A-Values are numerical values used in the determination of the most stable orientation of atoms in a molecule (Conformational Analysis), as well as a general representation of steric bulk. A-values are derived from energy measurements of a monosubstituted cyclohexane ring. [1] Substituents on a cyclohexane ring prefer to reside in the equatorial position to the axial. The difference in Gibbs free energy (ΔG) between the higher energy conformation (axial substitution) and the lower energy conformation (equatorial substitution) is the A-value for that particular substituent.

The A-value for a methyl group is 1.74 as derived from the equilibrium above. This means it costs 1.74 kcal/mol of energy to have a methyl group in the axial position compared to the equatorial position.

Utility edit

A-values help predict the conformation of cyclohexane rings. The most stable conformation will be the one which has the substituent or substituents equatorial. When multiple substituents are taken into consideration, the conformation where the substituent with the largest A-value is equatorial is favored.

 
A methyl substituent has a significantly smaller A-value than a tert-butyl substituent; therefore the most stable conformation has the tert-butyl in the equatorial position.

The utility of A-values can be generalized for use outside of cyclohexane conformations. A-values can help predict the steric effect of a substituent. In general, the larger a substituent’s A-value, the larger the steric effect of that substituent. Methyl has an A-value of 1.74 while tert-butyl has an A-value of ~5. Because the A-value of tert-butyl is higher, tert-butyl has a larger steric effect than methyl. This difference in steric effects can be used to help predict reactivity in chemical reactions.

Free Energy Considerations edit

Sterics play a major role in the assignment of configurations in cyclohexanes. One can use steric hindrances to determine the propensity of a substituent to reside in the axial or equatorial plane. It is known that axial bonds are more hindered than the corresponding equatorial bonds. This is because substituents in the axial position are relatively close to two other axial substituents. This makes it very crowded when you have bulky substituents in the axial position, these types of steric interactions are commonly known as 1,3 diaxial interactions.[2] These types of interactions are not present with substituents at the equatorial position. There are generally considered three principle contributions to the conformational free energy: [3] 1.Bayer Strain, defined as the strain arising from deformation of bond angles. 2. Pitzer Strain, defined as the torsional strain arising from 1,2 interactions between groups attached to contiguous carbons, 3. Van der Waals interactions, which are similar to 1,3 diaxial interactions.

Enthalpic Components edit

When comparing relative stability, 6- and 7-atom interactions can be used to approximate differences in enthalpy between conformations. Each 6-atom interaction is worth 0.9 kcal/mol and each 7-atom interaction is worth 4 kcal/mol. [4]

 
The dashed lines indicate 6-atom interactions found in this conformation of ethyl cyclohexane, which amounts to approximately 2.7 kcal/mol in the enthalpic term of Free Energy.
 
The dashed lines here signify the 7 atom interactions , which contribute approximately 8 kcal/mol to the enthalpic term making this conformation unrealistically high in energy.

Entropic Components edit

Entropy also plays a role in a substituent’s preference for the equatorial position. The entropic component is determined by the following formula:

 

Where σ is equal to the number of micro states available for each conformation.

 
Possible axial conformations of ethyl cyclohexane.
 
Possible equatorial conformations of ethyl cyclohexane.

Due to the larger number of possible conformations of ethyl cyclohexane, the A value is reduced from what would be predicted based purely on ethalpic terms. Due to these favorable entropic conditions, the steric relevance of an ethyl group is similar to that of a methyl substituent.

Table of A-Values edit

The following table lists some common A-values in kcal/mol:[5][6][7][8]

Substituent A-Value Substituent A-Value Substituent A-Value
D 0.006 CH2Br 1.79 OSi(CH3)3 0.74
F 0.15 CH(CH3)2 2.15 OH 0.87
Cl 0.43 c-C6H11 2.15 OCH3 0.6
Br 0.38 C(CH3)3 >4 OCD3 0.56
I 0.43 Ph 3 OCH2CH3 0.9
CN 0.17 C2H 1.35 O-Ac 0.6
NC 0.21 CO2- 1.92 O-TFA 0.68
NCO 0.51 CO2CH3 1.27 OCHO 0.27
NCS 0.28 CO2Et 1.2 O-Ts 0.5
N=C=NR 1 CO2iPr 0.96 ONO2 0.59
CH3 1.7 COCl 1.25 NH2 1.6
CF3 2.1 COCH3 1.17 NHCH3 1
CH2CH3 1.75 SH 0.9 N(CH3)3 2.1
CH=CH2 1.35 SMe 0.7 NH3+ 1.9
CCH 0.41 SPh 0.8 NO2 1.1
CH2tBu 2 S- 1.3 HgBr ~0
CH2OTs 1.75 SOPh 1.9 HgCl 0.3
SO2Ph 2.5 Si(CH3)3 2.5

Applications edit

Predicting Reactivity edit

One of the original experiments performed by Winston and Holness6 was measuring the rate of oxidation in trans and cis substituted rings using a Chromium catalyst. The large tBu group used locks the conformation of each molecule placing it equatorial (cis compound shown).

 
Possible chair conformations of cis 4-tert-butyl-1-cyclohexanol


It was observed that the cis compound underwent oxidation at a much faster rate than the trans compound. The proposition was that the large hydroxyl group in the axial position was disfavored and formed the carbonyl more readily to relieve this strain. The trans compound had rates identical to those found in the monosubstituted cyclohexanol.

 
Chromium oxidation of cis 4-tertbutyl-1-cyclohexanol

Approximating Intramolecular Force Strength Using A-Values edit

Using the A-Values of the hydroxyl and isopropyl subunit, the energetic value of a favorable intramolecular hydrogen bond can be calculated.[9]

 
Possible chair conformations and the favorable hydrogen bond available in the conformation where both hydroxyl substituents are equatorial

Limitations edit

A-Values are measured using a mono-substituted cyclohexane ring, and are an indication of only the sterics a particular substituent imparts on the molecule. This leads to a problem when there are possible stabilizing electronic factors in a different system. The carboxylic acid substituent shown below is axial in the ground state, despite a positive A-Value. From this observation, it is clear that there are other possible electronic interactions that stabilize the axial conformation.

 
Equilibrium representation of a chair flip of a Carboxylic Acid. The axial position is preferred due to favorable electronic factors, despite a steric bias favoring the equatorial position.

Other Considerations edit

It is important to note that A-values do not predict the physical size of a molecule, only the steric effect. For example, the tert-butyl group (A-value=4.9) has a larger A-value than the trimethylsilyl group (A-value=2.5), yet the tert-butyl group actually occupies less space. This difference can be attributed to the longer length of the carbon-silicon bond as compared to the carbon-carbon bond of the tert-butyl group. The longer bond allows for less interactions with neighboring substituents, which effectively makes the trimethylsilyl group less sterically hindering, thus, lowering it’s A-value.[2] This can also be seen when comparing the halogens. Bromine, iodine, and chlorine all have similar A-values even though their atomic radii differ.[4] A-values then, predict the apparent size of a substituent, and the relative apparent sizes determine the differences in steric effects between compounds. Thus, A-values are useful tools in determining compound reactivity in chemical reactions.

References edit

  1. ^ Muller, P (1994). "Glossary of terms used in physical organic chemistry (IUPAC Recommendations 1994)". PAC. 66: 1077–1184. doi:10.1351/pac199466051077.
  2. ^ a b Hoffman, Robert V. (2004). Organic Chemistry (second ed.). New Jersey: John Wiley and Sons, Inc. p. 167. {{cite book}}: Unknown parameter |trans_title= ignored (|trans-title= suggested) (help)
  3. ^ Anderson, J. Edgar (1974). Dynamic Chemistry. Springer-Verlag. p. 139. doi:10.1007/3-540-06471-0. {{cite book}}: Unknown parameter |trans_title= ignored (|trans-title= suggested) (help)
  4. ^ a b Anslyn, Eric V.; Dougherty, Dennis A. (2006). Modern Physical Organic Chemistry. Sausalito, CA: University Science Books. pp. 104–105.
  5. ^ Eliel, E.L.; Wilen, S.H. (1994). Stereochemistry of Organic Compounds. New York: John Wiley & Sons, Inc. {{cite book}}: line feed character in |publisher= at position 19 (help)
  6. ^ Eliel, E.L.; Allinger, N.L.; Angyal, S.J.; G.A., Morrison (1965). Conformational Analysis. New York: Interscience Publishers.
  7. ^ Hirsch, J.A. (1967). Topics in Stereochemistry (first ed.). New York: John Wiley & Sons,Inc. p. 199.
  8. ^ Romers, C.; Altona, C.; Buys, H.R.; Havinga, E. (1969). Topics in Stereochemistry (fourth ed.). New York: John Wiley & Sons,Inc. p. 40.
  9. ^ Huang, C.-Y.; C.A.; Anslyn, E.V. (1994). JACS. 116: 2776–2792. {{cite journal}}: Missing or empty |title= (help)