Module 4 Lab Report
Module 4 Lab Report
Michael Wolz
Fall 2020
Abstract:
In the first reaction, 6.59 g (61.8 mmol, 67.7%) of 2-chloro-2-methylbutane was synthesized from 2-
methyl-2-butanol. The compound was identified and found to be pure by FTIR and 1H NMR
spectroscopy. After purification, 0.234 g (2.844 mmol, 15.0%) of cyclohexene was obtained from
cyclohexanol, and the same FTIR and 1H NMR analysis indicated it was also a pure product.
Introduction:
In this module, two reactions were performed. The first, an SN1 substitution seen in Scheme 1, was the
synthesis of 2-chloro-2-methylbutane from 2-methyl-2-butanol.
The second was an E1 elimination reaction seen in Scheme 2, the synthesis of cyclohexene from
cyclohexanol. In both cases, an acid was used to protonate the alcohol to form a leaving group. 1H NMR
and FTIR spectroscopy were used to analyze the products.
Scheme 2. Synthesis of cyclohexene from cyclohexanol.
For either type of reaction to proceed, the leaving group must be able to stabilize a negative charge well
enough that a carbocation intermediate will form. This formation is the rate limiting step, and due to the
presence of a carbocation intermediate in the mechanism these reactions may be subject to
Wolz 2
rearrangement if it is energetically favorable. Once the carbocation is formed, the geometry is trigonal
planar, which means the nucleophile can attack from either side. As such, any reaction with a
stereocenter will produce a racemic product.
An SN1 reaction, as shown in Figure 1, is a substitution reaction involving the spontaneous formation of a
carbocation by loss of a leaving group, and immediate nucleophilic attack on the carbocation to form a
new compound. Choosing a nucleophilic acid, like hydrochloric acid, aids in favoring the SN1 reaction.
An E1 reaction also involves the same formation of a carbocation, and the corresponding potential for
rearrangement and racemic product formation, but the reactant is not a nucleophile. Instead, the
reactant in this process acts as a base and attacks a beta proton, leaving a lone pair of electrons free to
form a pi bond to stabilize the carbocation. This process is depicted in Figure 2. The reactant is acting as
a base because it is taking a proton rather than attacking an electrophile. Using an acid whose conjugate
base is a weak nucleophile, such as sulfuric acid, helps to favor the E1 reaction.
These two reactions compete with each other, and the product is determined by the conditions of the
reaction. The acid used is important, as removing the strong nucleophiles heavily favors the E1 reaction,
but high temperatures can also be used to favor the E1 reaction. In this case, sulfuric acid is resonance
stabilized when deprotonated, and does not act as a nucleophile.
Wolz 3
The products of both reactions were massed and identified using FTIR and 1H NMR spectroscopy. The
SN1 reaction of 2-methyl-2-butanol with hydrochloric acid yielded 6.59 g (61.8 mmol, 67.7%) of 2-chloro-
2-methylbutane. At just over two thirds of the theoretical yield of 9.73 g, the reaction did not go
perfectly to completion, but it was within a reasonable threshold of success.
The FTIR spectrum for the SN1 reaction, Figure 1A, agrees that the reaction went as planned. It shows an
Sp3 C-H stretch between 2900 and 3000 cm-1 and an Sp3 C-Cl stretch at 798 cm-1, with the C-Cl stretch
being the major diagnostic band. The lack of an O-H stretch in the 3000s cm-1 indicates that there is not
enough alcohol left in the solution to register in the spectrum.
Figure 2A, the 1H NMR spectrum for the SN1 reaction, is also indicative of 2-chloro-2-methylbutane.
There is a triplet centered at 1.03 ppm, a singlet at 1.56 ppm, and a multiplet at 1.78 ppm, accounting
Wolz 4
for all three unique proton environments. There is no shift for the Cl atom because there are no protons
attached to it, or to its carbon. The absence of a fourth stretch indicates that there is not a significant
amount of alcohol left in the mixture, because it would be expected to produce a shift between one and
five ppm if it were present.
The E1 reaction of cyclohexanol with sulfuric acid yielded 0.228 g (2.774 mmol, 14.7%) of crude
cyclohexene. After drying, the yield of cyclohexene was 0.234 g (2.844 mmol, 15.0%). The
counterintuitive increase in yield after drying is likely due to contamination with the drying agent,
anhydrous magnesium sulfate. The low percent yield is partially associated with losses at each step of
the experiment, such as ‘hold up’ in the distillation or polymerization during reflux, but such a significant
loss means that the reaction probably did not go to completion, either. Without witnessing the reaction,
it is difficult to say with certainty.
The FTIR spectrum for the E1 reaction, Figure 3A, is indicative of cyclohexene. There is an R=C-H stretch
at about 3050 cm-1, the major diagnostic band, as well as a Sp3 C-H stretch just under 3000 cm-1, and a
C=C alkene stretch near 1650 cm-1. As with the first reaction, the absence of an O-H stretch above 3000
cm-1 indicates that the alcohol was consumed in the reaction.
Figure 4A, the 1H NMR spectrum for the E1 reaction, also indicates that the reaction was successful.
There is a multiplet centered at 1.61 ppm and a quartet centered at 1.99 ppm indicative of Sp3
hybridized carbons bound to those protons, and a triplet centered at 5.67 ppm which is indicative of an
alkene. There is no additional alcohol stretch present, suggesting that the product contains negligible
amounts of cyclohexanol.
Overall, the reactions were both successful but the SN1 reaction had a much higher yield than the E1
reaction. This is likely due to the competition between the E1 reaction and the hydration reaction, which
are competitive and dependent on acid concentration. The products were similar in purity, as neither
one demonstrated any spectroscopic characteristics of alcohol. They only varied in mass and percent
yield.
Such a low percent yield in the E1 reaction, without alcohol contamination, is possible. Cyclohexanol has
potential for hydrogen bonding between the alcohol groups, and as such would have a higher boiling
point than that of cyclohexene. The distillation process would efficiently separate the cyclohexene from
the cyclohexanol, even if only a small portion had reacted.
Conclusion:
While both reactions produced the desired product, the SN1 reaction was much more efficient in this
trial. Despite the use of heat and an acid whose conjugate base is a weak nucleophile, which should both
favor the E1 process, the second reaction did not produce even half of the theoretical amount of
product. Further examination of the remaining reactant could have indicated the amount of
cyclohexanol remaining in the reaction mixture. It is possible that using 12 M H2SO4 instead of 9 M
H2SO4, or a higher volume of acid, could push the reaction further to completion; because the
elimination reaction is favored under increasing acidity, following Le Chatlier’s Principle.
Wolz 5
Experimental:
All reactants were reagent grade and provided by the University of Colorado Denver Chemistry
Department. FTIR spectra were obtained with a Nicolet iS5 ATR FTIR spectrometer and all
measurements were in cm-1. The 1H NMR spectra were obtained using a Bruker 400 MHz spectrometer
using CDCl3 as the solvent. All measurements were in ppm and the following abbreviations were used: s
(singlet), d (doublet), t (triplet), q (quartet), and m (multiplet).
For the reaction forming 2-chloro-2-methylbutane, a 125 mL separatory funnel was charged with 10 mL
of 2-methyl-2-butanol and 25 mL of 12 M HCl, then mixed and vented for approximately 6 minutes. The
organic layer was washed twice with 10 mL of DI water, then twice with 5% sodium bicarbonate. The
organic layer was then dried with anhydrous magnesium sulfate and gravity filtered into a tared beaker
to be massed and run through FTIR and 1H NMR spectroscopy. 6.59 g (61.8 mmol, 67.7%) of 2-chloro-2-
methylbutane was recovered.
For the reaction forming cyclohexene, a 25 mL round bottom flask with a stir bar was charged with 2 mL
of cyclohexanol, then 1 mL of 9 M H2SO4. The flask was refluxed under medium heat for 15 minutes and
allowed to cool to room temperature. Simple distillation was used to extract the product from the
reaction mixture in the form of an azeotrope. The product was massed, then dried using anhydrous
magnesium sulfate and pipetted out of the drying agent and massed again. The product was run
through FTIR and 1H NMR spectroscopy.
2-chloro-2-methylbutane:
1
H NMR (ppm): 1.03 (t, 3H); 1.56 (s, 6H); 1.76 (q, 2H).
IR (cm-1): 2850-2950, 798.
Cyclohexene:
1
H NMR (ppm): 1.61 (m, 4H); 1.99 (q, 4H); 5.67 (t, 2H).
IR (cm-1): 3050, 2850-2950, 1650.
References:
Fishback, V., Schwartz, A., and Lopez, C., 2019. Organic Chemistry Lab Manual. Southlake,
TX: Fountainhead Press, LLC, p. 18, 22.
Wolz 6
Figure 2A. Annotated H+ NMR spectrum of 2-chloro-2-methylbutane from the SN1 reaction.
Wolz 7