14 Chapter 14 – Conjugated Compounds and Ultraviolet Spectroscopy

Chapter 14 – Conjugated Compounds and Ultraviolet Spectroscopy

Solutions to Problems

14.1 We would expect ΔHhydrog = –126 + (–126) = –252 kJ/mol for allene if the heat of hydrogenation for each double bond were the same as that for an isolated double bond. The measured ΔHhydrog, –298 kJ/mol, is 46 kJ/mol more negative than the expected value. Thus, allene is higher in energy (less stable) than a nonconjugated diene, which in turn is less stable than a conjugateddiene.

 

14.2  

image

 

14.3  

image

A and D, which are resonance-stabilized, are formed in preference to B and C, which are not. The positive charge of allylic carbocation A is delocalized over two secondary carbons, while the positive charge of carbocation D is delocalized over one secondary and one primary carbon. We therefore predict that carbocation A is the major intermediate formed, and that 4-chloro-2-pentene predominates. Note that this product results from both 1,2 and 1,4 addition.

 

14.4  

image

 

14.5  

image

Allylic halides can undergo slow dissociation to form stabilized carbocations (SN1 reaction). Both 3-bromo-1-butene and 1-bromo-2-butene form the same allylic carbocation, pictured above, on dissociation. Addition of bromide ion to the allylic carbocation then occurs to form a mixture of bromobutenes. Since the reaction is run under equilibrium conditions, the thermodynamically more stable 1-bromo-2-butene predominates.

 

14.6  

image

1,4 adducts are more stable than 1,2 adducts because disubstituted double bonds are more stable than monosubstituted double bonds (see Chapter 7).

 

14.7 Draw the reactants in an orientation that shows where the new bonds will form. Form the new bonds by connecting the two reactants, removing two double bonds, and relocating the remaining double bond so that it lies between carbon 2 and carbon 3 of the diene. The substituents on the dienophile retain their trans relationship in the product. The product is a racemic mixture.

image

 

14.8 Good dienophiles have an electron-withdrawing group conjugated with a double bond.

(a) (d)
Good dienophiles: image image
Poor dienophiles: (b) (c)
image image
(e)

image

Compounds (a) and (d) are good dienophiles because they have electron-withdrawing groups conjugated with a carbon–carbon double bond. Alkene (c) is a poor dienophile because it has no electron-withdrawing functional group. Compounds (b) and (e) are poor dienophiles because their electron-withdrawing groups are not conjugated with the double bond.

14.9  

(a) This diene has an s-cis conformation and should undergo Diels–Alder cycloaddition.
(b) This diene has an s-trans conformation. Because the double bonds are in a fused ring system, it is not possible for them to rotate to an s-cis conformation.
(c) Rotation can occur about the single bond of this s-trans diene. The resulting s-cis, however, has an unfavorable steric interaction of a methyl group with a hydrogen at carbon 1. Rotation to the s-cis conformation is possible but not favored.
image

 

14.10 Rotation of the diene to the s-cis conformation must occur in order for a reaction to take place.

image

 

14.11 The initiator may be either a radical or a cation. Diene polymerization is a 1,4 addition process that forms a polymer whose monomer units have a 4 carbon chain that contains a double bond every 4 bonds.

image

 

14.12  

image

 

14.13  

The energy of electromagnetic radiation in the region of the spectrum from 200 nm to 400 nm is 300–600 kJ/mol.

The energy required for UV transitions is greater than the energy required for IR or 1H NMR transitions.

14.14  

 

14.15 All compounds having alternating single and multiple bonds should show ultraviolet absorption in the range 200–400 nm. Only compound (a) is not UV-active. All of the compounds pictured below (b, c, d, e, and f) show UV absorptions.

image

 

Additional Problems

Visualizing Chemistry

14.16  

image

 

14.17  

image

 

14.18 In order to undergo Diels–Alder reaction, this s-trans diene would have to rotate to an s– cis arrangement. In an s-cis conformation, however, the two circled methyl groups experience steric strain by being too close to each other, preventing the molecule from adopting this conformation. Thus, Diels–Alder reaction doesn’t occur.

image

 

14.19  

image

Mechanism Problems

14.20 Note: The major product of the reaction depends on the reaction temperature. At higher temperatures (> 40 °C) the more substituted akene (the thermodynamic product) is favored, while at lower temperatures (0 °C) the less substituted alkene (the kinetic product) is favored.

(a)

image

Mechanism:

image

(b)

image

Mechanism:

image

(c)

image

Mechanism:

image

 

14.21 Diels–Alder reactions are reversible when the products are much more stable (of lower energy) than the reactants. In this case, the reactant is a nonconjugated diene, and the products are benzene (a stable, conjugated molecule) and ethylene.

image

 

14.22  

(a) A Diels–Alder reaction between α-pyrone (diene) and the alkyne dienophile yields the following product.

image

The double bonds in this product are not conjugated, and a more stable product can be formed by loss of CO2.

image

This process can occur in a manner similar to the reverse Diels–Alder reaction of the previous problem.

(b) image

Mechanism:

 

14.23  

image

Mechanism:

image

 

14.24  

image

Mechanism:

image

Conjugated Dienes

14.25 All of these compounds can exhibit E/Z isomerism.

(a) image (b) image
(c) image (d) image

 

14.26 Excluding double-bond isomers:

image

 

14.27  

(a)

image

(b)

image

(c)

image

(d)

image

(e)

image

(f)

image

 

14.28  

image

Tertiary/primary allylic carbocation A is more stable than secondary/primary allylic carbocation B. Since the products formed from the more stable intermediate predominate, 3,4-dibromo-3-methyl-1-butene is the major product of 1,2 addition of bromine to isoprene. In both cases, the product with the more substituted double bond (1,4 addition product) predominates.

 

14.29 Any unsubstituted cyclic 1,3-diene cyclic diene gives the same product from 1,2- and 1,4 addition. For example:

image

 

14.30  

image

image

Carbocation D is most stable because it can use the π systems of both the benzene ring and the allylic side chain to delocalize positive charge. 3-Chloro-1-phenyl-1-butene is the major product because it results from cation D and because its double bond can be conjugated with the benzene ring to provide extra stability.

Diels–Alder Reactions

14.31  

(a)

image

(b)

image

If two equivalents of cyclohexadiene are present for each equivalent of dienophile, you can also obtain a second product:

image

 

14.32 This conformation of 2,3-di-tert-butyl-1,3-butadiene, in which the tert-butyl groups have an s-cis relationship, suffers from steric strain due to the bulky substituents. Instead, the molecule adopts the s-trans conformation, which relieves the strain but does not allow Diels–Alder reaction to take place.

 

14.33 The diene rotates to the s-cis conformation. The trans relationship of the two ester groups in the dienophile is preserved in the product.

image

 

14.34  

image

Both pentadienes are more stable in s-trans conformations. To undergo Diels–Alder reactions, however, they must rotate about the single bond between the double bonds to assume s-cis conformations.

image

When cis-1,3-pentadiene rotates to the s-cis conformation, steric interaction occurs between the methyl-group protons and a hydrogen on C1. Since it’s more difficult for cis– 1,3-pentadiene to assume the s-cis conformation, it is less reactive in the Diels–Alder reaction.

 

14.35 HC≡CC≡CH can’t be used as a Diels–Alder diene because it is linear. The end carbons are too far apart to be able to react with a dienophile in a cyclic transition state.

Furthermore, the product of Diels–Alder addition would be impossibly strained, with two sp-hybridized carbons in a six-membered ring.

 

14.36  

image

Two different orientations of the dienophile ester group are possible, and two different products can form.

 

14.37 The most reactive dienophiles contain electron-withdrawing groups.

Most reactive ————————> Least reactive

(NC)2C=C(CN)2       >          H2C=CHCHO          >         H2C=CHCH3         >       (CH3)2C=C(CH3)2

Four electron-                   One electron-                    One electron-               Four electron-

withdrawing groups       withdrawing group           donating group           donating groups

The methyl groups of 2,3-dimethyl-2-butene also decrease reactivity for steric reasons.

 

14.38 The difference in reactivity of the three cyclic dienes is due to steric factors. As the non-diene part of the molecule becomes larger, the carbon atoms at the end of the diene portion of the ring are forced farther apart. Overlap with the π system of the dienophile in the cyclic transition state is poorer, and reaction is slower.

 

14.39 Although an electron-withdrawing group increases the reactivity of a dienophile, it decreases the reactivity of a diene.

 

14.40 First, find the cyclohexene ring formed by the Diels–Alder reaction. After you locate the new bonds, you should then be able to identify the diene and the dienophile.

(a)

image

(b)

image

(c)

image

(d)

image

 

14.41  

image

 

Diene Polymers

14.42  

image

Ozone causes oxidative cleavage of the double bonds in rubber and breaks the polymer chain.

 

14.43 Polycyclopentadiene is the product of successive Diels–Alder additions of cyclopentadiene to a growing polymer chain. Strong heat causes depolymerization of the chain and reformation of cyclopentadiene monomer units.

image

 

UV Spectroscopy

14.44 Note: In general, the more conjugated double bonds that a molecule has, the longer the wavelength at which it will absorb UV light.

(a)

image

(b)

image

(c)

image

 

 

14.45 Only compounds having alternating multiple bonds show ππ* ultraviolet absorptions in the 200–400 nm range. Of the compounds shown, only pyridine (b) absorbs in this range.

 

14.46 To absorb in the 200–400 nm range, an alkene must be conjugated. Since the double bonds of allene aren’t conjugated, allene doesn’t absorb light in the UV region.

 

14.47 The value of λmax in the ultraviolet spectrum of dienes becomes larger with increasing alkyl substitution. Since energy is inversely related to λmax, the energy needed to produce ultraviolet absorption decreases with increasing substitution.

image

Each alkyl substituent causes an increase in λmax of 3–6 nm.

 

14.48  

image

In Problem 14.41, we concluded that one alkyl group increases λmax of a conjugated diene by approximately 5 nm. Since 2,3-dimethyl-1,3,5-hexatriene has two methyl substituents, its UV λmax should be about 10 nm longer than the λmax of 1,3,5-hexatriene.

 

14.49  

(a) β-Ocimene, C10H16, has three degrees of unsaturation. Catalytic hydrogenation yields a hydrocarbon of formula C10H22. β-Ocimene thus contains three double bonds and no rings.
(b) The ultraviolet absorption at 232 nm indicates that β-ocimene is conjugated.
(c) The carbon skeleton, as determined from hydrogenation, is:

image

Ozonolysis data are used to determine the location of the double bonds. The acetone fragment, which comes from carbon atoms 1 and 2 of 2,6-dimethyloctane, fixes the position of one double bond. Formaldehyde results from ozonolysis of a double bond at the other end of β-ocimene.

Placement of the other fragments to conform to the carbon skeleton yields the following structural formula for β-ocimene.

image

(d) image

 

General Problems

14.50  

(a)

image

(b)

image

(c)

image

 

14.51 All of the questions can be answered with the figure below.

image

(a) The three pi-bonds that are conjugated each have two p-atomic orbitals. Thus, there are six p-atomic orbitals in the conjugated system.

(b) The total number of molecular orbitals must equal the number of atomic orbitals used to create them. Therefore there are six molecular orbitals. (Ψ1, Ψ2, Ψ3, Ψ4, Ψ5, Ψ6)

(c) The bonding molecular orbitals are lower in energy than the p-atomic orbital energies. Thus, there are three bonding orbitals (Ψ1, Ψ2, Ψ3).

(d) The antibonding molecular orbitals are higher in energy than the p-atomic orbital energies. Thus, there are three antibonding orbitals (Ψ4*, Ψ5*, Ψ6*; antibonding orbitals are denoted with an *).

(e) Because orbitals fill from lowest energy to highest energy and there are six electrons in the bonding Ψ1, Ψ2 and Ψ3 molecular orbitals.

(f) The electron would move from the Ψ3 (the highest occupied molecular orbital; HOMO) to Ψ4* (the lowest unoccupied molecular orbital; LUMO).

 

14.52  

CH3CH2C≡CCH2H3

3-Hexyne

CH3CH=CHCH=CHCH3

2,4-Hexadiene

CH3CH2CH=C=CHCH3

2,3-Hexadiene

1H

NMR:

2 peaks (triplet, quartet) below 2.0 δ 3 peaks, two in region 4.5 – 6.5 δ 5 peaks, two in region 4.5 – 6.5 δ
13C

NMR

3 peaks,

8 – 55 δ (2)

65 – 85 δ (1)

3 peaks,

8 – 30 δ (1)

100 – 150 δ (2)

6 peaks,

8 – 55 δ (3)

100 – 150 δ (2)

~ 200 δ (1) (sp carbon)

UV absorption? no yes no

2,4-Hexadiene can easily be distinguished from the other two isomers because it is the only isomer that absorbs in the UV region. The other two isomers show significant differences in their 1H and 13C NMR spectra and can be identified by either technique.

 

14.53  

image

Conjugation with the                        Reaction with HCl yields                    Addition of Cl leads to the

oxygen lone pair electrons              a cation intermediate that                  observed product.

makes the double bond                   can be stabilized by the oxygen

more nucleophilic                              electrons.

There are two reasons why the other regioisomer is not formed: (1) Carbon 1 is less nucleophilic than carbon 2; (2) The cation intermediate that would result from protonation at carbon 1 can’t be stabilized by the oxygen electrons and does not form.

14.54  

image

 

14.55  

image

 

14.56 The first equivalent of maleic anhydride adds to the s-cis bond of the triene.

image

The new double bond has an s-cis relationship to the remaining double bond of the triene starting material. A second equivalent of maleic anhydride adds to the diene to form the product shown.

image

 

14.57 Much of what was proven for β-ocimene (Problem 14.49) is also true for myrcene, since both hydrocarbons have the same carbon skeleton and contain conjugated double bonds. The difference between the two isomers is in the placement of the double bonds.

The ozonolysis fragments from myrcene are 2-oxopentanedial (five carbon atoms), acetone (three carbon atoms), and two equivalents of formaldehyde (one carbon atom each). Putting these fragments together in a manner consistent with the data gives the following structural formula for myrcene:

image

 

14.58  

(a) Hydrocarbon A must have two double bonds and two rings, since the sole ozonolysis product contains all the carbons and a diketone–dialdehyde is formed.

image

(b) Rotation about the central single bond of II allows the double bond to assume the s– cis conformation necessary for a Diels–Alder reaction. Rotation is not possible for I.
(c) image

 

14.59  

image

 

14.60  

 

14.61  

image

 

14.62  

image

The stereochemistry of the product resulting from Diels–Alder reaction of the (2E,4Z) diene differs at the starred carbon from that of the (2E,4E) diene. Not only is the stereochemistry of the dienophile maintained during the Diels–Alder reaction, the stereochemistry of the diene is also maintained.

  • Although it is usually best to work backwards in a synthesis problem, it sometimes helps to work both forwards and backwards. In this problem, we know that the starting materials are a diene and a dienophile. This suggests that the synthesis involves a Diels– Alder reaction. The product is a dialdehyde in which the two aldehyde groups have a cis relationship, indicating that they are the products of ozonolysis of a double bond that is part of a ring. These two pieces of information allow us to propose the following synthesis:

image

The –CHO groups are cis to the ester in the product.

  • The lone pair electrons from nitrogen can overlap with the double bond π electrons in a manner similar to the overlap of the π electrons of two conjugated double bonds. This electron contribution from nitrogen makes an enamine double bond electron-rich.

image

The orbital picture of an enamine shows a 4 p-electron system that resembles the system of a conjugated diene.

  • Double bonds can be conjugated not only with other multiple bonds but also with the lone-pair electrons of atoms such as oxygen and nitrogen. p-Toluidine has the same number of double bonds as benzene, yet its λmax is 31 nm greater. The electron pair of the nitrogen atom can conjugate with the π electrons of the three double bonds of the ring, extending the π system and increasing λmax.

This file is copyright 2023, Rice University. All Rights Reserved.

License

Share This Book