03.07.2013 Views

Mosher Ester Analysis - Department of Chemistry and Physics

Mosher Ester Analysis - Department of Chemistry and Physics

Mosher Ester Analysis - Department of Chemistry and Physics

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>Mosher</strong> <strong>Analysis</strong> <strong>of</strong> Enantiomeric Purity <strong>and</strong> Assignment <strong>of</strong> Absolute Configuratoin<br />

by NMR spectroscopy.<br />

References:<br />

Patterson, J.; Sigurdsson, S.T. J. Chem. Educ. 2005, 82, 1049-1050.<br />

Sullivan, G. R.; Dale, J.A.; <strong>Mosher</strong>, H.S. J. Am. Chem. Soc. 1973, 95, 512-519.<br />

Dale, J.A.; <strong>Mosher</strong>, H.S. J. Org. Chem. 1973, 38, 2143-2147.<br />

Prelab: Prepare your reagent table <strong>and</strong> procedure sheet, including the reaction. The<br />

reaction is described below. Have them ready to do the experiment when you come into<br />

the lab. You can find all <strong>of</strong> the reagents at www.sigmaaldrich.com. Also get NMR<br />

spectra for the starting materials that you can use to compare with your crude product.<br />

Background<br />

While we <strong>of</strong>ten discuss chiral compounds as single enantiomers, it is common to<br />

encounter these compounds as mixtures <strong>of</strong> enantiomers. This occurs both in the<br />

laboratory <strong>and</strong> in nature. In the laboratory, this most commonly arises from the use <strong>of</strong><br />

less than enantiopure starting materials or reagents during synthesis or by racemization<br />

during reactions. In nature, this can be due to competing enantomeric biosynthetic<br />

pathways <strong>and</strong> by enzymatic or chemical racemization. Mixtures <strong>of</strong> enantiomers can be<br />

racemic (50:50 mixture <strong>of</strong> enantiomers) or scalemic (any other proportion).<br />

The analysis <strong>of</strong> enantiomeric mixtures can be divided into two objectives. The<br />

first is the analysis <strong>of</strong> enantiomeric excess <strong>and</strong> the second is the establishment <strong>of</strong> absolute<br />

configuration. Enantiomeric excess is the amount <strong>of</strong> one enantiomer over the other <strong>and</strong> is<br />

usually expressed as a percentage value. Percent e.e. is calculated as:<br />

% e.e. = (100%)|R – S|<br />

R + S<br />

Establishment <strong>of</strong> configuration is simply determining which enantiomer is R <strong>and</strong><br />

which is S, <strong>and</strong> determination <strong>of</strong> which is in excess. The latter is a difficult task as,<br />

enantiomers are identical in all physical <strong>and</strong> spectral properties. Enantiomers do differ in<br />

the sign <strong>of</strong> their optical rotation <strong>and</strong> certain compounds can be distinguished by<br />

spectropolarimetry (the study <strong>of</strong> variation <strong>of</strong> rotation with wavelength, <strong>of</strong>ten referred to<br />

as circular dichroism (CD) or optical rotatory dispersion (ORD)). However, this<br />

approach is <strong>of</strong>ten limited to certain classes <strong>of</strong> compounds with large changes in their<br />

rotation <strong>and</strong> it is not generally applicable.<br />

In many cases both assignment <strong>of</strong> configuration <strong>and</strong> determination <strong>of</strong> e.e. can be<br />

achieved by reacting the molecule in question with a chiral compound <strong>of</strong> known absolute<br />

configuration <strong>and</strong> enantiomeric purity (preferably a single enantiomer). This produces a<br />

mixture <strong>of</strong> diastereomers. The requirements for such a method are that the chemistry<br />

involved does not cause the chiral centers <strong>of</strong> either <strong>of</strong> the reactants to racemize <strong>and</strong> that<br />

the reaction is quantitative in reaction with both enantiomers <strong>of</strong> the compound you are<br />

analyzing. That is, in some cases, one enantiomer reacts faster than the other. This is


called kinetic resolution. Usually this is avoided by using an excess <strong>of</strong> the enantiopure<br />

reactant <strong>and</strong> using reactions that are fast <strong>and</strong> irreversible.<br />

In this lab, you will prepare a chiral ester <strong>of</strong> a chiral secondary alcohol, 1phenylethanol.<br />

The ester is derived from the chiral carboxlic acid -methoxy-trifluoromethylphenylacetic<br />

acid (MTPA, <strong>Mosher</strong>’s acid), shown below.<br />

F3C<br />

MTPA<br />

Diastereomers, unlike enantiomers, have distinguishable physical <strong>and</strong> spectral<br />

properties because they are not mirror images. Diastereomers have two (or more) chiral<br />

centers. The added chiral center(s) can be the same or opposite configuration as that <strong>of</strong><br />

the original molecule. The matched <strong>and</strong> mismatched diastereomers will have different<br />

spectra <strong>and</strong> can thus be distinguished as shown below for one <strong>of</strong> the enantiomers <strong>of</strong><br />

MTPA <strong>and</strong> scalemic 1-phenylethanol.<br />

F3C<br />

O<br />

OCH 3<br />

OH<br />

+<br />

OH<br />

CH 3<br />

single enantiomer mixture <strong>of</strong> enantiomers<br />

mixture <strong>of</strong> diastereomers<br />

MTPA is a useful chiral acid for analyzing enantiomeric excess for several<br />

reasons. First, both enantiomers are available in high enantiomeric purity. Second, the<br />

chiral center is adjacent to the carboxyl bringing it in close proximity to the chiral center<br />

<strong>of</strong> the alcohol when the ester is prepared. Also, the chiral center is not able to be<br />

racemized as it is quaternary <strong>and</strong> thus contains no acidic protons. Finally, the presence <strong>of</strong><br />

the trifluoromethyl group allows for clean analysis <strong>of</strong> enantiomeric excess by 19 F NMR,<br />

which is uncomplicated by signals from the chiral alcohol, unlike the 1 H NMR spectrum.<br />

However, 1 H NMR can also be used if the signals are sufficiently resolved. This also<br />

allows for assignment <strong>of</strong> configuration based on a preferred conformation <strong>of</strong> the ester. In<br />

this experiment, you will team up with a partner. You will be given a sample <strong>of</strong> 1phenylethanol<br />

<strong>of</strong> unknown enantiomeric composition. One <strong>of</strong> you will prepare the (R)-<br />

(+)-MTPA ester <strong>of</strong> your 1-phenylethanol sample <strong>and</strong> the other will prepare the (S)-(-)-<br />

MTPA ester <strong>of</strong> the sample. You will then compare your 19 F NMR <strong>and</strong> 1 H NMR data <strong>and</strong><br />

determine both the enantiomeric excess <strong>and</strong> the major enantiomer present according to<br />

the published methods by <strong>Mosher</strong>.<br />

We will prepare our esters using N,N’-dicyclohexylcarbodiimide (DCC) to<br />

activate the acid toward substitution <strong>and</strong> 4-dimethylaminopyridine (DMAP) as a catalyst.<br />

*<br />

O<br />

OCH 3<br />

OH<br />

F3C<br />

F3C<br />

O<br />

OCH 3<br />

O<br />

+<br />

OCH 3<br />

H<br />

O<br />

H<br />

O<br />

CH 3<br />

CH 3


This is an exceptionally mild method <strong>of</strong> carboxyl activation <strong>and</strong> is used extensively for<br />

the synthesis <strong>of</strong> both esters an amides, particularly for peptide synthesis.<br />

O<br />

R OH<br />

R'OH, DCC, DMAP<br />

CH 2Cl 2<br />

O<br />

R OR'<br />

The reaction proceeds in two steps. First, the the acid reacts with DCC, forming<br />

an O-acylisourea. This species is similar to an anhydride <strong>and</strong> the carbonyl is highly<br />

susceptible to substitution. In the second step, the alcohol attacks the carbonyl to produce<br />

the ester <strong>and</strong> eliminates dicyclohexylurea as shown below.<br />

Step 1.<br />

O<br />

R O<br />

Step 2.<br />

O<br />

N<br />

c-Hx<br />

c-Hx<br />

R O N<br />

H<br />

O-acylisourea<br />

(similar to an anhydride)<br />

O<br />

H<br />

+<br />

N C N<br />

Dicyclohexylcarbodiimide<br />

(DCC)<br />

O<br />

HN<br />

c-Hx<br />

+<br />

c-Hx<br />

R OR'<br />

O N<br />

H<br />

Dicyclohexylurea (DCU)<br />

N<br />

c-Hx<br />

O<br />

R O<br />

O<br />

R O<br />

R' O H +<br />

c-Hx<br />

R' O +<br />

R O N<br />

H<br />

+<br />

+<br />

O<br />

H<br />

R O<br />

OR'<br />

H<br />

N C N<br />

H<br />

N C N<br />

H-bond<br />

O<br />

R O<br />

While this process can proceed by itself or with added bases such as<br />

triethylamine, it is somewhat sluggish. In order to speed the process, a nucleophilic<br />

N<br />

c-Hx<br />

N<br />

H<br />

H<br />

N<br />

c-Hx<br />

c-Hx<br />

N<br />

H<br />

c-Hx


catalyst, 4-dimethylaminopyridine (DMAP) is used. DMAP facilitates the process both<br />

by acting as a base <strong>and</strong> by forming activated acylpyridinium derivatives (below) from<br />

DCC <strong>and</strong> from the O-acylisourea. These are much more reactive as the pyridine is an<br />

excellent leaving group. This accelerates the reaction severalfold over the uncatalyzed<br />

reaction.<br />

H 3C<br />

N<br />

CH 3<br />

N<br />

N<br />

c-Hx<br />

N<br />

H<br />

c-Hx<br />

Procedure<br />

Wear gloves. DCC <strong>and</strong> DMAP are very toxic.<br />

You will be provided with a culture tube containing a scalemic solution <strong>of</strong> 6.0 L<br />

<strong>of</strong> 1-phenylethanol in 1 mL <strong>of</strong> dry CH2Cl2. To this add a stir bar <strong>and</strong> 13 mg <strong>of</strong> (R)-(+)- or<br />

(S)-(-)-MTPA as directed by your instructor. Be sure to record which enantiomer <strong>of</strong><br />

MTPA you use. Also be sure to cap the tube quickly to minimize any moisture entry<br />

after each addition <strong>of</strong> reagent. A nitrogen balloon <strong>and</strong> septum can be used if you wish.<br />

Cool the flask in an ice bath <strong>and</strong> add 12 mg <strong>of</strong> DCC <strong>and</strong> a spatula point (~1-2 mg) <strong>of</strong><br />

DMAP in sequence. Add the DCC to the reaction as soon as it is weighed as it tends to<br />

pick up moisture in the air. Stir the solution for 5 minutes in the ice bath, <strong>and</strong> then 25<br />

minutes at room temperature, during which time dicyclohexylurea will precipitate.<br />

While you are waiting on the reaction to complete, experiment with different TLC<br />

systems to determine if your reaction has finished. Spot 1-phenylethanol, DCC, DMAP,<br />

MTPA, <strong>and</strong> your reaction separately on the TLC plate <strong>and</strong> develop. Visualize the plate<br />

with UV light <strong>and</strong> circle the spots with a pencil. Begin with 1:1 hexane-ethyl acetate <strong>and</strong><br />

evaluate several mixtures until you find one that separates all <strong>of</strong> the components well,<br />

especially the product. This mixture will be used for preparative TLC to purify the<br />

product.<br />

Also while you are waiting on the reaction to complete, prepare two large test<br />

tubes for exraction <strong>of</strong> the reaction mixture. To one tube add 2 mL <strong>of</strong> 5% NaHCO3 <strong>and</strong> to<br />

the second add 2 mL <strong>of</strong> saturated NaHCO3.<br />

When the reaction is judged to be complete or 30 minutes passed, add 2 drops <strong>of</strong><br />

water <strong>and</strong> stir an additional 10 minutes. Then, add one mL <strong>of</strong> CH2Cl2 <strong>and</strong> 2 mL <strong>of</strong> 5%<br />

acetic acid to the culture tube <strong>and</strong> swirl vigorously. Using a pipet, transfer the organic<br />

layer to the test tube containing 5% NaHCO3 <strong>and</strong> swirl as before. Repeat this for the tube<br />

containing saturated NaHCO3. Prepare a pipet drying column plugged with glass wool<br />

<strong>and</strong> filled with ~3 cm <strong>of</strong> anhydrous Na2SO4. Clamp the tube in place <strong>and</strong> place a tared 10<br />

mL round bottom flask below to collect the eluent. Add the organic phase from the last<br />

tube <strong>and</strong> allow it to pass through the column <strong>and</strong> collect in the flask. Once all <strong>of</strong> the<br />

organic phase has been added <strong>and</strong> the top <strong>of</strong> the liquid reaches the Na2SO4, wash the<br />

column twice with 2 mL CH2Cl2. Remove the solvent by rotary evaporation <strong>and</strong> remove<br />

any traces under high vacuum for no more than a minute or two (The instructor will show<br />

R<br />

O<br />

N<br />

N<br />

CH 3<br />

CH 3


you how to do this). Obtain a crude yield, dissolve the product in CDCl3 <strong>and</strong> obtain a 1 H<br />

NMR spectrum <strong>of</strong> the crude ester.<br />

Preparative Thin-Layer Chromatography<br />

The <strong>Mosher</strong> ester will be purified by preparative thin-layer chromatography. This<br />

is essentially the same as what you use for monitoring your reactions, only on larger<br />

scale. We will use a large developing tank, <strong>and</strong> large, thickly coated plates. The plates<br />

we are using are capable <strong>of</strong> separating around 10-20 mg, depending on separation<br />

efficiency. Unlike flash chromatography, no specific Rf value is required. Simply<br />

identify a solvent system which will adequately separate your compounds <strong>and</strong> develop<br />

the plate. The technique does have limitations, however. The large surface area <strong>of</strong> the<br />

silica, combined with its mildly acidic nature tend to promote air oxidation <strong>of</strong> sensitive<br />

functionalities such as aldehydes, amines, <strong>and</strong> reactive double bonds. Also, the method is<br />

rather expensive to perform on other than small scale separations. Nonetheless, it is <strong>of</strong><br />

great utility in difficult separations <strong>of</strong> small amounts <strong>of</strong> material, such as extracts <strong>of</strong><br />

natural products <strong>and</strong> separation <strong>of</strong> diastereomers. In the second part <strong>of</strong> this experiment,<br />

you will separate the products <strong>of</strong> your <strong>Mosher</strong> ester reaction by preparative TLC <strong>and</strong><br />

characterize your product by 1 H <strong>and</strong> 19 FNMR.<br />

Preparative TLC Procedures<br />

1. Set up TLC Developing tank.<br />

a. Examine TLC’s in various solvent systems (TLC hood) if you have not yet<br />

determined an optimal composition.<br />

b. Dissolve your sample in 1.0 mL CH2Cl2 or use the NMR sample directly.<br />

c. Perform a TLC on your material using that system. (Don’t forget to cospot<br />

with your starting material).<br />

d. Make up 100 mL <strong>of</strong> your chosen solvent system <strong>and</strong> put it in the tank.<br />

2. Prepare your TLC plate.<br />

a. Set up the solvent streaker (Have the instructor help you.)<br />

b. Draw a pencil line about 1” from the bottom <strong>of</strong> the plate.<br />

c. Set the streaker up so that the tip <strong>of</strong> the capillary traverses this line as you<br />

draw it along the benchtop.<br />

d. Fill the streaker with your solution. The reservoir will only hold around<br />

0.25-0.5 mL.<br />

e. Streak the plate on the line you drew. Try to keep the thickness <strong>of</strong> your<br />

b<strong>and</strong> between 1-3mm. Wave your h<strong>and</strong> over the plate to assist drying.<br />

f. Refill the reservoir again <strong>and</strong> repeat.<br />

g. Clean the reservoir with acetone. Dry it under a nitrogen flow. Give it to<br />

the next person to use. Be gentle with the streaker. We only have one <strong>and</strong><br />

if broken, it will end our experiment.<br />

3. Develop your plate.<br />

a. Place your plate in the developing tank.<br />

b. Develop the plate until the solvent front reaches within one inch <strong>of</strong> the top.<br />

This will take about an hour. Be sure to check the plate every 15 minutes<br />

to see its progress.


c. Remove the plate from the tank <strong>and</strong> mark the solvent front with a pencil.<br />

d. Wait for the plate to dry. You can assist this by fanning the plate with a<br />

folder or using a flow <strong>of</strong> dry nitrogen. While you are waiting, you should<br />

be cleaning glassware or something else productive. Do not wait an<br />

excessive amount <strong>of</strong> time. Many compounds are sensitive to oxidation on<br />

a TLC plate as the surface area is very large.<br />

4. Isolate the product.<br />

a. When your plate is dry enough, visualize the b<strong>and</strong>s using UV light. Mark<br />

them lightly with a pencil.<br />

b. Using an exacto knife or the flat edge <strong>of</strong> a spatula, scrape the b<strong>and</strong>s <strong>of</strong>f<br />

onto a lengthwise folded piece <strong>of</strong> clean, white paper.<br />

c. Prepare a vacuum filtration using a fritted glass funnel as shown below.<br />

d. Place the scrapings into a fritted glass funnel <strong>and</strong> press them into a pad.<br />

e. Wash the compound <strong>of</strong>f the silica into a round bottomed flask using 3 x 5<br />

mL <strong>of</strong> TLC solvent.<br />

f. Remove the solvent by rotary evaporation <strong>and</strong> determine the mass <strong>of</strong> your<br />

product. While this is occurring have your sample analyzed by chiral GC<br />

or be cleaning your glassware.<br />

5. Characterize your product by 1 H <strong>and</strong> 19 FNMR.<br />

Since one member <strong>of</strong> your team will prepare the R-MTPA ester <strong>and</strong> the other will<br />

prepare the S-MTPA ester, compare your results <strong>and</strong> assign the stereochemistry <strong>of</strong> the<br />

alcohol <strong>and</strong> the % e.e. according to the two <strong>Mosher</strong> references.<br />

Questions:<br />

1. Compare the % e.e. based on the integration <strong>of</strong> the 1 H <strong>and</strong> the 19 F spectra <strong>of</strong> the<br />

<strong>Mosher</strong> ester you made. For 1 H, calculate for all resonances that are sufficiently<br />

resolved to integrate reliably. Do they agree? If not, suggest a reason for the<br />

discrepancy.<br />

2. Calculate the % e.e. from your crude product <strong>and</strong> compare with that from the<br />

purified sample. Do they agree? If not, suggest a reason for the discrepancy.<br />

3. Compare the % e.e. you obtained to that obtained by your partner. Do they agree?<br />

If not, suggest a reason for the discrepancy.<br />

4. Suggest two alternate methods for determining the % e.e. <strong>of</strong> your alcohol.


Plate<br />

20 cm x 20 cm<br />

2.5 mm layer<br />

thickness<br />

Bench Guide/<br />

Reservoir Holder<br />

Streak line<br />

~1" from bottom<br />

1-3mm thickness<br />

Glass Plate Cover<br />

Developing chamber<br />

Solvent<br />

~3/4" from bottom<br />

About 100 mL<br />

TLC Plate Streaker<br />

Bench<br />

Preparative TLC Plate<br />

Development<br />

Reservoir<br />

Capillary<br />

TLC Plate<br />

to aspirator<br />

25 mL<br />

Claisen<br />

head

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!