Bending Energy = 28.88162
Total Bond Stetching Energy = 15.48796
Total Energy = 102.31850



NMR Analysis of dithiol Taxol Derivative

Modelling the Sulphur bonded Carbon correction constant

The carbon shift correction for the ketone functionality was calculated using the provided formula in the script. The carbon shift correction for the sulphur bonded carbon was calculated using the method proposed by Braddock et al.[7]

This involved using dimethylsulphide and methane as model NMR complexes, optimising these structures using the MMFF94s force field with conjugate gradient algorithm in Avodagro to total energies of 3.23095 and 0.02638 kcal/mol respectively, which are shown below.

DMS total energy 3.23095 kcal/mol

Methane total energy 0.02638 kcal/mol

These optimised structures were used as input to optimisation and NMR simulation calculations at the DFT B3LYP 6-31G(d,p) level of theory in Gaussian09. This returned optimised DMS and methane structures as shown below, where DMS numbering is documented.

Numbering of DMS used in NMR calculations

DMS optimised to DFT B3LYP 6-31G(d,p) level of theory Methane optimised to DFT B3LYP 6-31G(d,p) level of theory The DMS and Methane 13C NMR spectra of the optimised DMS and methane structures are as shown.

DMS 13C NMR spectrum
Methane 13C NMR spectrum

The carbon shift in dimethylsulphide was taken, and from this the methane carbon shift was subtracted, giving the sulphur correction. Parameters involved in this calculation are tabulated below.

Dimethylsulphide Methane δcorr (ppm)
Carbon Number δcalc (ppm) Carbon Number δcalc (ppm)
C1 21.59386267 C1 -3.842204108 25.43606571
C2 21.59386054
Average 21.5938616
Analysis of Taxol Derivative NMR calculations

An initial NMR at B3LYP 6-31G(d) level of theory was run to compare with that obtained using B3LYP 6-31G(d,p) level of theory. This gave a structure of energy -1651.84055438 a.u, minimised to an RMS Gradient Norm = 0.00000896 a.u. and dipole moment of 1.8244 Debye.

Taxol,B ring eclipsed, chair 1 lowest energy conformer optimised to DFT B3LYP 6-31G(d) level of theory

The numbering of this structure used in 13C and 1H NMR spectra analysis is shown below.

Atom numbering used in 13C and 1H NMR spectrum analysis

Carbon and Hydrogen shifts are as noted.

13Carbon NMR 1H NMR
Carbon Number δcalc (ppm) δcorr (ppm) Hydrogen Number δcalc (ppm)
C1 26.161857 H20 5.3808011259
C2 29.86709953 H18 2.1739491359
C3 147.5631791 H22 1.3884441820
C4 55.88038826 H23 2.0328648211
C5 39.98869248 H24 2.1101413932
C6 212.0205765 H25 1.5686530476
C7 55.208629 H26 1.5772279533
C8 121.0582038 H27 2.2212727976
C9 49.97235076 H28 1.1796658407
C10 24.56279499 H29 0.8114765715
C11 28.49392089 H30 0.5602421860
C12 33.98456806 H31 1.1097325593
C13 66.72226844 H32 0.5843857641
C14 92.90788938 H33 1.0914291282
C15 42.8563047 H34 0.5171291442
C16 25.65950574 H35 2.2921989860
C17 49.53367818 H36 2.3678804139
C19 35.59996883 H37 1.1055670964
C44 46.94962623 H38 0.8801533872
C45 45.28853047 H39 0.9134712328
H40 1.9115711768
H41 1.9390786429
H42 1.5109610099
H47 2.4266532587
H49 2.3604387856
H50 2.5928863775
H51 1.7256286136
H52 0.8754000000
H53 0.2364156789

The hydrogen shifts are more removed from the experimental values than the carbon shifts. L. Paquette, N. A. Pegg, D. Toops, G. D. Maynard, R. D. Rogers, J. Am. Chem. Soc.,, 1990, 112, 277-283. DOI:10.1021/ja00157a043 This is as expected as the two p-orbitals on hydrogen have not been taken into account at this level of theory.

The same molecule was entered into the NMR calculation again, using the expanded DFT B3LYP 6-31G(d,p) force field and basis set. These results, when compared with the 6-31G(d) basis set calculations gave a closer fit to the literature values, L. Paquette, N. A. Pegg, D. Toops, G. D. Maynard, R. D. Rogers, J. Am. Chem. Soc.,, 1990, 112, 277-283. DOI:10.1021/ja00157a043 and those carbons bonded to sulphur and oxygen atoms were entered into sulphur and oxygen corrections. This gave results as follows.

13Carbon NMR
Carbon Number δcalc (ppm) δcorr (0.96δcalc + 12.2) (ppm) Sulphur Correction (ppm) δcorr (ppm) ∆δ(ppm) Average ∆δ (ppm)
C1 24.45394017 24.45394017 2.243940174 4.20520568
C2 28.35963979 28.35963979 2.799639787
C3 147.8656084 147.8656084 -0.854391587
C4 54.93343559 54.93343559 3.633435588
C5 38.5101642 38.5101642 1.730164202
C6 211.9222905 215.6453988 215.6453988 4.155398844
C7 54.75635105 54.75635105 3.45635105
C9 49.5311262 49.5311262 6.251126196
C10 22.58014307 22.58014307 2.750143067
C11 26.50329918 26.50329918 1.153299184
C12 32.4674164 32.4674164 2.467416396
C13 65.93908338 65.93908338 5.40908338
C14 92.84098207 50.87213142 143.7131135 22.8131135
C15 41.47012634 41.47012634 2.740126336
C16 24.00755679 24.00755679 2.617556794
C17 48.03676504 48.03676504 7.216765045
C18 33.69603749 33.69603749 2.856037493
C19 33.69603749 33.69603749 -1.773962507
C44 45.64885345 25.43606571 71.08491916 3.525080837
C44 44.00772411 25.43606571 69.44378983 8.913789825
Difference Table

Part 2

Crystallographic Data

The structure of 21 (Shi asymmetric Fructose Catalyst), synthesised according to the scheme below

D-Fructose Shi catalyst synthesis scheme
was found in the Crystallographic Database using conquest, returning NELQEA

[8] and NELQEA01[9] structures, which were analysed using Mercury. An anomeric centre is a stereocentre resulting from intramolecular acetal (or ketal) formation of an aldehyde (or ketone) group and a hydroxyl group. Two stereochemistries at the anomeric centre are allowed, with each anomer being a diastereoisomers of the other.[10] The α-anomer, with the anomeric substituent axial on the ring is stable and can exist as the dominant isomer over the β-isomer (anomeric substituent equatorial), depending on the other ring substituents. Considering steric arguments only, this is not the expected dominance as in chair conformation, bulky axial substituents cause destabilisation due to unfavourable 1,3-diaxial interactions. The anomeric effect may occur by stabilisation of the axial –OR via donation into the C-O σ* (antibonding) orbital by the lone pair of the pyranose ring oxygen. This hyperconjugation, a stereoelectronic effect acts against, and is observed over steric effects, which render axial -OR stereochemistry as repulsive. The anomeric effect increases in strength with the electronegativity of the anomeric substituent, thus for the OR group, the effect would be expected to be relatively strong in the Shi catalyst. It is reported that the anomeric effect is stronger in pyranose.[11]

For analysis, all C-O bond lengths of the three acetal centres of the Shi catalyst were analysed. The pyranose ring acetal centre displayed bond lengths with a higher equivalence, whilst the acetal centres associated with the 2-dimethyl-(1,3)-dioxolane ring displayed the classical electronic stabilisation of the anomeric centre. This results in one C-O bond being significantly shorter than the other, due to the increase in bond strength of the ring C-O bond associated with π-like overlap of the oxygen p-orbital and the C-O σ*, and the decrease in bonding character of the C-O bond outside the ring, associated with electron donation from the oxygen into the C-O σ*.[11]
The Shi fructose catalyst shows high enantioselectivity for the epoxidation of tri- and trans-substituted olefins. It has been reported that the pyranose oxygen is important for catalytic activity as electronically it inductively activates the carbonyl group to oxone attack due to the high electronegativity of the pyranose oxygen. [9] This effect is reinforced through analysis of the the pyranose ring C-O and 2,2-dimethyl-(1,3)-dioxolane ring C-O bond lengths, which have both axial orientations at the acetal carbon centre, so allowing for the double anomeric effect, where both C-O bonds are stabilised by the anomeric effect. [11]Assuming a double anomeric effect, with equivalent electron donation by each C-O axial lone pair into C-O σ*, one would expect equivalent bond lengths, as observed for molecule 1 in the NELQEA structure. However, NELQEA01 gives a longer 2,2-dimethyl-(1,3)-dioxolane ring C-O σ bond length, which may be indicative of the reported inductive electron withdrawal from the ketone by the pyranose oxygen. This effect increases the electron density of the pyranose C-O bond, increasing the bond strength and reducing bond length. An additional interaction between the ketone C=O π-cloud and 2,2-dimethyl-(1,3)-dioxolane ring C-O σ* bond may also be present, which may reduce electron density at the ketone further through the favourable π-like resonance effect between the C-O σ*-cloud and the C=O π-cloud, which is not present with the pyranose ring C-O bond.

The anomeric O-C-O bond lengths for each structure are noted below.

NELQEA

NELQEA crystal structure of Shi catalyst
NELQEA crystal structure of Shi catalyst

Shi catalyst crystal structure NELQEA

Molecule 1 '
Anomeric Centre bonds Bond Length (A)
C7-01 1.413
C7-O2 1.441
C10-O5 1.409
C10-O4 1.439
Molecule 2
Anomeric Centre Bonds
C19-O7 1.38
C19-O8 1.443
C22-O11 1.416
C22-O10 1.438

Bond angles of the anomeric centres are as noted below.

Bond angle Angle (theta)
Molecule 1
O1-C7-O2 103.2
O5-C10-O4 105.6
Molecule 2
O7-C19-O8 105.3
O10-C22-O11 105.1

Non-anomeric bonds and angles were analysed as below.

Non-anomeric bond lengths Bond Length (A)
Molecule 1
C2-O6 1.403
C2-O2 1.403
Molecule 2
C14-O12 1.408
C14-O8 1.39
Bond angle Angle (theta)
Molecule 1
O2-C2-O6 113.03
Molecule 2
O8-C14-O12 112.48

NELQEA01

This structure was analysed in the same way, with bond lengths and angles as tabulated below.

NELQEA crystal structure of Shi catalyst

Shi catalyst crystal structure NELQEA01

'
Molecule 1
Anomeric Centre bonds Bond Length (A)
C10-O5 1.428
C10-O4 1.456
C9-O1 1.423
C9-O2 1.454
Molecule 2
C22-O11 1.43
C22-O10 1.442
C21-O8 1.46
C21-O7 1.406
Bond angle Angle (theta)
Molecule 1
O4-C10-O5 105.8
O1-C9-O2 103.6
Molecule 2
O11-C22-O10 105.7
O7-C21-O8 103

Non-anomeric bond lengths and angles are as noted in the following table.

Non-anomeric bond lengths Bond Length (A)
Molecule 1
C2-O6 1.415
C2-O2 1.423
Molecule 2
C14-O8 1.42
C14-O12 1.405
Bond angle Angle (theta)
Molecule 1
O8-C14-O12 113.26
Molecule 2
O6-C2-O2 112.78

The ketals are at both α-positions, and so act as chiral controls and activating groups. These ketals are sterically bulky enough and in close vicinity to the carbonyl group, allowing for high enantiomeric excess and reaction promotion at the carbonyl group. The lack of any R groups on the 5-membered rings reduces sterics around the carbonyl, allowing for oxone nucleophilic attack. [12]


Molecule 23,[Mn(III)(salen)(Cl)],(1R,2R)-(-)-[Bis(3,5-di-tert-butylsalicylidene)-1,2-cyclohexanediamine]chloromanganese(III), An Intermediate in Jacobsen's Catalyst Synthesis

Two structures of the intermediate ([Mn(III)(salen)(Cl)]) to Jacobsen's catalyst were obtained, TOVNIB01 and TOVNIB02. These are as reported below.

TOVNIB01

TOVNIB01 crystal structure of the Jacobsen catalyst. Blue lines represent intramolecular close contacts and red lines represent to intermolecular close contacts
TOVNIB01 space fill crystal structure of the Jacobsen catalyst.

Jacobsen catalyst crystal structure TOVNIB01

TOVNIB02

TOVNIB02 space fill crystal structure of the Jacobsen catalyst.
TOVNIB02 crystal structure of the Jacobsen catalyst. Blue lines represent intramolecular close contacts and red lines represent to intermolecular close contacts

Jacobsen catalyst crystal structure TOVNIB02


Jacobsen’s ligand, (R,R)- or (S,S)-N,N′-bis(3,5-di-tert-butylsalicylidene)-1,2-cyclohexanediamine, (salen) contains a trans-cyclohexane-1,2-diamine linkage as the sole chiral unit. Salicylidene rings are each bonded to two sterically bulky tert-butyl groups. The enantiopure C2 symmetrical trans-cyclohexane-1,2-diamine serves as a sterodirecting group, creating chiral centres in substrate molecules. Previous X-ray crystal structural studies have shown that metal complexes of the salen ligand exist in an almost planar structure, with the tert-butyl groups of the salicylidene rings not being sufficiently sterically demanding to cause N2O2 coordination distortion. It has been reported that small deviations from planarity may however be required to predetermine the course of asymmetric reactions. [10] As is apparent from the crystal structures TOVNIB01 [13] and TOVNIB02[14] documented, the N2O2 geometry exists with near planarity, indicating that the tert-butyl groups do not have sufficient steric bulk to cause distortion to a stepped geometry, as has been observed for similar Ni(II) complexes.[10] Analysing the close contacts of the tert-butyl groups with the two Mn(III) coordinated oxygens reveals that four interactions are made between each tert-butyl group and each oxygen. Two close contact interactions are made with attack from the chlorine side of the N2O2 coordination plane, with two close contact interactions between each tert-butyl and the oxygen from the under-side of this coordination plane. The tert-butyl groups have sufficient steric bulk to form four close contacts each with one phenyl functionality.


TOVNIB01 ' ' ' ' ' TOVNIB02 ' ' ' ' '
tBu-tBu Interactions Attack from Cl-side Attack from under Cl-side
Intramolecular Interactions Interaction Distance (A) Intramolecular Interactions Average Interaction Distance (A) Intramolecular Interactions Interaction Distance (A) Intramolecular Interactions Average Interaction Distance (A) Intramolecular Interactions Average Interaction Distance (A)
TOVNIB01 (C-H, underside of Cl) 2.895 tBu(H)-O 2.193 tBu(C)-O 2.884 tBu(H)-O 2.55795 tBu(C)-O 3.017
TOVNIB02 (H-H, underside of Cl) 2.328 tBu(H)-O 2.335 tBu(C)-O 2.909 tBu(H)-O 2.497 tBu(C)-O 3.119

Analysis of the Transition State for Shi Catalyst Catalysed Epoxidation of Styrene

To find the styrene transition state in the Shi catalyst catalysed asymmetric epoxidation, the Gibbs free energy of the four transition states given for the R and S-isomers were analysed. This was carried out by accessing the .log file of each transition state structure, where the Gibbs Free Energy was found as the "Sum of Electronic and Thermal Free Energies". Results of the four R and four S-styrene oxide transition states are noted below.

R-styrene series ' ' S-styrene series ' '
Conformer DOI Gibbs Free Energy (Hartree/Particle) Gibbs Free Energy +Thermal Correction to Gibbs Free Energy(Hartree/Particle) Conformer DOI Gibbs Free Energy (Hartree/Particle) Gibbs Free Energy + Thermal Correction to Gibbs Free Energy (Hartree/Particle)
10.6084/m9.figshare.822152 -1303.730703 -1303.34381 10.6084/m9.figshare.823545 -1303.73383 -1303.348909
10.6084/m9.figshare.828520 -1303.730238 -1303.344628 10.6084/m9.figshare.822136 -1303.72418 -1303.338169
10.6084/m9.figshare.822135 -1303.736813 -1303.35075 10.6084/m9.figshare.828519 -1303.72767 -1303.340216
10.6084/m9.figshare.822137 1303.738044 1304.124393 10.6084/m9.figshare.826003 -1303.73850 -1303.353713

This data reveals the structure with the lowest Gibb's Free Energy exists in the R-styrene oxide conformation, which was submitted for NCI and QTAIM analysis.

NCI and QTAIM Analysis of the S-styrene series

Non-covalent interactions refer to intra- and inter-molecular interactions that involve dispersed electromagnetic interactions and do not involve the sharing of electrons. [15] Such interactions are of the order of 1-5 kcal/mol. [15] Non-covalent interactions can be classified into four main categories: Van der Waals forces, π-effects, electrostatic,and hydrophobic effects. For the transition state of the Shi catalyst epoxidation of styrene, the R-styrene oxide conformation provides the lowest Gibb's Free Energy, a result supported by literature observation. [16] [17] [17] [17] Simply, this is due to non-covalent interactions in the transition state, which are greatest for the R-isomer product than for the S-isomer products, leading to the R-isomer pathway having a lower activation energy than the S-isomer pathway, which results in the R-isomer pathway being favoured and so having greater enantiomeric excess experimentally with the Shi catalyst. Both R and S-series transition states operate with a spiro conformation.

Spiro vs. planar epoxidation transition state

As illustrated below, the spiro A structure is observed for the transition states.

Spiro A vs. Spiro B epoxidation transition state


This result complies with literature, and this is favoured over the planar transition state due to the stabilising interaction of the dioxirane oxygen lone pair with the olefin pi* orbital. [17]


The S-styrene series NCI and QTAIM analysis results are as shown below. SM11 S isomer of styrene.cub.xyz

NCI analysis of the transition state for the Shi catalyst catalysed epoxidation towards the S-styrene oxide product 10.6084/m9.figshare.826003
NCI analysis of the transition state for the Shi catalyst catalysed epoxidation towards the S-styrene oxide product 10.6084/m9.figshare.826003

[image:Screenshot NCI 3.PNG|centre|200px|frame|NCI analysis of the transition state for the Shi catalyst catalysed epoxidation towards the S-styrene oxide product 10.6084/m9.figshare.826003]]

NCI analysis of the transition state for the Shi catalyst catalysed epoxidation towards the S-styrene oxide product 10.6084/m9.figshare.826003

The S-styrene series transition state corresponding to 10.6084/m9.figshare.82600 of Gibbs Free Energy (without thermal correction constant) of -1303.73850 Hartree/Particle was also submitted to QTAIM analysis. Results are as below.

[[image:Qtaim 1.png|centre|200px|frame|Shi catalyst catalysed epoxidation of styrene 10.6084/m9.figshare.826003 S-series]

Shi catalyst catalysed epoxidation of styrene 10.6084/m9.figshare.826003 S-series
Shi catalyst catalysed epoxidation of styrene 10.6084/m9.figshare.826003 S-series
Shi catalyst catalysed epoxidation of styrene 10.6084/m9.figshare.826003 S-series

Shi catalyst catalysed epoxidation of styrene 10.6084/m9.figshare.826003 S-series

QTAIM and NCI analysis revealed the spiro transition state structure. This structure indicates significant C-O bond formation between the oxirane and C1 of the styrene ethyl substituent, which lies at an interaction distance of 1.969 Å, and as revealed by the NCI analysis, can be considered as "half-covalent". This spatial separation is greater than the C-O bond length of the oxirane, which lies at a distance of 1.414 Å and indicates incomplete product C-O bond formation, a conclusion reinforced by the product C-O bond lying at a distance greater than both the epoxide and the hydroxyl C-O typical bond length. The interaction between the β-carbon and the oxirane oxygen occurs over an increased distance of 2.349 Å, and the terminal hydrogen makes an interaction with the ketal oxygen with an interaction distance of 2.489 Å. Additional interactions between the ketal oxygen and the hydrogen at the β-carbon position lies at 2.572 Å. The ketal oxygen makes further interactions with the benzyl ring carbons with strongest interaction distances for each oxygen lying at 3.729 and 5.769 Å. The NCI analysis indicates two major attractive non-covalent interactions. One is intramolecular, and occurs between the eclipsed bonds of the benzene ring and the alkene C=C bond of styrene. The otehr is intermolecular and occurs between the alkene C=C bond of styrene and the oxygens of the dioxirane and the dioxacyclopentane ring which projects closest to the styrene C=C alkene bond.

R-styrene series QTAIM Lowest Energy Transition State

The lowest energy transition state of the Shi catalyst catalysed epoxidation reaction of styrene corresponded to the R-styrene oxide product pathway and was analysed by QTAIM analysis, which revealed the spiro transition state structure, as can be seen in the figures below.

Shi catalyst catalysed epoxidation of styrene lowest energy transition state R-series
Shi catalyst catalysed epoxidation of styrene lowest energy transition state R-series
Shi catalyst catalysed epoxidation of styrene lowest energy transition state R-series
Shi catalyst catalysed epoxidation of styrene lowest energy transition state R-series
Shi catalyst catalysed epoxidation of styrene lowest energy transition state R-series
Shi catalyst catalysed epoxidation of styrene lowest energy transition state R-series

Shi catalyst catalysed epoxidation of styrene lowest energy transition state R-series

This structure indicates significant C-O bond formation between the oxirane and C1 of the styrene ethyl substituent, which lies at an interaction distance of 1.998 Å. This spatial separation is greater than the C-O bond length of the oxirane, which lies at a distance of 1.423 Å and indicates incomplete product C-O bond formation on styrene but significant C-O bond breaking at the dioxirane, a conclusion reinforced by the product C-O bond lying at a distance greater than both the epoxide and the hydroxyl C-O typical bond length. The interaction between the β-carbon and the oxirane oxygen occurs over an increased distance of 2.349 Å, and the terminal hydrogen makes an interaction with the two ketal oxygens with interaction distances of 2.780 and 2.506 Å. Additional interactions between the ketal oxygen and the benzyl ring hydrogen lies at 2.915 and 4.194 Å. The hydrogen of the dioxacyclopentane interacts with the benzyl carbon at an interaction distance of 2.457 Å. The ketal oxygens make further interactions with the benzyl ring carbons with the strongest interaction distances for each oxygen lying at 3.491 and 5.523 Å, and it is these interactions, caused by better overlap of the benzyl ring with the dioxacyclopentane Shi catalyst ring which causes the R-styrene oxide product conformation to be favoured. The S-styrene oxide product pathway causes a significant reduction in the interactions between the benzyl functionality and the ketal oxygens due to the increase in spatial separation, thus causing the higher energies of the transition states for the S-styrene oxide product pathways. In summary, the R-styrene oxide product pathway is favoured as the transition state allows for reduced steric clashing between the bulky phenyl styrene ring and the Shi catalyst dioxacyclopentane ring at the 2-position on the pyranose ring. This also allows for increased interactions between the phenyl ring of styrene and the oxygens of the second dioxacyclopentane functionality at the 4-position on the pyranose ring of the Shi catalyst. Drawing from the NCI analysis of the higher energy S-styrene transition state, the favourable interactions here would be expected to arise between the alkene C=C bond and the oxygens of the dioxirane and dioxacyclopentane, and between the alkene C=C bond and the eclipsed benzene bonds. As the eclipsed benzene-alkene C=C interactions are available to the oxygens of the dioxirane and the dioxacyclopentane, this allows for an increased favourability of interaction for the R-series as the benzene ring is closer to the oxygens of the Shi catalyst, allowing for extended benzene pi-conjugation with the dioxirane and the dioxacyclopentane orbitals of the Shi catalyst. This conclusion is in accord with literature, which reports an attractive interaction between the oxazolidinone moiety of the Shi catalyst and styrene benzyl-π group in the transition state giving the spiro A transition state.[18] Such non-covalent dispersive interactions have been previously reported to influence the stereoselectivity of reactions subject to enantiomeric differentiation.[19]

Analysis of the Transition State for Jacobsen's Catalyst Catalysed Epoxidation of Styrene

The Gibb's Free Energy of each diastereomeric transition state was found from the .log file, with application of the thermal correction to free energy, giving result tabulated below.

R-styrene series ' ' S-styrene series ' '
Conformer DOI Gibbs Free Energy (Hartree/Particle) Gibbs Free Energy +Thermal Correction to Gibbs Free Energy(Hartree/Particle) Conformer DOI Gibbs Free Energy (Hartree/Particle) Gibbs Free Energy + Thermal Correction to Gibbs Free Energy (Hartree/Particle)
10.6084/m9.figshare.860441 -3343.9691970 -3343.3060230 10.6084/m9.figshare.860446 -3343.960889 -3343.297060
10.6084/m9.figshare.860445 -3343.9631910 -3343.2987320 10.6084/m9.figshare.860449 -3343.962162 -3343.297940

This analysis revealed that the lowest energy transition state occurred for the R-series, a conclusion supported by literature, [20] [21] [22] and so this transtition state, which corresponds to the cis-breakdown pathway of the radical intermediate shown below, was used in QTAIM analysis.

Jacobsen's catalyst epoxidation scheme
Jacobsen's catalyst radical intermediate breakdown pathway

S-styrene series QTAIM Lowest Energy Transition State

Jacobsen's catalyst catalysed epoxidation of styrene lowest energy transition state S-series

The S-series transition state indicated the significant formation of the C–O bond between the terminal carbon of the ethyl styrene substituent and the (salen)Mn-O oxygen. The C-O bond, at 1.779 Å is greater in length than the Mn-O bond which lies at 1.594 Å. The typical C-O bond length found in hydroxyl bonds lies at 1.43 Å , and 1.47 Å for epoxide C-O bonds. <refname="ja98253389760782">R. C Weast, Handbook of Chemistry and Physics, CRC Press, 65th Edition, 1984 Template:ISBN</ref> As the C-O interaction distance present in the transition is much greater than either of such bond lengths, it can be deduced that C-O bond formation in the transition state is not complete at C1. The β-carbon is oriented by the salen ligand oxygen, with an interaction distance of 3.054 Å. Additionally, the transition state is stabilised by π-overlap between the parallel displaced phenyl functionalities of styrene and the salen ligand, which are oriented with two strong interactions at distances of 3.290 and 3.388 Å.

R-styrene series QTAIM Lowest Energy Transition State

Jacobsen's catalyst catalysed epoxidation of styrene lowest energy transition state R-series

Styrene R-series QTAIM analysis
Styrene R-series QTAIM analysis
Styrene R-series QTAIM analysis
Styrene R-series QTAIM analysis

The lowest energy transition state of the Jacobsen’s catalyst catalysed epoxidation reaction of styrene, corresponded to the R-product pathway, and was analysed using QTAIM analysis. This agrees with literature reports carried out with [(R,R)-(–)-N,N’-bis(3,5-di-tert-butylsalicylidene)-1,2-cyclohexanediaminomanganese(III) chloride]((R,R)-Jacobsen's catalyst). <refname="ja7696574643">J. Cubillosa, W. Hölderich, Rev. Fac. Ing. Univ. Antioquia, 2007,41, 31-47Template:ISSN</ref> This indicated the significant formation of the C–O bond between the terminal carbon of the ethyl styrene substituent and the (salen)Mn-O oxygen. The C-O bond, at 1.803 Å is greater in length than the Mn-O bond which lies at 1.592 Å. The typical C-O bond length found in hydroxyl bonds lies at 1.43 Å, and 1.47 Å for epoxide C-O bonds. <refname="ja98253389760782">R. C Weast, Handbook of Chemistry and Physics, CRC Press, 65th Edition, 1984 Template:ISBN</ref> As the C-O interaction distance present in the transition is much greater than either of such bond lengths, it can be deduced that C-O bond formation in the transition state is not complete at C1. The β-carbon is oriented by the salen ligand oxygen, with an interaction distance of 3.054 Å. Additionally, the transition state is stabilised by π-overlap between the eclipsed phenyl functionalities of styrene and the salen ligand, which are oriented with two strong interactions at distances of 3.294 and 3.421 Å. The eclipsed benzyl functionalities have an increased spatial separation in the R-styrene oxide product pathway conformation, but have increased overlap of π–clouds as in the S-series transition states, the benzyl rings exhibit parallel displacement, and so reduced stabilisation. Thus the S-styrene oxide product transition states exist at higher energies than the transition states for the R-styrene oxide pathways, disfavouring the S-styrene oxide product at low temperatures and inducing the enantioselectivity for the R-Styrene oxide in the Jacobsen’s catalyst catalysed epoxidation. Additionally, the side-on approach of the Mn=O bond with the styrene oxide is as reported in previous studies, Heterogeneous asymmetric epoxidation of cis-ethyl cinnamte over <refname="ja053389760782">J. A. C. Lobo, Jacobsen’s catalyst immobilized in inorganic porous materials, El Bagre, Kolumbien, 2005</ref> and it can be seen that the non-planarity of the salen-Mn bonding plane reported previously is present in the lowest energy transition state.

Analysis of the Transition State for Shi Catalyst Catalysed Epoxidation of 1,2-Dihydronapthalene

The Gibb's Free Energy of each diastereomeric transition state was found from the .log file, with application of the thermal correction to free energy. This revealed that the lowest energy transition state occurred for the S,R-series, and this was used in QTAIM analysis.

R,S-series ' ' S,R-series ' '
Conformer DOI Gibbs Free Energy (Hartree/Particle) Gibbs Free Energy +Thermal Correction to Gibbs Free Energy(Hartree/Particle) Conformer DOI Gibbs Free Energy (Hartree/Particle) Gibbs Free Energy + Thermal Correction to Gibbs Free Energy (Hartree/Particle)
DOI:10.6084/m9.figshare.832492 -1381.120782 -1,380.697745 DOI:10.6084/m9.figshare.832538 -1381.131343 -1380.709885
DOI:10.6084/m9.figshare.832510 -1381.125886 -1,380.704512 DOI:10.6084/m9.figshare.832536 -1381.116109 -1380.696143
DOI:10.6084/m9.figshare.832511 -1381.134059 -1,380.711914 DOI:10.6084/m9.figshare.832545 -1381.126039 -1380.704465
DOI:10.6084/m9.figshare.832512 -1381.126722 -1,380.703543 DOI:10.6084/m9.figshare.832544 -1381.136239 -1380.714346

S,R-Dihydronapthalene oxide series QTAIM Highest Energy Transition State

QTAIM analysis of highest energy 1,2-dihydronapthalene oxide transition state DOI:10.6084/m9.figshare.832536


The transition state of the highest energy S,R-series Shi catalyst epoxidation of 1,2-dihydronapthalene corresponds to the S,R-product pathway. QTAIM analysis allows visualisation of the spiro transition state structure and significant C-O bond formation between the oxirane and C10 of the olefinic bond, which lies at a distance of 2.069 Å. The oxirane oxygen lies at an increased distance of 2.190 Å from C9, which lies closer to the benzyl functionality. QTAIM analysis further reveals the formation of strong interactions between ketal oxygen and the hexane-benzene bridging carbon, which lies at an interaction distance of 3.495 Å, and a much weaker interaction over a distance of 4.826 Å with the C3 ring bridging carbon and the ketal oxygen. Both of these interactions likely involve the overlap between the benzyl π-cloud and oxygen p-orbital lone pairs. A weaker interaction occur between the axial hydrogen of C1 and the pyranose ring oxygen at an interatomic distance of 2.088 Å and a further interaction exists between the C3 ring bridging carbon and the pyranose ring equatorial hydrogen separated by a spatial distance of 2.858 Å. For the transition states corresponding to the R,S-conformation, these interactions are reduced, as the available C1 hydrogen points equatorial, and is thus spatially further from the pyranose oxygen. Additionally, the benzyl group lies beneath the dioxacyclopentane ring at the 2-position on the pyranose. This ring projects down towards the benzyl group, but with the dioxacyclopentane carbon making the closest interaction with the benzyl group of 1,2-dihydronapthalene. This reduces the favourability of this orientation as the carbon lacks p-orbital lone pairs with which it can use to interact with the benzyl π-cloud.

S,R-Dihydronapthalene oxide series QTAIM Lowest Energy Transition State

QTAIM analysis of highest energy 1,2-dihydronapthalene oxide transition state DOI:10.6084/m9.figshare.832536

QTAIM analysis of lowest energy transition state
QTAIM analysis of lowest energy transition state
QTAIM analysis of lowest energy transition state

The transition state of lowest Gibb’s free energy of the Shi catalyst epoxidation of 1,2-dihydronapthalene corresponds to the S,R-product pathway, with an endo-structure between the phenyl functionality of the 1,2-dihydronapthalene reagent and the dioxyacyclopentane oxygen of the Shi catalayst. QTAIM analysis allows visualisation of the spiro transition state structure and significant C-O bond formation between the oxirane and C10 of the olefinic bond, which are spatially separated by 2.067 Å. The strong interaction here is due to the upward projection of the C1-C10 bond towards the dioxirane oxygen, increasing availability of the olefinic carbon for attack. The hydrogen of C1 is involved in a hydrogen-bonding like interaction with one oxygen of the dioxyacyclopentane at the C2-position on the pyranose ring, and a Van der Waals radii favourable overlap interaction with the dioxyacyclopentane methyl group. These interactions occur over distances of 2.445 and 2.233 Å respectively. The other dioxyacyclopentane oxygen is involved in an interaction with the other olefinic hydrogen at an interaction distance of 2.496 Å. The dioxirane oxygen lies at an increased distance of 2.189 Å from C9 c.f C10. QTAIM analysis further reveals the formation of strong interactions between ketal oxygen and the hexane-benzene bridging carbon, which lies at an interaction distance of 3.879 Å, and a much weaker interaction over a distance of 5.912 Å with the C3 ring bridging carbon and the other dioxyacyclopentane oxygen. Both of these interactions likely involve the overlap between the benzyl π-cloud and oxygen p-orbital lone pairs. A weaker interaction occur between the axial hydrogen of C2 and the dioxirane oxygen not involved in the epoxide formation, with an interatomic distance of 3.719 Å. For the transition state corresponding to the highest energy S,R-product, with the opposite direction of C10-C1 bond slope, the carbon involved in initial C-O bond formation with the dioxirane points downwards away from the dioxirane oxygen, reducing the interaction with the dioxirane oxygen due to an increased interaction distance of 2.069 Å. Additionally, the formation of the epoxide C-O bond occurs with the other dioxirane oxygen, which reduces steric clashing between the aromatic group of 1,2-dihydronapthalene oxide and the Shi catalyst dioxyacyclopentane rings, and allows for increase electrostatic interactions between the olefin and the dioxyacyclopentane oxygens.


NMR and Modelling of Alkenes and the Associated Epoxides

The two alkenes, and their corresponding epoxides chosen for molecular modelling were styrene (2-phenyloxirane) and 1,2-dihydronapthalene (1,2-dihydronapthalene oxide).


1,2-Dihydronapthalene Oxide

1,2-dihydronapthalene oxide has two chiral centres at each epoxy carbon, as shown in the following figure.

The two chiral centres and one conformational chiral plane in 1,2-dihydronapthalene oxide

The two chiral centres associated with the epoxide must exist as (R,S) or (S,R) due to the syn-substitution of the epoxy oxygen across the alkene double bond, and thus the two chiral centres can be treated as one chiral centre. One conformational chiral plane arises from the geometry of the C9-C10 bond in the hexane ring, which can either slope up or down with respect to the benzene ring plane. The epoxy chiral centres gives rise to 22 (four) enantiomers, two of which cannot exist favourably due to the syn-oxygen substitution. For each of the two allowed enantiomers, (R,S) and (S,R), two chiral conformations can exist. The chiral conformers can interconvert into each other through chair-antichair inversions of the cyclohexane ring. These are described in the figure below.

The two enantiomers of of 1,2-dihydronapthalene oxide and their conformational isomers

Each of the configurational and conformational isomers was modelled in Avogadro to the MMFF94s, conjugate gradient level of theory. Initial structures were minimised to obtain a minimum energy geometry for each isomer. These are as tabulated below.

Isomer Conformer 1 Conformer 2 Conformer 3
Structure Energy Contribution (kcal/mol) Structure Energy Contribution (kcal/mol) Structure Energy Contribution (kcal/mol)
(S,R)-1,2-dihydronapthalene oxide (C9-C10 slope up) 1,2-dihydronapthalene oxide 30.68348 kcal/mol Total Electronic Energy = 5.07558

Total Van der Waals = 19.97117
Total Out-of-Plane Bending Energy = 0.02988
Total Torsional Energy = 1.01016
Total Stretch Bending Energy = -0.80380
Total Angle Bending Energy = 2.61823
Total Bond Stetching Energy = 2.78225
Total Energy = 30.68348

1,2-dihydronapthalene oxide 30.68347 kcal/mol Total Electronic Energy = 5.07650

Total Van der Waals = 19.95799
Total Out-of-Plane Bending Energy = 0.03054
Total Torsional Energy = 1.01520
Total Stretch Bending Energy = -0.80232
Total Angle Bending Energy = 2.62588
Total Bond Stetching Energy = 2.77968
Total Energy = 30.68347

1,2-dihydronapthalene oxide 30.68364 kcal/mol Total Electronic Energy = 5.07403

Total Van der Waals = 19.97645
Total Out-of-Plane Bending Energy = 0.02967
Total Torsional Energy = 1.01202
Total Stretch Bending Energy = -0.80469
Total Angle Bending Energy = 2.61375
Total Bond Stetching Energy = 2.78242
Total Energy = 30.68364

(S,R)-1,2-dihdronapthalene oxide (C9-C10 slope down) 1,2-dihydronapthalene oxide 30.22445 kcal/mol Total Electronic Energy = 5.15408

Total Van der Waals = 19.38146
Total Out-of-Plane Bending Energy = 0.00922
Total Torsional Energy = 1.67327
Total Stretch Bending Energy = -0.86999
Total Angle Bending Energy = 2.17888
Total Bond Stretching Energy = 2.69754
Total Energy = 30.22445

1,2-dihydronapthalene oxide 30.22440 kcal/mol Total Electronic Energy = 5.15300

Total Van der Waals = 19.37922
Total Out-of-Plane Bending Energy = 0.00939
Total Torsional Energy = 1.67403
Total Stretch Bending Energy = -0.86951
Total Angle Bending Energy = 2.18029
Total Bond Stretching Energy = 2.69798
Total Energy = 30.22440

(R,S)-1,2-dihydronapthalene oxide (C9-C10 slope down) 1,2-dihydronapthalene oxide 30.68345 kcal/mol Total Electronic Energy = 5.07563

Total Van der Waals = 19.96768 =
Total Out-of-Plane Bending Energy = 0.03003
Total Torsional Energy = 1.01200
Total Stretch Bending Energy = -0.80345
Total Angle Bending Energy = 2.61984
Total Bond Stretching Energy = 2.78171
Total Energy = 30.68345

1,2-dihydronapthalene oxide 30.68344 kcal/mol Total Electronic Energy = 5.07608

Total Van der Waals = 19.96371
Total Out-of-Plane Bending Energy = 0.03025
Total Torsional Energy = 1.01298
Total Stretch Bending Energy = -0.80303
Total Angle Bending Energy = 2.62238
Total Bond Stretching Energy = 2.78106
Total Energy = 30.68344

(S,R)-1,2-dihydronapthalene oxide (C9-C10 slope up) 1,2-dihydronapthalene oxide 30.22442 kcal/mol Total Electronic Energy = 5.15329

Total Van der Waals = 19.37677
Total Out-of-Plane Bending Energy = 0.00950
Total Torsional Energy = 1.67492
Total Stretch Bending Energy = -0.86888
Total Angle Bending Energy = 2.18145
Total Bond Stretching Energy = 2.69737
Total Energy = 30.22442

1,2-dihydronapthalene oxide 30.22441 kcal/mol Total Electronic Energy = 5.15317

Total Van der Waals = 19.37800
Total Out-of-Plane Bending Energy = 0.00944
Total Torsional Energy = 1.67446
Total Stretch Bending Energy = -0.86923
Total Angle Bending Energy = 2.18084
Total Bond Stretching Energy = 2.69775
Total Energy = 30.22441

As expected, the conformational isomers, (S,R)-1,2-dihydronapthalene oxide (C9-C10 slope up) and (S,R)-1,2-dihydronapthalene oxide (C9-C10 slope down), and (R,S)-1,2-dihydronapthalene oxide (C9-C10 slope up) and (R,S)-1,2-dihydronapthalene oxide (C9-C10 slope down),give the same energy to within 0.00001 kcal/mol, whilst of the two (S,R) and (R,S)-1,2-dihydronapthalene oxide enantiomers,the (S,R) enantiomer gives a lower energy, and thus is predicted to be more stable than the (R,S) configuration.

NMR

Calculation of Oxide Correction Constants

To allow for the correction factors of the epoxy carbons, ethane and ethylene oxide were modelled using the MMFF94s force field with conjugate gradients in Avogadro. These input structures were optimised to DFT B3LYP level of theory using the 6-31G(d,p) basis set, and NMR calculations carried out on the optimised structures.

  • Ethane

The MMFF94s optimised structure used as input to the DFT B3LYP level of theory optimisation is shown below.

Ethane geometry optimised using MMFF94s force field

Ethane


Energy parameters associated with this structure are as tabulated.

Energy Contribution Energy (kcal/mol)
Total Electrostatic Energy 0.00000
Total Van der Waals Energy 0.11175
Total Out-of-Plane Bending Energy 0.00000
Total Stretch Bending Energy -0.00153
Total Torsional Energy -4.95900
Total Angle Bending Energy 0.01595
Total Bond Stretching Energy 0.00742
Total Energy -4.82541
  • Ethylene oxide

Ethylene oxide was modelled using the MMFF94s force field in Avogadro, optimised to the DFT B3LYP 6-31G(d,p) level of theory with an NMR simulation carried out. The structure used as DFT B3LYP optimisation input is demonstrated below.

Ethylene oxide geometry optimised using MMFF94s force field

Ethylene oxide

Energy parameters associated with this structure are tabulated below.

Energy Contribution Energy (kcal/mol)
Total Electrostatic Energy 3.49639
Total Van der Waals Energy 0.01963
Total Out-of-Plane Bending Energy 0.00000
Total Stretch Bending Energy -0.00982
Total Torsional Energy 2.35781
Total Angle Bending Energy 0.18271
Total Bond Stretching Energy 0.00462
Total Energy 6.05133

Correction Factors

The ethane and ethylene structures gave 13C NMR peaks corrected with reference: TMS B3LYP/6-31G(d,p) Chloroform of shielding 192.17 ppm with NMR Degeneracy Tolerance of 0.05. Numbering used when referring to ethane atom shifts are made with refernce to the figure below.

Atomic numbering used in NMR spectral analysis
Atomic numbering used in NMR spectral analysis


The calculated C-O oxide δcorr (ppm) was given by subtracting the ethylene shift from the ethane shift for the equivalent atoms and are as documented in the table below.

Ethane 6-31G(d,p) Ethylene 6-31G(d,p)
Carbon Number δcalc (ppm) Average δcalc (ppm) Carbon Number δcalc (ppm) Average δcalc (ppm) δcorr (ppm)
C1 9.4675 9.4675 C1 41.194 41.19395 31.72645
C2 9.4675 C2 41.1939

The equivalent analysis for 1H NMR spectra corrected with reference: TMS B3LYP/6-31G(d,p) Chloroform of shielding reference 31.7462 ppm and NMR Degeneracy Tolerance of 0.05

Ethane 6-31G(d,p) Ethylene 6-31G(d,p)
Hydrogen Number δcalc (ppm) Average δcalc (ppm) Hydrogen Number δcalc (ppm) Average δcalc (ppm) δcorr (ppm)
H4 0.9506 0.949933333 H1 2.6096 2.6096 1.659666667
H7 0.9504 H2 2.6096
H5 0.9504 H3 2.6096
H3 0.9488 H4 2.6096
H6 0.9488
H8 0.9504

It is well reported that the 1H NMR spectrum of ethane has one signal which is an average of all hydrogen signals in the rapidly interchanging conformations adopted by ethane in solution occurring by rotation around the central C-C bond. [23] However the presently calculated spectrum exhibits three different signals, one for each equivalent hydrogen pair in ethane. This reinforces the static nature of the simulation, which takes into account only the one conformation. In this case, this conformer is the lowest energy staggered conformation, which provides the most accurate representation of the molecule as this conformer, residing at an energy minimum, is the dominant conformation found for a selection of ethane molecules at any given time. The six signals were averaged to give one signal. This provides a rather poor representation of the conformations in solution however, as the experimental final shift is an average of those observed for each conformation. The δcorr (ppm) was given by subtracting the ethylene shift from the ethane shift.

1,2-dihydronapthalene oxide NMR

The NMR simulation was carried out on the lowest energy conformer as this structure represents the dominant structure at any one time in solution. Referring to the table above, this gives the 30.22440 kcal/mol isomer.


Higher Level of Theory

Numbering used in NMR analysis
Numbering used in NMR analysis

The NMR was also carried out using the DFT B3LYP cc-pVTZ level of theory.This gave a structure with thermal free energies of:

Sum of electronic and thermal Energies= -462.280088
Sum of electronic and thermal Enthalpies= -462.279
Sum of electronic and thermal Free Energies= -462.320756


The 13C NMR and 1H NMR spectra were corrected for TMS B3LYP/cc-pVTZ GIAO are as shown below, respectively.

13C NMR of 30.22441 kcal/mol isomer
13C NMR of 30.22441 kcal/mol isomer
1H NMR of 30.22441 kcal/mol isomer
1H NMR of 30.22441 kcal/mol isomer

These results were compared with literature for the (+)-enantiomer, [24] and disparity with experimental shifts calculated and tabulated below.


Carbon Number δcalc (ppm) δexp (ppm) δcorr (ppm) ∆δ (ppm) Average ∆δ (ppm) Mode ∆δ (ppm) ∆δcorr (ppm) Average ∆δ (ppm) Mode ∆δ (ppm)
C1 26.3666 22.2 4.1666 5.14473 8.1021 4.1666 -0.30897 -23.77355
C2 60.3019 55.5 31.72645 4.8019 -23.77355
C3 57.688 55.2 31.72645 2.488 -23.47355
C4 141.0021 132.9 8.1021 8.1021
C5 145.0906 137.1 7.9906 7.9906
C6 29.297 24.8 4.497 4.497
C7 133.4602 128.8 4.6602 4.6602
C8 133.5733 128.8 4.7733 4.7733
C9 131.242 126.5 4.742 4.742
C10 135.1256 129.9 5.2256 5.2256
13C NMR difference graph of 30.22441 kcal/mol isomer with literature(+)-enantiomer
13C NMR difference graph of 30.22441 kcal/mol isomer with literature(+)-enantiomer


The carbon NMR spectrum was compared with literature for the (-)-enantiomer, [25] giving results tabulated below.

Carbon Number δcalc (ppm) δexp (ppm) δcorr (ppm) ∆δ (ppm) Average ∆δ (ppm) Mode ∆δ (ppm) ∆δcorr (ppm) Average ∆δ (ppm) Mode ∆δ (ppm)
C1 26.3666 20.8 5.5666 6.77473 9.5021 5.5666 1.32103 -22.37355
C2 60.3019 54.1 31.72645 6.2019 -22.37355
C3 57.688 51.7 31.72645 5.988 -19.97355
C4 141.0021 131.5 9.5021 9.5021
C5 145.0906 135.6 9.4906 9.4906
C6 29.297 23.4 5.897 5.897
C7 133.4602 127.3 6.1602 6.1602
C8 133.5733 127.4 6.1733 6.1733
C9 131.242 125.1 6.142 6.142
C10 135.1256 128.5 6.6256 6.6256
13C NMR difference graph of 30.22441 kcal/mol isomer with literature(-)-enantiomer
13C NMR difference graph of 30.22441 kcal/mol isomer with literature(-)-enantiomer

The calculated carbon shifts have a 1.63 ppm larger disparity with the (-)-enantiomer [25] than with the (+)-enantiomer [24] literature carbon shifts. The disparity between literature shifts and the calculated shifts is due to the static nature of the molecule being run in the calculation, which is of one configuration and conformation. In reality the molecule is dynamic, and making transitions between different energetically available conformers on the submillisecond timescale, and so the different proton shifts that occur in each conformer are completely averaged out on the NMR timescale, which is ~1 sec. Thus a spectrum that is an average of the exchanging states is yielded. The interconversion between conformers is fast on the NMR time scale as it occurs at a much greater rate than the frequency difference between the two conformers. The difference in frequency depends on the Larmor frequency of the observed nucleus and the magnetic field strength, thus changing the magnetic field strength and cooling the solution undergoing the NMR experimental analysis would allow recreation of the static calculated spectrum of one conformer which lies at the energy minimum.
Additionally, intermolecular coupling between molecules is not taken account of in the calculated spectrum, which considers only one molecule. Thus coupling constants and peak multiplicity in the observed spectra would be expected to differ with those present in the calculated spectra.

The equivalent analysis was undertaken for the 1H NMR spectrum. Below is documented the calculated 1H NMR shifts calcualted at the DFT B3LYP cc-pVTZ level of theory and their disparity with literature shifts for the (+)-enantiomer

Hydrogen Number δcalc (ppm) δexp (ppm) δcorr (ppm) ∆δ (ppm) Average ∆δ (ppm) Mode ∆δ (ppm) ∆δcorr (ppm) Average ∆δ (ppm) Mode ∆δ (ppm)
H11 7.4035 7.018 0.3855 0.281076667 0.8442 0.3855 -0.13544 -2.122333333
H12 7.5534 7.1754 0.378 0.378
H13 7.5182 7.1356 0.3826 0.3826
H14 7.7166 7.322 0.3946 0.3946
H15 2.5507 2.2555 0.2952 0.2952
H16 2.8556 2.715666667 0.139933333 0.139933333
H17 2.3642 1.52 0.8442 0.8442
H18 1.6547 1.6988 -0.0441 -0.0441
H20 3.6844 3.667666667 1.659666667 0.016733333 -2.008
H21 3.8001 3.782 1.659666667 0.0181 -2.122333333

Th 1H NMR shift were compared with literature for the (-)-enantiomer.[25] Results are as tabulated below.

Hydrogen Number δcalc (ppm) Average δcalc (ppm) δexp (ppm) δcorr (ppm) ∆δ (ppm) Average ∆δ (ppm) Mode ∆δ (ppm) ∆δcorr (ppm) Average ∆δ (ppm) Mode ∆δ (ppm)
H11 7.4035 7.01 0.3935 0.19252 0.0142 0.3935 -0.223996667 -2.120333333
H12 7.5534 7.5358 7.1356 0.4178 0.4178
H13 7.5182 7.5358 7.1356 0.3826 0.3826
H14 7.7166 7.34 0.3766 0.3766
H15 2.5507 2.45 0.1007 0.1007
H16 2.8556 2.675 0.1806 0.1806
H17 2.3642 2.35 0.0142 0.0142
H18 1.6547 1.65 0.0047 0.0047
H20 3.6844 3.65 1.659666667 0.0344 -1.990333333
H21 3.8001 3.78 1.659666667 0.0201 -2.120333333


1H NMR difference graph between calculated hydrogen shifts and literature shifts for the (-)-enantiomer
1H NMR difference graph between calculated hydrogen shifts and literature shifts for the (-)-enantiomer

A Higher Energy Enantiomer, 30.22441 kcal/mol input energy

To allow for a comparison of the lowest energy isomer NMR, an additional NMR of a higher entropy isomer was carried out on (S,R)-1,2-dihydronapthalene oxide, of input energy 30.22441 kcal/mol, modelled at the DFT B3LYP 6-31G(d,p) level of theory in chloroform solvent, giving the structure shown below.


Numbering used for the 1H NMR and 13C NMR spectra of 1,2-dihydronapthalene oxide 30.22441 kcal/mol at DFT B3LYP 6-31G(d,p)level of theory
Electron density surface of 1,2-dihydronapthalene oxide 30.22441 kcal/mol at DFT B3LYP 6-31G(d,p)level of theory, coloured by electrostatic potential
HOMO of 1,2-dihydronapthalene oxide 30.22441 kcal/mol at DFT B3LYP 6-31G(d,p)level of theory, coloured by electrostatic potential

Following TMS/DFT B3LYP 6-31G(d,p) (chloroform) correction application, this gave 1H and 13C NMR spectra as documented below.

1H NMR data

1H NMR spectrum of 1,2-dihydronapthalene oxide 30.22441 kcal/mol at DFT B3LYP 6-31G(d,p)level of theory

Comparison of B3LYP 6-31G(d,p) optimised structure with MMFF94s force field, conjugate gradient optimisation algorithm

The 30.2441 kcal/mol (S,R)-1,2-dihydronapthalene oxide conformer was optimised through the HPC using Gaussian 09 to the DFT B3LYP 6-31G (d,p) level of theory, and NMR, IR and VCD simulations run on this optimised configuration. Energetics of the optimised molecule are reported in the table below. The lower total energy of the molecule when run to the DFT B3LYP level of theory is lower than at the MMFF94s, conjugate gradient level of theory.

Energy Contribution Energy (kcal/mol)
Total Electrostatic Energy -4.01005
Total Van der Waals Energy 18.10016
Total Out-of-Plane Bending Energy 0.01243
Total Stretch Bending Energy -0.46660
Total Torsional Energy 1.92867
Total Angle Bending Energy 3.35949
Total Bond Stretching Energy 3.11745
Total Energy 22.04156


Comparison of 1H NMR data with (+)-1,2-Dihydronapthalene Oxide Literature

Comparing the calculated NMR shifts with D. Xiong et al. (USE REFERENCE IN FIGURE), the disparity between calculated and experimental shifts could be found. This is reported in the following table and figure.

Hydrogen Number δcalc (ppm) δexp (ppm) ∆δ (ppm)
H11 7.251492731 7.01 0.241492731
H12 7.391151507 7.18 0.211151507
H13 7.387571667 7.185 0.202571667
H14 7.615118252 7.34 0.275118252
H15 2.267285883 2.45 -0.182714117
H16 2.946631263 2.675 0.271631263
H17 2.209025512 2.35 -0.140974488
H18 1.873442614 1.65 0.223442614
H20 3.483067617 3.65 -0.166932383
H21 3.559560436 3.78 -0.220439564
1</>H NMR ∆δ (ppm) of (S,R)-1,2-Dihydronapthalene oxide, 30.22441 kcal/mol conformer with literature D. Xiong, X. Hu, S. Wang, C.-X. Miao, C. Xia, W. Sun,E. J. Org. Chem., 2011, 2011, 23, 4289–4292
1</>H NMR ∆δ (ppm) of (-)-1,2-Dihydronapthalene oxide 30.22441 kcal/mol

Carbon NMR Comparison with (-)-Rotamer

Carbon Number δcalc (ppm) δexp (ppm) δcorr (ppm) ∆δ (ppm) Average ∆δ (ppm) Mode ∆δ (ppm) ∆δcorr (ppm) Average ∆δ (ppm) Mode ∆δ (ppm)
C1 29.06348904 20.8 8.263489036 0.03509053 8.263489036 8.263489036 -4.120980016 -22.37355
C2 52.82116189 54.1 31.72645 -1.278838112 -22.37355
C3 52.19244356 51.7 31.72645 0.492443564 -19.97355
C4 130.3705995 131.5 -1.129400532 -1.129400532
C5 135.38776 135.6 -0.212240025 -0.212240025
C6 30.18028122 23.4 6.780281225 6.780281225
C7 123.533413 127.3 -3.76658704 -3.76658704
C8 123.7910881 127.4 -3.60891185 -3.60891185
C9 121.7441924 125.1 -3.355807577 -3.355807577
C10 126.6664766 128.5 -1.833523393 -1.833523393
1</>H NMR ∆δ (ppm) of (+)-1,2-Dihydronapthalene oxide 30.22441 kcal/mol

(-)-1,2-Dihydronapthalene oxide Literature Comparison

Further literature comparison with the (-)-rotamer was undertaken as a comparative study with the (+)-rotamer.

13C NMR data=

At the B3LYP 6-31G level of theory, 13C NMR spectrum is as below.


The Carbon shifts and disparity with literature[26] is as noted below.

Carbon Number δcalc (ppm) δexp (ppm) δcorr (ppm) ∆δ (ppm) Average ∆δ (ppm) Mode ∆δ (ppm) ∆δcorr (ppm) Average ∆δ (ppm) Mode ∆δ (ppm)
C1 29.06348904 20.8 8.263489036 0.03509053 8.263489036 8.263489036 -4.120980016 -22.37355
C2 52.82116189 54.1 31.72645 -1.278838112 -22.37355
C3 52.19244356 51.7 31.72645 0.492443564 -19.97355
C4 130.3705995 131.5 -1.129400532 -1.129400532
C5 135.38776 135.6 -0.212240025 -0.212240025
C6 30.18028122 23.4 6.780281225 6.780281225
C7 123.533413 127.3 -3.76658704 -3.76658704
C8 123.7910881 127.4 -3.60891185 -3.60891185
C9 121.7441924 125.1 -3.355807577 -3.355807577
C10 126.6664766 128.5 -1.833523393 -1.833523393

The literature disparity is summarised in the following table.

Disparity of 13C NMR shifts with literature
Disparity of 13C NMR shifts with literature

Comparison of 13C NMR data with Literature

Repeat of NMR Analysis Using a Different Conformer and Higher Level of Theory

The MMFF94s force field, conjugate gradient algorithm optimised structures of the 30.2244 kcal/mol configurational isomer and the 30.6834 kcal/mol isomer were optimised further using Gaussian09 at the DFT B3LYP 6-31G(d,p) level of theory. These structures were input into optimisation at the higher DFT B3LYP cc-pVTZ level of theory, and NMR simulations carried out on both optimised structures. Results are as documented below.

30.68344 kcal/mol isomer at the 6-31G (d,p) level of theory

The 13C NMR spectrum of this isomer was corrected with reference TMS B3LYP/6-31G(d,p) Chloroform and reference shielding 192.17 ppm and NMR Degeneracy Tolerance of 0.05.

Atomic numbering used for the 30.68344 kcal/mol isomer is shown below.

Atomic numbering for the 30.68344 kcal/mol isomer used in NMR analysis
Atomic numbering for the 30.68344 kcal/mol isomer used in NMR analysis

The 13C NMR spectrum is shown below, with corrected shifts and comparison with literature,[26] tabulated.

13C NMR spectrum of the 30.68344 kcal/mol isomer at the B3LYP 6-31G(d,p) level of theory
13C NMR spectrum of the 30.68344 kcal/mol isomer at the B3LYP 6-31G(d,p) level of theory
Carbon Number δcalc (ppm) δexp (ppm) δcorr (ppm) ∆δ (ppm) Average ∆δ (ppm) Mode ∆δ (ppm) ∆δcorr (ppm) Average ∆δ (ppm) Mode ∆δ (ppm)
C1 24.94531918 20.8 4.145319183 -0.267577089 -4.145319183 4.145319183 6.077712911 34.85481801
C2 56.4348396 54.1 88.1612896 2.334839596 34.0612896
C3 54.82836801 51.7 86.55481801 3.128368014 34.85481801
C4 130.383024 131.5 -1.116975979 -1.116975979
C5 133.8418328 135.6 -1.758167178 -1.758167178
C6 28.03616498 23.4 4.63616498 4.63616498
C7 123.8353838 127.3 -3.464616185 -3.464616185
C8 123.425036 127.4 -3.974963965 -3.974963965
C9 121.4413171 125.1 -3.658682863 -3.658682863
C10 125.5529435 128.5 -2.947056492 -2.947056492
13C NMR difference spectrum of the 30.68344 kcal/mol isomer at the B3LYP 6-31G(d,p) level of theory with literature
13C NMR difference spectrum of the 30.68344 kcal/mol isomer at the B3LYP 6-31G(d,p) level of theory with literature

Applying the oxide C-O correction in fact increases the disparity with literature,[26] as can be seen in the following figure.

13C NMR difference spectrum of the 30.68344 kcal/mol isomer at the B3LYP 6-31G(d,p) level of theory with literature
13C NMR difference spectrum of the 30.68344 kcal/mol isomer at the B3LYP 6-31G(d,p) level of theory with literature

This may be due to the poor modelling of the simple ethylene oxide environment used for carbon correction factors when compared to the increased complexity of benzylic environment present for 1,2-dihydronapthalene. TALK ABOUT NAPTHALENE HERE!!!!

The equivalent analysis was carried out for the 1H NMR spectrum, and is documented below.

1H NMR spectrum of the 30.68344 kcal/mol isomer at the B3LYP 6-31G(d,p) level of theory
1H NMR spectrum of the 30.68344 kcal/mol isomer at the B3LYP 6-31G(d,p) level of theory

X-Axis: Shift (ppm)

  1. Y-Axis: Degeneracy
  2. Reference: TMS B3LYP/6-31G(d,p) Chloroform
  3. Reference shielding: 31.7462 ppm
  4. NMR Degeneracy Tolerance: 0.05
  5. X Y Degeneracy

14-H 7.4903000000 1.0000000000 1.0000000000 12-H 7.4035000000 1.0000000000 2.0000000000 13-H 7.3689000000 2.0000000000 2.0000000000 11-H 7.1952000000 1.0000000000 1.0000000000 21-H 3.6098000000 1.0000000000 2.0000000000 20-H 3.5657000000 2.0000000000 2.0000000000 16-H 2.8801000000 1.0000000000 1.0000000000 15-H 2.2677000000 1.0000000000 1.0000000000 18-H 2.2162000000 1.0000000000 1.0000000000 17-H 1.5628000000 1.0000000000 1.0000000000

To allow for comparison with literature, [27] the original spectra were used to refer to peaks notated as a "multiplet" in literature, and the peaks corresponding to one hydrogen environment were identified. For those hydrogens which observed coupling, each peak present in the multiplet was identified and these peaks averaged to find the centre of the multiplet, which corresponds to the chemical shift of the hydrogen.

Hydrogen Number δcalc (ppm) δexp (ppm) δcorr (ppm) ∆δ (ppm) Average ∆δ (ppm) Mode ∆δ (ppm) ∆δcorr (ppm) Average ∆δ (ppm) Mode ∆δ (ppm)
H11 7.1952 7.018 0.1772 0.126956667 6.8408 0.1772 0.45889 -0.050243333
H12 7.4035 7.1754 0.2281 0.2281
H13 7.3689 7.1356 0.2333 0.2333
H14 7.4903 7.322 0.1683 0.1683
H15 2.2677 2.2555 0.0122 0.0122
H16 2.8801 2.715666667 0.164433333 0.164433333
H17 1.5628 1.52 0.0428 0.0428
H18 2.2162 1.6988 0.5174 0.5174
H20 3.5657 3.667666667 5.225366667 -0.101966667 1.5577
H21 3.6098 3.782 5.269466667 -0.1722 1.487466667
1H NMR spectrum of the 30.68344 kcal/mol isomer at the B3LYP 6-31G(d,p) level of theory
1H NMR spectrum of the 30.68344 kcal/mol isomer at the B3LYP 6-31G(d,p) level of theory

As for the carbon spectrum, applying the oxide hydrogen correction in fact increases the disparity between calculated results and literature, albeit the increase between these results is much lower, ~1.6 ppm for the hydrogen c.f ~32 ppm for the carbon NMR spectrum. This is most probably due to the increase distance between the hydrogens and epoxide oxygen, which gives the smaller correction factor calculated in the ethane model. Nevertheless, the increase in disparity reinforces the conclusion that the ethane environment is too oversimplified to allow for accurate correction factors. In general, the difference between the calculated and literature 1H NMR peaks[27] is lower than that observed for the 13C NMR spectrum.

30.68344 kcal/mol isomerAt the cc-pVTZ level of theory

H nmr spec napth oxide cc-pvtz 30.68344.PNG

Reference: TMS B3LYP/6-31G(d,p) Chloroform Reference shielding: 31.7462 ppm NMR Degeneracy Tolerance: 0.05


14-H 7.7291000000 1.0000000000 1.0000000000 12-H 7.5659000000 1.0000000000 2.0000000000 13-H 7.5307000000 2.0000000000 2.0000000000 11-H 7.4160000000 1.0000000000 1.0000000000 21-H 3.8126000000 1.0000000000 1.0000000000 20-H 3.6970000000 1.0000000000 1.0000000000 16-H 2.8681000000 1.0000000000 1.0000000000 15-H 2.5632000000 1.0000000000 1.0000000000 18-H 2.3767000000 1.0000000000 1.0000000000 17-H 1.6672000000 1.0000000000 1.0000000000


C nmr spec napth oxide cc-pvtz 30.68344.PNG X-Axis: Shift (ppm)

  1. Y-Axis: Degeneracy
  2. Reference: TMS B3LYP/6-31G(d,p) Chloroform
  3. Reference shielding: 192.17 ppm
  4. NMR Degeneracy Tolerance: 0.05
  5. X Y Degeneracy
5-C      152.8919000000        1.0000000000        1.0000000000
4-C      148.8033000000        1.0000000000        1.0000000000

10-C 142.9270000000 1.0000000000 1.0000000000

8-C      141.3746000000        1.0000000000        1.0000000000
7-C      141.2615000000        1.0000000000        1.0000000000
9-C      139.0433000000        1.0000000000        1.0000000000
2-C       68.1034000000        1.0000000000        1.0000000000
3-C       65.4894000000        1.0000000000        1.0000000000
6-C       37.0983000000        1.0000000000        1.0000000000
1-C       34.1679000000        1.0000000000        1.0000000000

30.22440 kcal/mol isomer

C nmr spec napth oxide cc-pvtz 30.2244.PNG

X-Axis:  Shift (ppm)
  1. Y-Axis: Degeneracy
  2. Reference: TMS B3LYP/6-31G(d,p) Chloroform
  3. Reference shielding: 192.17 ppm
  4. NMR Degeneracy Tolerance: 0.05
  5. X Y Degeneracy
5-C      152.8919000000        1.0000000000        1.0000000000
4-C      148.8034000000        1.0000000000        1.0000000000

10-C 142.9269000000 1.0000000000 1.0000000000

8-C      141.3746000000        1.0000000000        1.0000000000
7-C      141.2615000000        1.0000000000        1.0000000000
9-C      139.0433000000        1.0000000000        1.0000000000
2-C       68.1032000000        1.0000000000        1.0000000000
3-C       65.4893000000        1.0000000000        1.0000000000
6-C       37.0983000000        1.0000000000        1.0000000000
1-C       34.1679000000        1.0000000000        1.0000000000


H nmr spec napth oxide cc-pvtz 30.2244.PNG

X-Axis: Shift (ppm)

  1. Y-Axis: Degeneracy
  2. Reference: TMS B3LYP/6-31G(d,p) Chloroform
  3. Reference shielding: 31.7462 ppm
  4. NMR Degeneracy Tolerance: 0.05
  5. X Y Degeneracy

14-H 7.7291000000 1.0000000000 1.0000000000 12-H 7.5659000000 1.0000000000 2.0000000000 13-H 7.5307000000 2.0000000000 2.0000000000 11-H 7.4160000000 1.0000000000 1.0000000000 21-H 3.8126000000 1.0000000000 1.0000000000 20-H 3.6969000000 1.0000000000 1.0000000000 16-H 2.8681000000 1.0000000000 1.0000000000 15-H 2.5632000000 1.0000000000 1.0000000000 17-H 2.3767000000 1.0000000000 1.0000000000 18-H 1.6672000000 1.0000000000 1.0000000000

IR and VCD analysis

IR and VCD analysis was carried out for the lowest energy isomer at input 30.2244 kcal/mol, optimised to the DFT B3LYP cc-pVTZ level of theory, with results as below.

Napth ox 30.2244 IR cc-pVTZ im.PNG

IR for the lowest energy isomer of 1,2-dihydronapthalene oxide otpimised to DFT B3LYP cc-pVTZ levle of theory
IR for the lowest energy isomer of 1,2-dihydronapthalene oxide otpimised to DFT B3LYP cc-pVTZ levle of theory
VCD for the lowest energy isomer of 1,2-dihydronapthalene oxide otpimised to DFT B3LYP cc-pVTZ levle of theory
VCD for the lowest energy isomer of 1,2-dihydronapthalene oxide otpimised to DFT B3LYP cc-pVTZ levle of theory


Infrared spectrum of 1,2-dihydronapthalene oxide 30.22441 kcal/mol at DFT B3LYP 6-31G(d,p)level of theory


VCD can be used to elucidate the enantioeric excess (ee) of a sample consisting of a mixture of enantiomers of a chiral molecule. The sample % ee is defined as:

(NA - NB) x 100%/ (NA + NB)

where NA and NA is a measure of moles of enantiomer A and enantiomer B respectively. [28]This expression is equivalent to the excess amount of one enantiomer over the other relative to the total amount of both enantiomers. Optically pure samples of A have 100% ee, racemic mixtures 0% and optically pure B sample have -100% purity. VCD can be used for ee determination from the multiplicity of spectral bands that can be measure simultaneously in one spectrum. The intensity of the VCD bands gives an indication of the ee, as each VCD spectrum for each enantiomer is a mirror image of the other. [29]

VCD spectrum of 1,2-dihydronapthalene oxide 30.22441 kcal/mol at DFT B3LYP 6-31G(d,p)level of theory

Styrene oxide Configurations

Styrene oxide has a chiral center at the benzylic carbon atom, giving (R)-styrene oxide and (S)-styrene oxide enantiomers. The energies of these two enantiomers were minimised to the MMFF94s, conjugate gradient level of theory.

Enantiomer Conformer 1 Conformer 2 Conformer 3
Structure Energy Contribution (kcal/mol) Structure Energy Contribution (kcal/mol) Structure Energy Contribution (kcal/mol)
R-styrene oxide R-styrene oxide 21.61506 kcal/mol Total Electronic Energy = 3.07365

Total Van der Waals = 13.67376
Total Out-of-Plane Bending Energy = 0.00164
Total Torsional Energy = 2.34616
Total Stretch Bending Energy = -0.74666
Total Angle Bending Energy = 1.42461
Total Bond Stetching Energy = 1.84188
Total Energy = 21.61506

R-styrene oxide 21.61547 kcal/mol Total Electronic Energy = 3.08594

Total Van der Waals = 21.61547
Total Out-of-Plane Bending Energy = 0.00169
Total Torsional Energy = 2.34606
Total Stretch Bending Energy = -0.74594
Total Angle Bending Energy = 1.42450
Total Bond Stetching Energy = 1.84036
Total Energy = 30.68347

S-styrene oxide S-styrene oxide 21.61503 kcal/mol Total Electronic Energy = 3.07620

Total Van der Waals = 13.67135
Total Out-of-Plane Bending Energy = 0.00166
Total Torsional Energy = 2.34617
Total Stretch Bending Energy = -0.74656
Total Angle Bending Energy = 1.42457
Total Bond Stretching Energy = 1.84164
Total Energy = 21.61503

S-styrene oxide 21.61507 kcal/mol Total Electronic Energy = 3.07301

Total Van der Waals = 13.67434
Total Out-of-Plane Bending Energy = 0.00164
Total Torsional Energy = 2.34620
Total Stretch Bending Energy = -0.74667
Total Angle Bending Energy = 1.42456
Total Bond Stretching Energy = 1.84199
Total Energy = 21.61507

S-styrene oxide 21.61508 kcal/mol Total Electronic Energy = 3.07293

Total Van der Waals = 13.67443
Total Out-of-Plane Bending Energy = 0.00164
Total Torsional Energy = 2.34616
Total Stretch Bending Energy = -0.74669
Total Angle Bending Energy = 1.42463
Total Bond Stretching Energy = 1.84199
Total Energy = 21.61508

NMR analysis of S-styrene oxide

The NMR of (S)-styrene oxide was modelled at the DFT B3LYP 6-31G(d,p) level of theory in chloroform solvent, giving the structure as shown below.

(S)-styrene oxide at DFT B3LYP 6-31G(d,p)level of theory

Numbering used for the 1H NMR and 13C NMR spectra of S-styrene oxide at DFT B3LYP 6-31G(d,p)level of theory

Following TMS/DFT B3LYP 6-31G(d,p) (chloroform) correction application, this gave 1H and 13C NMR spectra as documented in the figures below.

13C NMR spectra of S-styrene oxide at DFT B3LYP 6-31G(d,p)level of theory
1H NMR spectra of S-styrene oxide at DFT B3LYP 6-31G(d,p)level of theory

Numbering of atoms used in the following NMR analysis is demonstrated below.

Atomic numbering of S-styrene lowest energy conformation use in NMR analysis
Atomic numbering of S-styrene lowest energy conformation use in NMR analysis

Comparison of S-styrene Oxide with Literature

The carbon shifts were compared with both S-styrene oxide NMR and R-styrene oxide literature NMR shifts.

Comparison with R-styrene oxide literature shifts (K. Sarma, N. Bhati et al.) is as documented below.

Carbon Number δcalc (ppm) δexp (ppm) δcorr (ppm) ∆δ (ppm) Mode ∆δ (ppm) ∆δcorr (ppm) Average ∆δ (ppm) Mode ∆δ (ppm)
C2 122.9659755 128.1 -5.134024522 -7.231939222 5.134024522 11.7736977 33.982242
C3 123.4152252 128.5 -5.0847748 5.0847748
C5 122.9534464 125.5 -2.546553554 2.546553554
C7 124.1326453 128.5 -4.367354719 4.367354719
C9 118.2680608 125.5 -7.231939222 7.231939222
C11 135.1343367 137.6 -2.465663325 2.465663325
C12 54.05057945 52.4 85.77702945 1.650579448 33.37702945
C13 53.455792 51.2 85.182242 2.255791999 33.982242
S-Styrene oxide carbon NMR shift difference graph with R-Styrene oxide literature (K. Sarma, N. Bahati et al.)
S-Styrene oxide carbon NMR shift difference graph with R-Styrene oxide literature (K. Sarma, N. Bahati et al.)

Comparison with S-styrene oxide NMR literature was carried out,(Synthesis 2011(7): 1092-1098) with results as follows:

13C NMR

Carbon Number δcalc (ppm) δexp (ppm) δcorr (ppm) ∆δ (ppm) Average ∆δ (ppm) Mode ∆δ (ppm) ∆δcorr (ppm) Average ∆δ (ppm) Mode ∆δ (ppm)
C2 122.9659755 128.1 -5.134024522 -2.865492337 -7.231939222 5.134024522 11.7736977 33.982242
C3 123.4152252 128.5 -5.0847748 5.0847748
C5 122.9534464 125.5 -2.546553554 2.546553554
C7 124.1326453 128.5 -4.367354719 4.367354719
C9 118.2680608 125.5 -7.231939222 7.231939222
C11 135.1343367 137.6 -2.465663325 2.465663325
C12 54.05057945 52.4 85.77702945 1.650579448 33.37702945
C13 53.455792 51.2 85.182242 2.255791999 33.982242
S-Styrene oxide carbon NMR shift difference graph with S-Styrene oxide literature (synthesis)
S-Styrene oxide carbon NMR shift difference graph with S-Styrene oxide literature (synthesis)

Theoretically, no difference should be observed between the experimental NMR of the R- and S- enantiomers. However, comparing the carbon NMR shifts obtained through B3LYP 6-311G(d,p) optimisation and simulation, the shift differences of the carbon NMR spectra with literature reveal that a closer fit is obtained with the literature S-styrene oxide NMR. (Synthesis) The disparity shown between the calculated and experimental shifts is indicative of the conformational transitions and intermolecular coupling [1] in solution that are not modelled computationally.

1H NMR

The hydrogen shifts were compared with literature experimental spectra. The S-isomer literature spectrum [30] when compared to the calculate R-isomer spectrum gave parameters as shown.

Hydrogen Number δcalc (ppm) Multiplet δcalc (ppm) δexp (ppm) δcorr (ppm) ∆δ (ppm) Average ∆δ (ppm) Mode ∆δ (ppm) ∆δcorr (ppm) Average ∆δ (ppm) Mode ∆δ (ppm)
H1 7.514585625 7.451002894 7.28 0.171002894 -0.069801232 -0.251973718 0.171002894 1.174948768 1.644740884
H4 7.483516459
H6 7.446452754
H8 7.512661831
H10 7.2977978
H15 3.662441679 3.84575 5.322108346 -0.183308321 1.476358346
H16 2.534026282 2.786 4.193692949 -0.251973718 1.407692949
H17 3.114074217 3.129 4.773740884 -0.014925783 1.644740884

This is summarised in the shift difference graph below.

S-Styrene oxide hydrogen NMR shift difference graph with literature[30]
S-Styrene oxide hydrogen NMR shift difference graph with literature[30]

Comparison with R-styrene oxide literature (K. Sarma, N. Bhati) gave a slightly closer fit of calculated hydrogen shifts to experimental shifts.

Hydrogen Number δcalc (ppm) Multiplet δcalc (ppm) δexp (ppm) δcorr (ppm) ∆δ (ppm) Average ∆δ (ppm) Mode ∆δ (ppm) ∆δcorr (ppm) Average ∆δ (ppm) Mode ∆δ (ppm)
H1 7.514585625 7.451002894 7.31 0.141002894 -0.094613732 -0.275973718 0.141002894 1.150136268 1.623740884
H4 7.483516459
H6 7.446452754
H8 7.512661831
H10 7.2977978
H15 3.662441679 3.87 5.322108346 -0.207558321 1.452108346
H16 2.534026282 2.81 4.193692949 -0.275973718 1.383692949
H17 3.114074217 3.15 4.773740884 -0.035925783 1.623740884

The closer fit is summarised in the shift difference graph below.

S-Styrene oxide hydrogen NMR shift difference graph with literature (K. Sarma, N. Bhati et al.)
S-Styrene oxide hydrogen NMR shift difference graph with literature (K. Sarma, N. Bhati et al.)

The closer fit of carbon NMR to the S-Styrene oxide literature conflicts with the closer fit of the hydrogen NMR to the R-Styene oxide literature, hinting towards the theoretical equivalence of the R- and S-enantiomer NMR spectra. Further literature would need to be carried out to gain a wider appreciation of such an effect. Additionally, it can be seen that correcting for the oxide hydrogens and carbons in fact increases disparity with literature.

NMR analysis of R-Styrene Oxide Lowest Energy Conformer

The 20.61505 kcal/mol conformer was optimised using the DFT B3LYP 6-31G(d,p) level of theory in Gaussian09 and NMR simulation calculatioins were carried out. This gave NMR spectra as shown below.

1H NMR spectra of R-styrene oxide at DFT B3LYP 6-31G(d,p)level of theory
13C NMR spectra of S-styrene oxide at DFT B3LYP 6-31G(d,p)level of theory

Atomic numbering used when referring to specified atoms in NMR analysis is as shown below.

Numbering of R-styrene oxide used during NMR analysis

Comparison of spectra with (S)-Styrene oxide

Literature NMR spectra of S-Styrene oxide, [31] was compared with the calculated NMR spectra. Results for 13C NMR spectra are as shown below in the following table and shift difference graph.

Carbon Number δcalc (ppm) δexp (ppm) δcorr (ppm) ∆δ (ppm) Average ∆δ (ppm) Mode ∆δ (ppm) ∆δcorr (ppm) Average ∆δ (ppm) Mode ∆δ (ppm)
C2 118.2706444 127.7 -9.429355626 -2.463345741 -9.429355626 -9.429355626 1.502460509 2.573722263
C3 124.1340577 128.1 -3.965942316 -3.965942316
C5 122.9525029 125.1 -2.147497079 -2.147497079
C7 123.4136518 128.1 -4.686348199 -4.686348199
C9 122.9586151 125.1 -2.141384907 -2.141384907
C11 135.1341726 137.2 -2.065827374 -2.065827374
C12 54.05586731 51.9 85.78231731 2.155867307 33.88231731
C13 53.47372226 50.9 85.20017226 2.573722263 2.573722263
Difference graph of calculated R-styrene oxide 1C NMR spectra, at DFT B3LYP 6-31G(d,p)level of theory, with experimental literature S-styrene oxide spectrum

Comparison of spectra with (R)-Styrene oxide

Literature NMR spectra of R-Styrene oxide,[32] was compared with the calculated NMR spectra. Results for 13C NMR spectra are as shown below in the following table and shift difference graph.

Carbon Number δcalc (ppm) δexp (ppm) δcorr (ppm) ∆δ (ppm) Average ∆δ (ppm) Mode ∆δ (ppm) ∆δcorr (ppm) Average ∆δ (ppm) Mode ∆δ (ppm)
C2 118.2706444 128.1 -9.829355626 -2.863345741 -9.829355626 -9.829355626 5.068266759 34.00017226
C3 124.1340577 128.5 -4.365942316 -4.365942316
C5 122.9525029 125.5 -2.547497079 -2.547497079
C7 123.4136518 128.5 -5.086348199 -5.086348199
C9 122.9586151 125.5 -2.541384907 -2.541384907
C11 135.1341726 137.6 -2.465827374 -2.465827374
C12 54.05586731 52.4 85.78231731 1.655867307 33.38231731
C13 53.47372226 51.2 85.20017226 2.273722263 34.00017226
Difference graph of calculated R-styrene oxide 1C NMR spectra, with experimental literature R-styrene oxide spectrum

IR and VCD analysis of styrene oxide

S-Styrene Oxide

IR spectrum of S-styrene oxide at DFT B3LYP 6-31G(d,p)level of theory

The IR peaks of the lowest energy S-styrene oxide enantiomer, of energy 20.61503 kcal/mol, were compared with literature,[1] with results tabulated below.

ν (cm-1 εcalc ∆νcalc-exp Average ∆ν Mode ∆ν
3031 0.8672364229 -4 0.785714286 12
3038 1.1258300005 3
2975 0.2298337644 0
2919 0 -2
2926 0 5
2933 0.0777506685 12
1603 0.3872326267 -2
1456 2.1668723937 1
1281 5.3363537201 -5
1288 24.6060915808 2
980 2.5214957845 -1
875 6.9168187410 2
763 35.0258886184 3
693 3.9809019461 -3

=R-Styrene Oxide

The IR spectrum for the R-styrene oxide lowest energy isomer was modelled at the DFT B3LYPP 6-31G(d,p) level of theory, and is as shown below.

IR spectrum of R-styrene oxide at DFT B3LYP 6-31G(d,p)level of theory

Through infrared spectroscopy, molecules absorb specific frequencies of non-polarised infrared red that are characteristic of their bonding structure, so creating normal modes with resonant frequencies related to bond strength and atomic mass of the atoms involved in bonding. For non-linear molecules like styrene-oxide, the degrees of vibrational freedom are 3N-6, where N signifies the number of atoms in the molecule, Thus, as IR spectra are indicative of bonding in the molecule, it is well documented that configurational enantiomers are indistinguishable by IR spectroscopy, due to no change in bonding structure occurring between them. However, the R-styrene oxide calculated spectrum is somewhat offset from the calculated S-styrene oxide spectrum and experimental literature spectra.[1] [1]

This disparity occurs most dramatically in the 2700-3000 cm-1 region, in which the R-isomer has an absence of characteristic peaks. This is indicative of the static nature of the calculation, which takes into account only the one supplied conformation of the configurational isomer being investigated. In reality, the experimental spectrum occurs with rapid interchange between conformers of the configuration. Thus the total bond energies are averaged over a selection of conformations, and the total bond energies of both R and S-styrene oxide have equivalence. Additionally, this disparity between IR spectra may be indicative of a poor R- or S-styrene model used as input for the NMR calculation.

VCD of S-enantiomer

The VCD of the S-enantiomer was run at the DFT B3LYP 6311++G(2df,p) level of theory also, giving a spectrum as shown.

VCD spectrum of S-styrene oxide at DFT B3LYP 6-31G(d,p)level of theory

As reported by G.-J. Kim, D.-W. Park et al.[29] the VCD of S-(-)-styrene oxide, at the maximum intensity, gives a positive intensity at 1250 cm-1, whilst R-(+)-styrene oxide gives a negative intensity of equal magnitude at equivalent frequencies. Thus, as the assigned S-enantiomer gives a maximum positive intensity in the present results, some agreement with literature is observed.[29] However, the maximum peak, lying at 1500 cm-1 is removed by ~250 cm-1 from the expected maximum peak, albeit the reduced intensities in the 1200-1000 cm-1 region agree well with literature.[29]

VCD of R-enantiomer

The VCD of the R-enantiomer was run at the DFT B3LYP 6311++G(2df,p) level of theory also, giving a spectrum as shown.

VCD spectrum of R-styrene oxide at DFT B3LYP 6-31G(d,p)level of theory

As reported by G.-J. Kim, D.-W. Park et al.[29] the VCD of R-(+)-styrene oxide, at the maximum intensity, gives a negative intensity at 1250 cm-1, whilst S-(+)-styrene oxide gives a positive intensity of equal magnitude at equivalent frequencies. Thus, as the assigned R-enantiomer gives a maximum positive intensity in the present results, some agreement with literature is observed.[29] However, the maximum peak, lying at 1500 cm-1 is removed by ~250 cm-1 from the expected maximum peak, albeit the reduced intensities in the 1200-1000 cm-1 region agree well with literature.[29]

Optical Rotations, ECD and UV-Vis of the Epoxides

For conformational isomers, stereochemistry is only important if one conformation cannot inter-convert freely with the other. Purification of one member of the enantiomeric pair or chiral diastereomers must be achieved to observe an optical rotation; members of the pair cancel each other out if purification is not achieved. Enantiomers of a known compound have optical rotations opposite in sign but identical in intensity. However it is not possible to predict the sign or magnitude of an enantiomer from its stereochemistry. Two diastereomers however have no reported relationship with respect to optical activity sign or magnitude, and have differences in other characteristic properties e.g. reactivity, melting point and solubility that cannot be predicted. For NMR and IR, enantiomers have the same spectra, but diastereomers have different spectra.


Optical Rotation of 1,2-Dihydronapthalene Oxide

The optical rotation of the 1,2-Dihydronapthalene oxide structure can be found from either the 1R,2S- or 1S,2R-enantiomer, as each is of the same magnitude. The lowest energy conformer of each configuration provides the most accurate optical rotation, at room temperature, of that configuration, as most time is spent in the lowest energy conformation, and so at any one time in a given sample, the dominant conformation exists at the energy minimum. The optical rotation was trialled for a number of isomers in order to check reliability of results. The higher energy (1S,2R)-isomer, of 30.68344 kcal/mol, was optimised to the DFT B3LYP 6-31G(d,p) level of theory, the optical rotation was calculated with the 6,311++G(2df,p) basis set. The 30.68344 kcal/mol isomer gave a negative rotation, of magnitude shown below.

[Alpha] ( 5890.0 A) = -155.82⁰
[Alpha] ( 3650.0 A) = -522.16⁰

The 30.2244 kcal/mol isomer, the lowest energy (1S,2R)-isomer, and thus the most accurate approximation for optical rotation was optimised to the DFT B3LYP 6-31G(d,p) level of theory and the optical rotation was calculated with the 6,311++G(2df,p) basis set, giving a negative rotation of magnitude shown below.

[Alpha] ( 5890.0 A) = -155.82⁰
[Alpha] ( 3650.0 A) = -522.15⁰

To increase accuracy of this calculation further the structure was optimised further using the higher force field cc-pVTZ and optical rotation simulation carried out. This returned reduced o.r as below.

[Alpha] ( 5890.0 A) = -155.22⁰
[Alpha] ( 3650.0 A) = -516.50⁰

The negative rotation for the 1S,2R-isomer compares favourably with literature reports,[33] but indicates the %ee assignment in such literature may be drawn to error due to contaminant molecules present in so-called "enantiopure" references. For example, S. Pedragosa-Moreau et al.[33] reported a 99% ee with o.r +129 for the (1R,2S)-isomer. This result implies a 100% enantiopure substance of o.r 130.303⁰, a value which is 25.17⁰ and 24.92⁰ lower than that calculated using the 30.2244 kcal/mol isomer, with input optimisations of DFT B3LYP 6-31G(d,p) and cc-pVTZ level of theory respectively. Indeed, the higher energy 30.68344 kcal/mol input isomer at optimised to B3LYP 6-31G(d,p) level of theory also returns an enantiopure o.r greater than the experimental reference. As calculations on the lowest energy isomer at both levels of theory return an enantiopure optical rotation greater in magnitude than the experimental enantiopure reference, this may indicate that the true enantiopure o.r lies significantly higher than the 130.303⁰ reported. However, optical activity is not an individual molecular property, but results from the random distribution of sample molecules. The calculation performed presently is performed on one static molecule only, and does not take into account inter-molecular and packing effects, which may cause changes to the molecular geometry of the oxides and thus cause deviations from the single molecule enantiopure o.r. Additionally, o.r measurements are sensitive to temperature fluctuations, and any disparity between the calculation parameters and the environment of the enantiopure substance may cause the observed experimental o.r reduction.

Optical Rotation of Styrene Oxide

Optical Rotation of S-(-)-Styrene Oxide

Using the NMR of the S-Styrene oxide, the lowest energy structure found in the present study, optimised to the DFT B3LYP 6-31G(d,p) level of theory, the optical rotation was calculated with the 6,311++G(2df,p) basis set. In accordance with literature,[34] [35]this gave a negative rotation for S-styrene oxide, of magnitude shown below.

[Alpha] ( 589.0 nm) = -30.07⁰
[Alpha] ( 365.0 nm) = -93.90⁰

This o.r compared poorly with literature values of ~23⁰,[34] [35] and so the enantiomer was optimised further to the cc-pVTZ level of theory. Using the NMR of this S-Styrene oxide optimised to the DFT B3LYP cc-pVTZ level of theory, the optical rotation was calculated with the 6,311++G(2df,p) basis set. The S-conformer gave a negative rotation, of magnitude shown below.

[Alpha] ( 589.0 nm) = -22.52⁰
Alpha] ( 365.0 nm) = -69.26⁰

The optical rotation at the higher level of theory agrees well with a -22.5⁰ enantiomeric pure reference in literature at 26⁰C[34], but differs by 1.42⁰ from reports by D. E. White et al. with an enantiomeric pure reference of -23.94⁰ at a temperature of 31⁰C.[34] [36] These differences may arise from temperature differences at measurement of the optical rotation but may also arise from an enantiomeric contamination present for the higher optical rotation. The present optical rotation is reduced compared to those obtained in previous reports due to the averaging effects of conformational transitions in solution, which are not accounted for by the static conformer model. Experimentally, this results in an average optical rotation of the conformations involved in conformational transitions. The present calculation has taken into account only one conformation. To reduce the effects of conformational transitions during computational analysis, the lowest energy conformer found for the S-Styrene oxide product was input into the optical rotation analysis as, this being the lowest energy conformer found presently, it was expected to exist predominantly in solution at any given time, and so give the greatest contribution to the optical rotation. However, the conformation found presently may reside at a local energy minima, in which case the trialed conformer would not be present as the predominant conformation.

Optical Rotation of R-(+)-Styrene Oxide

Using the NMR of R-Styrene oxide optimised to the DFT B3LYP 6-31G(d,p) level of theory, the optical rotation was calculated with the 6.311++G(2df,p) basis set. The higher energy 21.61547 kcal/mol configuration gave a positive rotation, of magnitude shown below.

[Alpha] ( 589.0 A) = 30.17⁰ [Alpha] ( 365.0 A) = 94.24⁰

Using the NMR of Styrene oxide optimised to the DFT B3LYP cc-pVTZ level of theory, the optical rotation was calculated with the 6,311++G(2df,p) basis set. The higher energy 21.61547 kcal/mol configuration gave a positive rotation, of magnitude shown below.

[Alpha] ( 589.0 A) = 22.47⁰ [Alpha] ( 365.0 A) = 69.10⁰

The R-styrene oxide conformer found with lower energy, 21.61506 kcal/mol was optimised to the DFT B3LYP 6-31G(d,p) level of theory, the optical rotation was calculated with 6.311++G(2df,p) basis set.

[Alpha] ( 589.0 A) = 30.43⁰ [Alpha] ( 365.0 A) = 95.05⁰

This conformer was optimised to the DFT B3LYP cc-pVTZ level of theory, and the optical rotation was calculated with 6.311++G(2df,p) basis set, giving a closer fit with literature.[37]

[Alpha] ( 589.0 A) = 23.06⁰ [Alpha] ( 365.0 A) = 71.02⁰

In general for the Styrene oxide enantiomers, the lower energy conformers give the best fit with literature, and increasing the optimisation level of theory to the cc-pVTZ level increases the agreement with literature.[37]

ECD and UV-Vis Simulations

Electronic Circular Dichroism spectroscopy (ECD) can be used to determine the absolute structure of solid and solution-phase molecules. For determining molecular 3D structure, configuration and conformation, it is a sensitive, quick and reliable technique, provided that the substance has chirality, is non-racemic, and absorbs circularly polarized light (CPL) in 180 - 1100 nm range available.[38]

ECD and UV-Vis of 1,2-Dihydronapthalene Oxide

ECD and UV-Vis simulation were run on the 30.22441 kcal/mol configuraiton isomer, at DFT B3LYP 6-31G level of theory giving spectra as shown below.

  • UV-vis napthalane lowest energy conformer.svg

Excitation Energy (nm) = 192.5000000000 ɛ = 57629.5162835934 d(epsilon)/d(excitation energy) = -294.5415960797

  • Ecd spectrum napthalene lowest energy.svg

Excitation Energy (nm)= 196.0000000000 ∆ɛ = 9.0192453436

ECD and UV-Vis simulations were also run on the lowest energy conformer, the 30.2244 kcal/mol of the (-)-1,2-dihydronapthalene oxide configuration, which was deemed to be a more accurate representation of the molecule in solution. Noted are the maxima and minima for the relevant spectra of the structure optimised to the B3LYP 6-31G(d,p) level of theory are as below.

  • UV-Vis spectrum:
UV-Vis at 6-31G level of theory for (-)-enantiomer
UV-Vis at 6-31G level of theory for (-)-enantiomer

Excitation Energy (nm)= 192.02
Oscillator Strength = 0.7778

Comparing this UV-Vis maximum with that obtained for the higher energy (+)-isomer at the same level of theory, there is a 0.48 nm decrease in absorption maximum associated with the 0.00001 kcal/mol decrease in initial input energy.

  • ECD spectrum:
ECD at 6-31G level of theory for (-)-enantiomer
ECD at 6-31G level of theory for (-)-enantiomer

Excitation Energy (nm)= 187.58
Rvel (10-40 esu2cm2 = -47.1
∆ɛ = ~-8

Excitation Energy (nm)= 173.5
Rvel (10-40 esu2cm2 = 26.872
∆ɛ = ~4.5

To increase the accuracy of these calculations, the structure was optimised to B3LYP cc-pVTZ level of theory and ECD and UV-Vis simulations run. Noted are the maxima and minima for the relevant spectra of the structure optimised to the cc-pVTZ level of theory are as below.

  • UV-Vis spectrum:
UV-Vis at cc-pVTZ level of theory for (-)-enantiomer
UV-Vis at cc-pVTZ level of theory for (-)-enantiomer

Excitation Energy (nm)= 190.84
Oscillator Strength = 0.7719

Thus increasing the force field increases the energy of the maximum absorption excitation energy, which is to be expected given the reduction in total energy that the cc-pVTZ force field brings about during optimisation c.f. the 6-31G force field.

  • ECD spectrum:

Excitation Energy (nm)= 158.78
Rvel (10-40 esu2cm2 = 0.441
∆ɛ = ~-3

ECD at cc-pVTZ level of theory for (-)-enantiomer
ECD at cc-pVTZ level of theory for (-)-enantiomer

Excitation Energy (nm)= 173.53
Rvel (10-40 esu2cm2 = 7.1378
∆ɛ = ~3

Excitation Energy (nm)= 187.18
Rvel (10-40 esu2cm2 = -530.165
∆ɛ = ~-8

At the higher level of theory, one extra peak at a higher energy of 158.78 nm is observed, and all peaks are shifted to lower energy.

The ECD spectra are in general the mirror image of the UV-Vis spectrum, as expected, and each configurational isomer ECD spectrum is the mirror image of the other. However, no literature could be found for either enantiomer to compare with experimental values. Literature reports involving napthalene derivatives exhibit ECD peaks within the 200-250 nm region.[39] The calculated values at the DFT B3LYP 6-31G level of theory lie in the 150-200 nm region, and thus provide a reasonable similarity given the errors associated with such a comparison.

ECD and UV-Vis of Styrene Oxide

Using the NMR calculation carried out on the R-styrene oxide highest energy configurational isomer, at the DFT B3LYP 6-31G(d,p) level of theory, ECD and UV-Vis simulations were run, giving results as shown below.

  • UV-Vis Spectrum
Caption
Caption

ʎmax Excitation Energy (nm) = 187.98
Oscillator Strenght = 0.867
∆ɛ = ~59000

  • ECD Spectrum
Caption
Caption

ʎmax Excitation Energy (nm)= 185.54
∆ɛ = ~-12
Rvel = ~-30

ʎmax Excitation Energy (nm)= 160.09
∆ɛ = ~15
Rvel = ~40

Using the NMR calculation carried out on the S-styrene oxide configurational isomer, of lowest energy, at the DFT B3LYP 6-31G(d,p) level of theory, ECD and UV-Vis simulations were run, giving results as shown below.

  • UV-Vis Spectrum
Caption
Caption

ʎmax Excitation Energy (nm) = 188.2400000000
∆ɛ = 58894.6631990984
d∆ɛ/dExcitation Energy =320.5154439262

  • ECD Spectrum
Caption
Caption

ʎmax Excitation Energy (nm)= 185.6000000000
∆ɛ = 13.8016611316
d∆ɛ/dExcitation Energy = -0.4784170592

ʎmin Excitation Energy (nm) = 160.0800000000
∆ɛ = -15.8501295345
d∆ɛ/dExcitation Energy = 0.3810660432

Suggesting New Candidates for Investigation=

A suitable epoxide of o.r > 500 or < 500 is provided by 2-(4-methylphenyl)-3-phenyl oxirane trans-(2R,3R)-2-(4-methylphenyl)-3-phenyl oxirane [Alpha] ( 5890.0 A) = -320.35˚ [Alpha] ( 4360.0 A) = -730.74˚

cis [Alpha] ( 5890.0 A) = -25.48 deg [Alpha] ( 4360.0 A) = -49.25 deg


QTAIM image 4

Dansette,P.M. et al. Tetrahedron, 1976 , vol. 32, p. 2071 - 2074


The trans-(2R,3R)-2-(4-methylphenyl)-3-phenyl oxirane was optimised to a minimum in Avogadro using the MMF94s force field and conjugate gradients. The structure and parameters of the optimised molecule are shown below.

Trans methyl stilbene for hpc.cml

Energy Contribution Energy (kcal/mol)
Total Electrostatic Energy 5.04177
Total Out-of-Palne Bending Energy 0.0201
Total Torsional Energy 2.17991
Total Stretch Bending Energy -1.50356
Total Angle Bending Energy 2.83423
Total Bond Stretching Energy 3.9429
Total Energy 41.55575

This structure was submitted to optimisation at the DFT B3LYP 6-31G(d,p) level of theory, giving a structure and energy as shown below.

Sum of electronic and zero-point Energies= -655.026348

Sum of electronic and thermal Energies=              -655.012762
Sum of electronic and thermal Enthalpies=            -655.011818
Sum of electronic and thermal Free Energies=         -655.068821

The enantiomeric (1S,2R)-cis-2-(4-methylbenzenyl)-3-phenyl-oxirane, with a lower reported optical rotation, was optimised in Avogadro with the MMFF94s force field using the conjugate gradient algorithm. The structure and parameters of the lowest energy structure obtained using this technique are as below.

Trans methyl stilbene opt 4 (cis) new one.cml

'
Energy Contribution Energy (kcal/mol)
Total Electrostatic Energy 6.56941
Total Van der Waals Energy 28.65841
Total Out-of-Palne Bending Energy 0.02091
Total Torsional Energy 2.1782
Total Stretch Bending Energy -1.3851
Total Angle Bending Energy 3.20829
Total Bond Stretching Energy 4.08582
Total Energy 43.33594

NMR analysis was carried out, giving good agreement with literature for most carbon shifts, as documented below.

Carbon Number δcalc (ppm) δexp (ppm) ∆δcalc-exp (ppm) Average ∆δcalc-exp (ppm) Mode ∆δcalc-exp (ppm)
C1 122.9778008 125.5 -2.522199225 -4.017924319 -7.153745456
C2 123.513071 128.5 -4.986928969
C3 123.0827678 128.2 -5.117232162
C4 124.1918224 129.2 -5.008177643
C5 118.2462545 125.4 -7.153745456
C6 134.2028056 138.2 -3.997194358
C7 124.2666619 129.2 -4.933338102
C8 123.100 128.2 -5.100009147
C9 131.2212506 138.2 -6.978749362
C10 118.4044517 125.4 -6.99554825
C11 124.9491655 137.2 -12.25083451
C12 134.7213106 138.2 -3.478689364
C13 66.42459374 62.8 3.624593736
C14 66.27088104 62.7 3.57088104
C16 22.25830698 21.2 1.058306982

This was optimised to the DFT B3LYP 6-31G(d,p) level of theory, giving a structure of energetic parameters, expressed in Hartree/particle, as below.

Thermal correction to Gibbs Free Energy=               0.204857
Sum of electronic and zero-point Energies= -655.023912
Sum of electronic and thermal Energies= -655.010405
Sum of electronic and thermal Enthalpies= -655.009460
Sum of electronic and thermal Free Energies= -655.066618


  1. 1.0 1.1 1.2 1.3 1.4 1.5 1.6 W. T. Klooster , T. F. Koetzle , P. E. M. Siegbahn , T. B. Richardson , and R. H. Crabtree, "Study of the N-H···H-B Dihydrogen Bond Including the Crystal Structure of BH3NH3 by Neutron Diffraction", J. Am. Chem. Soc., 1999, 121, 6337–6343.DOI:10.1021/ja9825332 Cite error: Invalid <ref> tag; name "ja9825332" defined multiple times with different content
  2. 2.0 2.1 2.2 J. Sauer, R. Sustmann, Angew. Chem. Int. Ed. Engl., 1980, 19, 779 - 807.DOI:[1] Cite error: Invalid <ref> tag; name "ja9825334" defined multiple times with different content
  3. K. Alder, G. Stein , Angew. Chem., 1937, 50, 510.DOI:[2]
  4. L. Paquette, N. A. Pegg, D. Toops, G. D. Maynard, R. D. Rogers, J. Am. Chem. Soc., 1990, 112, 277-283.DOI:10.1021/ja9825332
  5. 5.0 5.1 5.2 5.3 S. W. Elmore, L. Paquette, Tet. Lett., 1991, 319, 1891DOI:10.1016/S0040-4039(00)92617-0
  6. 6.0 6.1 W. F. Maier, P.-V. R. Schleyer,J. Am. Chem. Soc., 1981, 103, 1891DOI:10.1021/ja00398a003
  7. C. Braddock, H. S. Rzepa, J. Nat. Prod., 2008, 71, 728-730DOI:10.1021/np0705918
  8. M. Durík, V. Langer, D. Gyepesová, J. Micová, B. Steiner M. Koós, Acta Cryst.,2001,E57, 672-674DOI:10.1107/S160053680101073X 10.1107/S160053680101073X
  9. 9.0 9.1 Z.-X. Wang , S. M. Miller, O. P. Anderson, Y. Shi, J. Org. Chem.,2001, 66, 2, 521–530.DOI:10.1021/jo001343i 10.1021/jo001343i
  10. 10.0 10.1 10.2 R. V. Stick, S. J. Williams, Carbohydrates, Amsterdam, Elsevier, 2009.Template:ISBN Cite error: Invalid <ref> tag; name "ja9825335" defined multiple times with different content
  11. 11.0 11.1 11.2 J. Clayden, N. Greeves, S. Warren, P. Wothers, Organic Chemistry, Oxford University Press,2nd Edition, 2012, 1126-1131.Template:ISBN
  12. Z.-X. Wang , S. M. Miller, O. P. Anderson, Y. Shi, J. Org. Chem.,2001, 66, 2, 521–530.DOI:10.1021/jo001343i 10.1021/jo001343i
  13. J. W. Yoon, T.-S. Yoon, S. W. Lee, W. Shin, Acta Cryst.,1999, C55, 1766-1769 .DOI:10.1107/S0108270199009397
  14. F. R. Fronczek, X. R. J. Bu, Private Comm.,2002DOI:[3]
  15. 15.0 15.1 E. Anslyn, Modern Physical Organic Chemistry, 2004, Sausalito, CA, University ScienceTemplate:ISBN Cite error: Invalid <ref> tag; name "ja9825332123" defined multiple times with different content
  16. Z.-X. Wang,Y. Shi, J. Org. Chem. 1997, 62, 8622-8623DOI:S0022-3263(97)01701-5
  17. 17.0 17.1 17.2 17.3 G.-A. Cao, Z.-X. Wang, Y. Tu, Y. Shi, Tetrahedron Letters, 1998, 39, 25, 4425–4428DOI:10.1016/S0040-4039(98)00838-7 Cite error: Invalid <ref> tag; name "ja982533672123" defined multiple times with different content
  18. M. Hickey, D. Goeddel, Z. Crane, Y. Shi, Proc. Natl. Acad. Sci. USA; 2004, 101, 16, 5794–5798DOI:10.1073/pnas.0307548101
  19. J. L. Arbour, H. S. Rzepa, J. Contreras-García, L. A. Adrio, E. M. Barreiro, K. K. Hii, Chemistry - A European Journal, 2012, 18, 36, 11317–11324DOI:10.1002/chem.201200547
  20. G. A. Morris, S. B. T. Nguyen, J. T. Hupp, J. Mol. Cat. A: Chemical 174, 2001, 15–20DOI:S1381-1169(01)00165-0
  21. W. Zhang, J. L. Loebach, S. R. Wilson, E. N. Jacobsen, J. Am. Chem. Soc., 1990, 112, 7, 2801-2803 DOI:10.1021/ja00163a052
  22. M. Palucki, P. J. Pospisil, W. Zhang, E. N. Jacobsen, J. Am. Chem. SOC. 1994,116, 9333-9334DOI:10.1021/ja00099a062
  23. C. Altona, J. H. Ippel, A. J. A. W. Hoekzema, C. Erkelens, M. Groesbeek, L. A. Donders, A. Lambertus, Magnetic Resonance in Chemistry, 1989 , 27, 6, 564 - 576DOI:10.1002/mrc.1260270609
  24. 24.0 24.1 P. C. Bulman Page, B. R. Buckley ,H. Heaney, A. J. Blacker, Org. Lett., 2005, 7, 3, 375–377DOI:10.1021/ol047836h
  25. 25.0 25.1 25.2 L. Paquette, N. A. Pegg, D. Toops, G. D. Maynard, R. D. Rogers, J. Am. Chem. Soc., 1990, 112, 277-283 DOI:10.1021/ja00157a043
  26. 26.0 26.1 26.2 D. Xiong, X. Hu, S. Wang, C.-X. Miao, C. Xia, W. E. Sun, J. Org. Chem., 2011, 23, 4289–4292DOI:[4]
  27. 27.0 27.1 P. C. B. Page, B. R. Buckley, H. Heaney , A. J. Blacker, Org. Lett., 2005, 7, 3, 375–377 DOI:10.1021/ol047836h
  28. C. Guo, R. D. Shah, R. K. Dukor, X. Cao, T. B. Freedman , L. A. Nafie, Anal. Chem., 2004, 76, 23,6956–6966DOI:10.1021/ac049366a
  29. 29.0 29.1 29.2 29.3 29.4 29.5 29.6 G.-J.Kim, D.-W. Park, Y.-S. Tak, Catalysis Letters, 2000, 65, 1-3, 127-133DOI:10.1023/A:1019040532103
  30. 30.0 30.1 D. Xiong, X. Hu, S. Wang, Shoufeng, C.-X. Miao, C. Xia, W. Sun, European Journal of Organic Chemistry, 2011 , 23, 4289 - 4292
  31. Y. Kon, H. Hachiya, Y. Ono, T. Matsumoto, K. Sato, Synthesis, 2011, 7, 1092-1098DOI:[5]
  32. K. Sarma,N. Bhati, N. Borthakur, A. Goswami, Tetrahedron, 2007, 63, 36, 8735,-8741
  33. 33.0 33.1 S. Pedragosa-Moreau, A Archelas, R Furstoss, Tetrahedron, 1996, 52, 13, 4593–4606 DOI:10.1016/0040-4020(96)00135-4
  34. 34.0 34.1 34.2 34.3 M. Hayashi, H. Miwata, N. Oguni, J. Chem. Soc., Perkin Transactions 1: Organic and Bio-Organic Chemistry, 1991, 5, 1167-1171 DOI:[6]
  35. 35.0 35.1 V. Capriati, S. Florio, R. Luisi, A. Salomone, Org. Lett., 2002, 4, 14, 244-2448DOI:10.1021/ja9825332
  36. D. E. White, E. N. Jacobsen, Tetrahedron: Asymmetry, 2003, 14, 22, 3633-3638DOI:10.1016/j.tetasy.2003.09.024
  37. 37.0 37.1 R. Imae, N. Itoh, H. Toda, Tetrahedron Asymm., 2012, 23, 22-23, 1542-1549DOI:[7]
  38. P. J. Stephens, F. J. Devlin, J. R. Cheeseman, VCD Spectroscopy for Organic Chemists, CRC Press, 2012, 1-30Template:ISBN
  39. B. M. Bulheller, G. D. Pantoş, J. K. M. Sanders, J. D. Hirst, Phys. Chem. Chem. Phys., 2009,11, 6060-6065DOI:10.1039/B905187B