0% found this document useful (0 votes)
14 views

Complete Labmanual 2021 Final

This document provides safety guidelines for students conducting experiments in an organic chemistry laboratory. It outlines proper personal protective equipment including wearing eye protection, closed-toe shoes, and protective clothing. Hazards such as open flames, food and drink, and chemical exposure are addressed. Procedures for injuries and emergencies are also summarized. The document aims to establish a safe laboratory environment for students to learn techniques of organic chemistry.

Uploaded by

kenidi green
Copyright
© © All Rights Reserved
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
14 views

Complete Labmanual 2021 Final

This document provides safety guidelines for students conducting experiments in an organic chemistry laboratory. It outlines proper personal protective equipment including wearing eye protection, closed-toe shoes, and protective clothing. Hazards such as open flames, food and drink, and chemical exposure are addressed. Procedures for injuries and emergencies are also summarized. The document aims to establish a safe laboratory environment for students to learn techniques of organic chemistry.

Uploaded by

kenidi green
Copyright
© © All Rights Reserved
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 138

LABORATORY MANUAL

for
Techniques of Organic Chemistry

Chemistry 2633

Department of Chemistry and Biochemistry


University of Missouri-St. Louis
St. Louis, MO 63121

Prepared by

James S. Chickos
Valerian T. D’Souza
David L. Garin

Spring 2019

1
Table of Contents

Safety .............................................................................................................................. 3
Glassware and Equipment ...................................................................................................................... 6
Other Equipment in the Laboratory ........................................................................................................ 9
Advanced Preparation and the Notebook ............................................................................................. 10
Notebook Calculations ......................................................................................................................... 11
Labels ............................................................................................................................ 13
Recrystallization or crystallization ....................................................................................................... 14
Decolorization: ............................................................................................................................ 15
Filtering ............................................................................................................................ 15
Solvent Polarity in decreasing order: ................................................................................................... 17
Melting Point (mp) Determination: ...................................................................................................... 17
Extraction ............................................................................................................................ 19
Drying Agents ............................................................................................................................ 20
Extraction Solvents ............................................................................................................................ 21
EXPERIMENT 1.  SYNTHESIS OF ASPIRIN ................................................................................ 22
EXPERIMENT 2.  SYNTHESIS OF ACETAMINOPHEN ............................................................. 25
EXPERIMENT 3.  CAFFEINE EXPERIMENT ............................................................................... 27
EXPERIMENT 4.  THIN LAYER CHROMATOGRAPHY (TLC) ................................................. 28
EXPERIMENT 5.  SYNTHESIS OF ISOPENTYL ACETATE (BANANA OIL)........................... 35
EXPERIMENT 6.  FRACTIONAL AND SIMPLE DISTILLATION OF A BINARY MIXTURE 48
EXPERIMENT 7. ISOLATION OF A NATURAL PRODUCT BY STEAM DISTILLATION .... 50
EXPERIMENT 8. PREPARATION OF METHYL SALICYLATE - VACUUM DISTILLATION
52
EXPERIMENT 9.  IDENTIFICATION OF UNKNOWNS .............................................................. 54
EXPERIMENT 10. MULTISTEP SYNTHESIS ............................................................................... 63
INFRARED SPECTROSCOPY........................................................................................................... 68
Analysis of IR Spectra .......................................................................................................................... 77
APPENDIX 1. Table of IR Spectra ...................................................................................................... 98
APPENDIX 2. Tables of Selected Unknowns and Derivatives ......................................................... 126

2
Safety
The teaching and practice of organic chemistry has undergone major changes in
the past few years with regards to the safe handling and disposal of chemicals and this is
likely to continue. Some chemicals such as benzene, commonly used as a solvent a few
years ago, have practically disappeared from the laboratory. The use of other chemicals
such as carbon tetrachloride are slowly being phased out. Nevertheless the use of many
substances of variable toxicity will continue both in industry and academia and it is the
obligation and responsibility of a laboratory course in organic chemistry to instruct the
student in the safe use and disposal of these substances.

The following safety rules constitute a beginning in establishing a safe laboratory


environment:

a. Eye protection must be worn at all times the laboratory is in session. This is a
regulation of the State of Missouri. The wearing of approved eye protection can be met by
several different types of glasses or goggles. All prescription safety glasses must meet
current ANSI standards and have side shields. Regular prescription glasses are not
adequate protection. Most plastic lenses meet the specifications of safety glass and are
acceptable. Students wearing plastic lenses should be aware that the lenses can be attacked
by some of the solvents used in the laboratory and some loss of the optical properties can
occur if the two come in contact. In case your prescription glasses do not meet current
ANSI standards, wear Safety glasses or goggles over your prescription glasses.

Students that do not normally wear glasses are encouraged to purchase a pair of
goggles or non-correcting glasses. Although goggles offer the best eye protection, they
also have a tendency to fog up thereby reducing vision. Goggles are intended to protect
the eyes, no the forehead! Be sure to purchase a pair that affords plenty of ventilation.

The wearing of contact lenses in the laboratory is strongly discouraged. Material


that comes in contact with the eye as a result of a splash is considerably more difficult to
remove if the person is wearing contact lenses. Photochromic glasses are acceptable but
sunglasses are not unless there are extenuating medical circumstances that require such
glasses.

Everyone should be aware of the location of the nearest eyewash; getting to an


eyewash quickly can help to save your sight.

b. Clothing that is worn offers important protection. In a chemical spill, if the


chemical is a liquid, it will be absorbed by the clothing and while the clothing should be
removed as quickly as possible, contact with the skin is kept to a minimum. The wearing
of old jeans is recommended, but clothing covering as much exposed surface as
reasonable should be worn. The wearing of shorts or sandals is not allowed in the
laboratory. No shoes, no shirt, no laboratory.

3
In addition to dressing properly for the laboratory, be sure to take advantage of the
protection that the pair of "disposable" gloves provided in the laboratory. Disposable
gloves are not suitable for handling aggressive or highly hazardous chemicals and should
never be re-used. Whenever a disposable glove comes in contact with hazardous
chemicals it should be removed followed by thorough hand washing and new gloves for
continued work. Nitrile gloves are generally considered more chemically resistant than
latex gloves (some people do have Latex glove allergies, consider nitrile gloves if
symptoms occur), but the safety data sheet (SDS) of the compounds you are handling or a
glove selection guide should be consulted.

c. No food or beverages are allowed in the laboratory. Be sure to wash your


hands when you leave the laboratory, particularly if you are going out for snack.

d. Open flames are not allowed in the laboratory. Fires have been one of the major
causes of accidents in the laboratory. For those experiments that require an open flame
(the preparation of glass capillaries used in thin layer chromatography), a special hood in
the laboratory will be designated for use. Hot plates, electric heating mantles and steam
baths will be used as sources of heat instead. While the use of other sources of heat has
eliminated the cause of most fires, the risk of fire has not been totally eliminated. Many
of the organic solvents in use today are quite volatile, quite flammable and can be ignited
given the right conditions. Heating solvents should always be conducted in a well-
ventilated area such as a hood. You should also be aware of the properties of the common
solvents you are using, particularly their flash points. Although not commonly used,
carbon disulfide will spontaneously ignite on a steam bath if the vapors are heated above
about 80 C. While the risk of fires has been reduced, you should be aware of the location
of the nearest fire extinguisher. Most laboratories are also equipped with a fire blanket and
safety shower. When you are in the laboratory, take a minute to read how to operate these
devices. However, only the instructor is designated to use the fire blanket and
extinguisher.

e. If, during the course of a laboratory, you are injured, no matter how minor,
you should bring this to the attention of the instructor or teaching assistant. If the
injury does require some medical attention, the instructor or teaching assistant will escort
you to the Campus Health Center. In case of a serious injury, you will be escorted to the
appropriate medical facility. Students should be aware that in the event that either the
teaching assistant and/or instructor must leave the laboratory, and the other is unavailable,
you will be asked to stop work until they return. In the event of emergencies, call 911
from any campus phone.

In case you spill a chemical on your skin, notify the instructor immediately and a)
Assist exposed person with use of safety shower and/or eye wash; b) Flush body and/or
eyes with copious amounts of water; c) Remove any contaminated clothing while under
the safety shower; d) Wash skin with mild soap and water, do not use neutralizing agents,
creams, lotions or salve; e) Seek medical attention.

4
A discussion of the different types of accidents that can happen in an organic
laboratory is beyond the scope of this manual. However, it may be useful to describe the
most common accidents in the laboratory that usually requires some sort of medical
attention. The most frequent type of accident is usually associated with inserting or
removing a thermometer or glass tubing in or from a rubber stopper. The need for rubber
stoppers and glass tubing has been greatly reduced with the use of ground glass joints.
Please heed the following advice if you need to insert or remove a glass rod into a rubber
stopper. Always use a lubricant such as stopcock grease to lubricate the shaft. Do not
force or apply excess pressure to dislodge or insert a stopper and use a towel as an
intermediary between you and the glass. Instead of pushing the glass rod or thermometer
in or out, try using a pulling, twisting motion. Don't hesitate to ask for assistance. In trying
to insert or remove your Teflon thermometer adapter from your thermometer, use the
steam in the hood to heat and lubricate. Once heated, the Teflon will expand and slide
much easier down the shaft of the thermometer.

In case you get injured, you to fill out an incident report. Instructions and forms
can be found in the links below.
www.umsystem.edu/ums/rules/bpm/bpm700/manual_704
www.umsystem.edu/ums/fa/management/records/forms/risk

f. Chemicals are generally not to be disposed of by flushing down the sink or


simply tossing in the trash. Follow the suggested disposal directions and ask your
instructor how to properly dispose of your wastes. Solid wastes will either be collected in
an appropriate container. Please note that the laboratory is equipped with special waste
containers for glass; do not place glass contaminated with hazardous chemicals into the
broken glass boxes. Liquids are generally collected in containers kept in a corner of the
laboratory. Acetone is usually used to wash and dry glassware in the laboratory, collect
your wash acetone in a beaker and dispose of it in the container specifically designated for
wash acetone. Acetone vapors do pose a fire risk.

g. Take only as much of the reagents as you need. In the event you have any
excess reagent left over, ask your instructor how to dispose of it properly. Accidental
contamination of reagents is not only unfair to your colleagues, it can be down-right
dangerous. Potassium permanganate in contact with glycerol and other organic materials
can spontaneously ignite.

h. Ingestion of chemicals can occur by several obvious and some less obvious
mechanisms. The tasting of laboratory chemicals, once a common practice, is no longer
an acceptable practice. Many organic solvents are quite volatile and a common exposure
to these substances is through the lungs. The vapors of some common solvents such as the
simple halogenated methanes, can cause liver damage and some are suspected of being
carcinogenic. Avoid breathing the vapors of these materials and always work with these
materials in a well-ventilated area such as a hood. Many organic solvents are also rapidly
absorbed through the skin. It is a good idea to use the disposable gloves when handling
these materials. Organic ethers pose fire hazards. The vapors of diethyl ether can
accumulate in troughs or on the floor and can be ignited by remote sparks or flames.

5
Additionally, the more highly substituted ethers react readily with oxygen to form
potentially explosive peroxides. Avoid distilling ethers to dryness.

i. Proper laboratory practices, neatness and organization in the laboratory are


considered closely associated with safety. A neat and well organized student is usually
less likely to have an accident and more likely to get a better grade. Come to laboratory
prepared. Be sure to read the experiment before coming to lab and run through how you
will complete the experiment in your mind before you try it with your hands. Look up the
properties of the materials you will be working with and be aware of any precautions
associated with any of these substances. Ask questions if any part of the experiment is
unclear to you.

Glassware and Equipment


The equipment in the laboratory at your work station is housed in three cabinets. Two of
the cabinets can be locked and contain a variety of glassware. The third cabinet contains a
variety of hardware that is relatively non-destructible. The two lockable cabinets contain
mostly glassware and are identified as a personal locker and as a community locker. You
will be assigned a key to your personal locker and with the exception of the stockroom
personnel, no one else will have access to its contents. You will be able to store your
personal belongings in this locker. To assist you and your colleagues working at your lab
station (as well as the teaching staff), we will ask you to attach your name on a label at
your work station. This will be provided to you during check-in.

Before using glassware, evaluate it for cracks, stars and chips (in which case the
glassware should not be used, because it can break). Use a brush and a dust pan and heavy
duty gloves to remove broken glass; dispose it in the special glass containers. All incidents
involving broken glassware must be reported to the instructor.

The community locker that you will be using will have limited access by a
maximum of three other students. This locker contains most of the glassware with the
ground glass joints. Since this glassware is expensive, and will be used by others, we
have devised the following system to insure that if any glassware is broken or lost, the
student responsible for the loss is held accountable. During the first five minutes of the
laboratory, this locker will be opened and available for inspection. At this time you will
also be asked to sign the log-in for the community locker. If you find the contents
incomplete or equipment broken, you are to bring this to the attention of the instructor.
Your instructor will replace the damaged or missing item and will instruct the storeroom
to charge the replacement costs to the individual whose name appears as the previous last
entry in the log-in. In order to insure that this process works properly, it will be necessary
that you familiarize yourself with the contents of this locker. The following list itemizes
each piece but you should become familiar with the shape and function of each. The price
of each item is indicated on the check-in sheet.

Community Cabinet Equipment

6
6 24/40 single neck round bottom flasks, assorted sizes
2 24/40 condensers
1 24/40 125 mL addition funnel with pressure equalizer with stopper
1 pear shaped separatory funnel (250 mL) and stopper
1 24/40 distilling head with a thermometer adapter
1 24/40 Claisen distilling head
1. 24/40 distillation connector with vacuum adapter
1 24/40 250 mL three neck round bottom flask
1 24/40 stopper
1 24/40 drying tube
1 24/40 thermometer adapter
3 24/40 plastic clips

The following equipment is included in your personal locker:

Personal Locker

Item Use

beakers, assorted sizes containers


Erlenmeyer flasks, assorted sizes containers for crystallization
vacuum filter flasks, assorted sizes containers for vacuum filtrations
graduated cylinders, containers for measuring volume of liquids
wide mouth bottles handing in samples
narrow stem funnels gravity filtration of liquids
powder funnel adding solids to flasks, gravity
filtering hot solutions
porcelain funnels vacuum filtration of cold
solutions containing solids
drying tube preventing moisture backup
stainless steel spatula handling small amounts of solids
stainless steel scoupula with handle handling larger amounts of solids
watch glasses for spot tests, evaporating small
amounts of liquids, covers
test tubes for tests
test tube with side arm used as a cold finger in sublimation
Teflon thermometer adapter for use in attachment of
thermometer to distilling head
–10 to 260 °C thermometer measurement of temperature
poylethylene wash bottle acetone wash bottle
filter vac for attaching porelein funnels to
filter flasks
stirring rods for stirring solutions and
scratching glass vessels during
crystallization

7
The following hardware will be found in your second community cabinet

Hardware in Community Cabinet

6 two and three prong clamps, assorted sizes


6 clamp holders
2 pinch clamps
2 screw clamps
12’ flexible rubber tubing, assorted lengths
6’ vacuum tubing, assorted lengths
3 cork rings, assorted sizes
1 test tube rack, 6 tube capacity
1 test tube rack, 12 tube capacity
1 test tube clamp
1 wire gauze
1 crucible tongs
3 iron rings, assorted sizes

Not included in your lockers but considered consumables and available in kit form
(or individually) at cost are the following:

Consumable Items (available as needed, subject to change)

Item Use

disposable pipettes for dispensing small volumes of liquids


pipette bulbs for dispensing small volumes of liquids
towel for cleaning and drying
sponge for cleaning
test tube brush for cleaning
glass slides for Fischer Johns melting point
apparatus
disposable gloves for handling toxic and corrosive
substances
vials, assorted sizes for samples

8
Other Equipment in the Laboratory
You will also find a Variac or variable voltage supply unit at each of your work stations.
These Variacs are usually attached to the racks by your work station and are to be used
whenever you wish to control the power delivered to an apparatus, usually a heating
mantle. Your instructor will demonstrate their use. Please note that heating mantles are
never to be attached directly to the 115 V receptacles.

You will also find other equipment in the laboratory that you will use from time to
time. This includes rotary evaporators that can be used to evaporate and recover solvents
quickly. These units operate on the principles of vacuum distillation which will be
discussed later in the semester. You will also find a vacuum pump and appropriate
condensing flasks on a movable cart that you may use later.

A variety of top loading electronic balances are available for use in the laboratory.
Although these are easy to use, if you are unfamiliar with them, take a minute to ask for
instructions. These few minutes could save you hours in having to repeat an experiment
later. In addition to the top loading balances which will weigh to 0.1, 0.01 and 0.001g
respectively, an analytical balance is also available that will weigh to 0.0001 g. This
balance should be used only when you require this accuracy. Generally for most
experiments, use of the top loading balances will suffice. Maintain the area around the
balance well; clean residual chemicals from the balances after each use. This will reduce
the risk of accidental exposures.

The laboratory does not contain a source of distilled water nor is distilled water
required unless you are specifically requested to use it. Distilled water is available in the
Physical Chemistry Laboratory and will be made available if necessary.

A variety of different sources of heat are available in the laboratory including


steam. Only the steam baths in the hoods are to be used. Hot plates and heating mantles
are available in various cabinets in the laboratory and will be made available when
necessary.

In addition to the equipment housed in the organic chemistry laboratory, additional


equipment that will be used during the semester can be found in the organic instrument
room that is directly attached to the laboratory. This room houses the Perkin Elmer
Fourier Transform Infrared Spectrometer and a variety of accessories that are used to
prepare samples for infrared analysis. A variety of computers are also available. The
spectrometer and the computers are for use by students enrolled in the organic laboratory
and are available for use during laboratory time. For use at other times, you may need to
make arrangements with your instructor.

9
Advanced Preparation and the Notebook
Before coming to the laboratory you will find that some advanced preparation can save
significant time and effort in the laboratory. The advanced preparation will vary
depending on the nature of the experiment, but some general rules will be applicable to all
experiments. Read the experiment first and identify all of the reagents and substances that
will be used and prepared during the course of the experiment. Next, collect the physical
properties of these materials before coming to the laboratory. Physical properties that will
be useful include melting point and boiling points (if available) of all the reagents and
expected products, their densities (if liquid) and some general solubility properties. Most
importantly, review the chemical safety data sheets (SDS) for each chemical used for
each lab and incorporate safety precautions into your lab procedures. Safety data sheets
provide procedures for handling or working with substances in a safe manner and can be
found online, for example at www.sigmaaldrich.com. Molecular weights should be
calculated for all the reagents. This information should be recorded in your notebook
which will be described below. While this may seem like a lot of work, many of these
same reagents will be used repeatedly. Once recorded in your notebook, this material can
be copied or a reference made to the page on which this information is recorded. In
addition to the physical property listed, some reference, including pagination, to the
source of the information should be cited. Included below are a few valuable sources of
physical property data. Copies of these sources of material can be found in the Thomas
Jefferson Library and also in the laboratory. A variety of editions are available for most.

"The Dictionary of Organic Compounds" Buckingham, J., editor, Chapman and Hall.
"CRC Handbook of Chemistry and Physics" Weast, R. C, editor, CRC Press.
"Aldrich Catalog Handbook of Fine Chemicals", Aldrich Chemical Co.
www.sigmaaldrich.com
"Merck Index", Windholz, M. editor, Rahway N.J.
"Lange's Handbook of Chemistry", Dean, J. A., editor, McGraw-Hill.

The laboratory notebook is the permanent record documenting your work in the
laboratory during a specific period of time. It should contain sufficient details and
documentation that an individual with similar training could repeat your work months or
years later. The laboratory notebook should contain the original data, not copies of the
data. The specific requirements for entries will vary from experiment to experiment and
from instructor to instructor. The following is intended to serve as a guideline of what to
include in the notebook. A substantial part of the grade earned in this course will be
derived from the contents of your notebook.

a. You will be using a bound notebook produced by Chemical Education Resources, Inc.,
available in the University Bookstore. This notebook produces a copy that can be removed
and handed in when each experiment is due. Suggestions on how to properly use the
notebook are available in the notebook on page ii.

10
b. You may leave some blank space on a page or additional blank pages in order to
complete an experiment. A notation instructing the reader to the location where the
remaining portion of the experiment is to be found, is also acceptable. You may wish to
include supporting documentation such as spectra that you have recorded. This
documentation can be attached to the notebook with staples or tape or preferably, in a
separate folder. In addition you may wish to sketch your thin layer chromatography results
in your notebook since the actual plates may not survive repeated handling.

c. Some notebook preparation should be completed before you come to the


laboratory. What is required will vary from instructor to instructor. However, the
following will apply to all.

Begin with a title of the experiment and a brief sentence summarizing the purpose
of the experiment and specify a reference (e.g., UMSL Chemistry 2633 Manual, pp. 18-
19). If the experiment involves a synthesis, include a balanced chemical equation, and
physical properties of all chemicals involved. Include comments on toxicity if appropriate,
any side reactions and anything out of the ordinary. Prepare an outline of the procedure
and what you plan to do in the laboratory. The outline should be completed before you
come to the laboratory but may be on a separate sheet of paper instead of the lab
notebook. This will enable you to work more efficiently.

In the laboratory, you should record in the notebook the masses of the
reagents you used. Include what you actually did and observed as you performed the
experiment. Be as brief as possible. Do not simply copy the experiment from the manual
to your notebook. Include a sketch of the apparatus, if one is involved. Always include
the date you did the work on the outside page margin. Be sure to include your
observations. You may also include any conclusions that you reached. When you
comment on specific tests, (e.g., the ferric chloride test) do not simply say the test was
negative, add that it means you do not have a phenol or that you do not have salicylic acid
present.

It may not be possible to follow this format for each experiment you will be
performing this semester. However, for those experiments that involve some synthetic
component, following this format will suffice. It is important to remember that your
notebook is the record of your work and observations in the laboratory. If you keep this in
mind you will stay on the right track. The notebook should contain sufficient information
so that a reader knowledgeable in the subject should be able to reproduce your results.

Notebook Calculations
The number of calculations you will need to perform in the organic laboratory is quite
small but those that you will be required to perform are very important. The calculations
are similar to the calculations you performed in introductory chemistry and be sure to

11
review them if you have forgotten the details and theory behind them. What follows is a
brief summary of some important definitions.
The calculations associated with conversion of the starting materials to
product is based on the assumption that the reaction will follow simple ideal
stoichiometry. For example, in the preparation of aspirin from salicylic acid and acetic
anhydride calculate the theoretical and actual yield by assuming that all of the starting
material is converted to product. This is not correct since some of the starting material
actually forms a polymer as a consequence of the reaction conditions and catalyst that is
used.

The first step in calculating yields is to determine the limiting reagent. The
limiting reagent in a reaction that involves two or more reactants is simply the reagent that
is present in lowest molar amount based on the stoichiometry of the reaction. Once this
reagent is consumed additional conversion to product will cease. In the reaction of
salicylic acid with acetic anhydride, the latter reagent is used in excess. There are several
reasons for doing so. An important consideration whenever any reagent is used in excess,
is to determine how this reagent can be separated from the product at the end of the
reaction. In this case, the acetic anhydride which is not water soluble can be allowed to
come in contact with water. During this time it will slowly hydrolyze to acetic acid. The
acetic acid is water soluble and the aspirin can be separated from the aqueous solution of
acetic acid by filtration.

Salicyclic acid is the limiting reagent in the synthesis of aspirin. As far as the
calculation of the theoretical yield is concerned, we assume that every mole of salicylic
acid is converted to aspirin. According to the stoichiometry of the reaction, one mole of
salicyclic acid will be converted to one mole of aspirin. Therefore, from the number of
moles of salicylic acid we used, we evaluate the maximum amount of aspirin that can be
obtained. Multiplying the number of moles of aspirin by its molecular weight results in the
theoretical yield. The theoretical yield is usually reported in grams. In all cases, the
calculation is performed similarly, regardless of how many products are formed. However,
we can now evaluate the actual yield by determining how much aspirin we have actually
isolated, experimentally. The % yield is simply the ratio of the actual yield divided by the
theoretical yield times 100.

In a multi-step synthesis, each step in the process is characterized by a


limiting reagent, an actual yield and a % yield. The calculation of the overall yield of
the entire process can be calculated by identifying the limiting reagent in each step. The
overall yield of the process will simply be the product of the actual yield divided by the
theoretical yield of the first step times the same ratio for each step in the process times
100. Thus in a two step process, if you get a 0.5 yield (50%) in the first step and the same
yield in the second step, your overall yield will be 0.5 × 0.5 × 100 = 25 %. You can
imagine that a multistep process operating at this efficiency is not likely to be very useful
commercially.

12
Labels
Properly label your sample vials when you submit your products for grading.

BENZOIC ACID (appropriate, unambiguous name of the compound)


mp 120-121 °C (melting point observed for this sample)
3.2 g (amount in vial)
D.L. Green (the name of the person submitting the sample)
1/15/14 (date submitted)
DLG 12-14 (which is a reference to notebook pages concerning this material).
A hazard statement for the chemical where applicable (take the information from the SDS
sheet)

Properly label your spectra. The infrared spectra you decide to save should be labeled
immediately. List the name of the compound or some identifying term if the name is
unknown, your name, and the method by which the spectrum was obtained (Nujol mull,
neat melt, KBr pellet, etc.). The instrument will automatically print the date but if it
doesn’t, be certain that you list the date it was recorded on the spectrum.

Basic Techniques in Organic Chemistry

Crystallization is used to purify a solid. The process requires a suitable solvent. A


suitable solvent is one which readily dissolves the solid (solute) when the solvent is hot
but much less so when it is cold. The best solvents exhibit a large difference in solubility
over a reasonable range of temperatures. (e.g., water can be a crystallization solvent
between 0-100oC; hydrocarbon solvents such as hexanes or petroleum ether have a
different temperature range since they can be cooled below 0 degrees but boil below 100
°C).

Characteristics of a solvent:
a) chosen for solubilizing power - solubility usually increases with increasing
temperature;
b) polarity is important--like dissolves like; polar compounds are more soluble in
polar solvents; non-polar compounds in non-polar solvents;
c) should be INERT but few are; e.g., acetic acid is sometimes used as a solvent
although it will certainly react with basic compounds;
d) most solvents are COMBUSTABLE - avoid flames;
e) mixed solvents (e.g.; 1:1 water/methanol) provide a huge range of possible
solvents but they must be miscible in one another.

Question: Is 95% ethanol a mixed solvent?

Getting solutes out of solution:

13
a) lower the temperature - solute will be less soluble;
b) concentrate the solution by removing solvent (in a hood) on a hot plate, steam
bath, or with using a rotary evaporator.

To remove solvent:
1. You must have bubbling to concentrate at atmospheric pressure-use a boiling
stone, a capillary tube, or agitation to prevent bumping (i.e. the very rapid boiling
of a liquid overheated above its boiling point, resulting in splattering).
2. If you used reduced pressure to concentrate the solution, use the water aspirator
with a TRAP in the line. DO NOT turn off the water until the pressure is released.
In general, CLAMP any flask that could conceivably tip over.
3. Do NOT use a boiling stone if using the rotary evaporator. The rotation provides
sufficient agitation.

Crystallization (summary)
Used to obtain pure crystalline solid. Use the proper solvent or solvents - test if necessary;
a proper solvent will exhibit a big solubility difference over a small temperature range.

Recrystallization or crystallization
a. use an Erlenmeyer flask, it is specifically designed for this purpose;
b. dissolve solid in a minimum amount of boiling solvent - add solvent in
small amounts. For example, if you add 5 mL and approx. half of the solid
dissolves, it should take only another 5 mL to dissolve the remaining half.
If some of the solid does not dissolve then filter it off or decant it.
c. are remaining particles your compound or insoluble material (e.g., sand,
old boiling stones)?
d. to determine this, add ca. 10% more hot solvent. If the material remains
insoluble, you can decant (carefully transfer solution into another flask
leaving the insoluble material behind) or filter.
e. if filtering is necessary, do so to remove suspended solids - the faster the
better - keep solution near or at the boiling point so crystallization does not
occur (this may require filtering on hot plate or other heating device).
f. to decolorize a solution, use a small amount of charcoal and filter with
filter aid (see below). (both (e) and (f))- be sure to rinse filter paper and
flask with a small amount of hot solvent.
g. let the filtered liquid (filtrate) cool to room temperature slowly in the
Erlenmeyer flask
h. cool the filtrate in an ice-water bath after the solution reaches room
temperature.
i. if crystals have not formed
i. seed with a small crystal of crude product, or
ii. scratch the flask with a glass rod which has not been fire polished at
the end (ask for a demonstration), or

14
iii. heat to boiling and add a second solvent dropwise until the cloud
point is reached; the cloudiness suggests that the solute has reached
a saturation point in this new mixed solvent and will start to come
out of solution.
j. if material oils out, you must re-dissolve by heating the solution and then
proceed again from part (g). Try adding a little more solvent.
k. if your material precipitates out, then you cooled the solution too quickly.

Safety note regarding hot glassware: use heat resistant gloves when touching hot
glassware. Avoid temperature extremes when heating and cooling glassware to prevent
unintended breakages.

Decolorization:
Most organic compounds are colorless. Highly conjugated compounds (e.g., dyes) will
absorb light in the visible region of the spectrum and thus be colored. If these highly
polar, large molecules are impurities, they can be removed by use of finely granulated
activated charcoal (“Norit”). Polar compounds (e.g., polar impurities) are adsorbed on the
charcoal which is insoluble in the solvent and can be filtered away from solution.
Unfortunately, some of your compound will also adsorb if there is enough charcoal so the
trick is to use just the right amount. Usually, a very small amount of charcoal will suffice
(there is a lot of surface area to these particles). The Norit is added in small amounts to the
hot (but not boiling) solution until sufficient decolorization has occurred. CAUTION--
trapped air in the Norit can cause rapid frothing when it hits the hot solution. The Norit
can be filtered from the hot solution using fine filter paper. The solution must be filtered
hot or your compound will crystallize out on the filter paper.

Filtering
Used to remove insoluble solids suspended in solution.
Use a GRAVITY filtration when you DON’T want the solid
Use a HIRSCH or BUCHNER funnel with vacuum when you do want the solid but....
Never use a Hirsch or Buchner funnel with a hot solution.

For the Hirsch and Buchner funnels, use a piece of round filter paper which fits the
funnel. A proper fit is a piece of paper that just covers the holes but does not touch the
sides of the funnel (ask for a demo). You will need to use reduced pressure, usually via a
water aspirator (be certain that the TRAP is clean). For aqueous solutions, you can use the
house vacuum line (be certain that the TRAP is clean).
Just prior to filtering, wet the filter paper with a few mL of solvent and apply the vacuum.
This will cause the filter paper to “stick” to the funnel.

For the gravity funnel, you can prepare a suitable filter from round filter paper by
folding into quarters or by folding it into more than quarters (fluted like a fan). Ask for a
demo. Another trick is to use a piece of cotton or glass wool loosely wedged into the cone

15
of the funnel. It must be tight enough to not move and to trap the solid particles but loose
enough to allow the solution to flow through. This procedure is often used to remove
drying agents and other coarse solids from the solution.

Receiver Flasks: For gravity filtration of a hot solution, use an Erlenmeyer flask. It has
been designed to prevent loss and rapid cooling of your solvent. For a Hirsch or Buchner
funnel, you will need to use a filter flask with a sidearm for a rubber tubing connection to
the TRAP and the reduced pressure source. Use an appropriately sized filter flask, one that
will not fill more than halfway and secure this flask so that it does not tip over with that
expensive funnel containing your valuable crystals. Be sure to use the thick walled
rubber tubing when using reduced pressure.

Trap: For vacuum filtration, you want a clean glass trap in between your filter flask and
the vacuum source. One reason is obvious--in the event that your filtrate is drawn out of
the filter flask, it can be trapped and recovered before it goes down the drain or into the
house vacuum line. Another reason is that a changing flow of water affects the pressure in
the water aspirator so that water can back up and flow into your filter flask. This way, the
trap will fill first and prevent contamination of your filtrate.

Rinsing: Gravity Filtration: You should rinse the gravity funnel with a small amount of
solvent to wash down the remaining solution that adheres to the filter paper and funnel.
Remember, you want the solution, you do not want the solid. Use a small amount of
solvent to prevent excess dilution of the filtrate.

Vacuum Filtration: You should rinse the crystals in the Buchner and Hirsch funnel with
a minimum amount of cold solvent. Remember, you want the crystals and do not want to
dissolve the crystals with excess solvent which will wash them into the filtrate.

The proper way to wash the crystals is to SHUT OFF the vacuum, add a minimum
amount of cold solvent so that the crystals are barely sitting in solvent for about 5 seconds
(the solvent will not drip through quickly) and then apply the vacuum. The solvent will be
drawn into the filtrate by the vacuum. Do this one or two times for each solid. You can
air dry the solid by drawing air through the solids. To dry a large amount of solid more
rapidly, press another piece of filter paper on top of the solid.

Filtrate from Buchner and Hirsch filtration: The filtrate will probably still contain some
of your desired compound and is called the mother liquor. Often, by concentrating this
solution further and cooling the solution, one can obtain more crystals. This may happen
while you are filtering the original crystals under reduced pressure. To collect this second
batch of crystals, filter as before but do not combine them with the first batch of crystals.
The purity of this second crop will usually be different from the first batch. Use the
melting point or other methods (e.g., thin layer chromatography) to determine whether the
purity of the second crop is equal to that of the first. If so, they can be combined. If not,
keep them separate.

16
Solvent Polarity in decreasing order:
water
amides (N,N-dimethylformamide)
alcohols (methanol, ethanol)
ketones (acetone, methyl ethyl ketone)
esters (ethyl acetate)
chlorocarbons (methylene chloride, chloroform)
ethers (diethyl ether)
aromatics (toluene)
alkanes (hexanes, petroleum ether)

Heating:

There are different methods used for heating material in the laboratory. Flames are never
used in the laboratory except in controlled situations (e.g., isolated in fume hoods).
Electric hot plates and heating mantles are most commonly used. Be careful not to turn
this equipment to its highest setting initially since the mantle can get very hot. It does take
several minutes for these instruments to reach the desired temperature. The heating
mantles are plugged into a variable rheostat which provides a temperature control. Heating
mantles are used for round-bottom flasks (rbf); choose an appropriate size to fit the flask
you plan to use. Steam is often used for heating volatile, non-aqueous, flammable solvents
in the fume hoods. This is preferable to using hot plates which can lead to flash fires. The
100 ºC maximum temperature from steam is sufficient for most commonly used organic
solvents including many mixed aqueous solvents.

Steam bath vs hot plate:

Advantages of steam baths:


 can only heat to 100 °C so cannot overheat the compounds in solution
 flash fires are unlikely with most solvents
 provides instant heat.

Disadvantages of steam baths:


 can only heat to 100 °C so may not be suitable for higher boiling solvents
 steam condenses to water which can get into your flask
 steam line is not as portable as a hot plate
 temperature control is more difficult

Melting Point (mp) Determination:


A standard physical property of a solid is its melting point. The melting point (mp)
is usually reported as a melting point range. It is used to help determine the purity of a

17
solid and to help verify the identity of the compound. A pure compound should melt over
a narrow temperature range. Impurities usually cause the melting point range to widen and
result in a lower value. To obtain the melting point range, you need to record the
temperature at which the first crystals begin to melt (solid to liquid phase) and the
temperature at which the last crystal melts, e.g.; mp 124-126 °C. Important
considerations include: 1. do you have a representative sample of your compound or have
you fished out the best looking crystal for a melting point; and 2. is the thermometer you
are using correctly calibrated. If not, you will obtain incorrect values.

There are several different mp apparati in the laboratory. The most common is a
Fisher-Johns apparatus whereby the crystalline sample is sandwiched between two glass
discs and placed on a small hot plate. The temperature of the hot plate can be adjusted
using the rheostat on the instrument. Temperatures respond fairly quickly to settings of the
rheostat and can be read from the thermometer inserted into to the hot plate. You assume
in using this apparatus that the temperature of the crystals is equal to the temperature of
the hot block as measured by the thermometer. This is only correct if the rate of heating is
low, 1-3 degrees/minute. You may assume that the thermometers read correctly although
this can be tested and calibrated by determining the mp of standard compounds.

Preparing your sample for the Fisher-Johns apparatus: Take a small amount of your
crystalline material (less than 1 mg!) and place it on a glass disc. Cover with another glass
disc and carefully crush the two together, rotating one disc over the other to produce a
finely ground material. Place this sandwich on the hot plate and heat rapidly (10-20
degrees/min) until the material melts. This will give you a rapid, but inaccurate reading.
Now cool the hot plate down to about 20 degrees below the observed inaccurate mp and
repeat the experiment. Increase the temperature rapidly until you are 10-15 degrees below
the expected mp at which time lower the rate of heating to 1-2 degrees/minute. Common
errors: rate of heating is too rapid (results in a lower mp); amount of material is too large
(will increase your mp range); crystals are not ground flat (results in a higher mp and
broad mp range since an air gap is present which helps to cool the crystals); crystals are
not dry. To rapidly dry a few wet crystals, place them onto a small piece of filter paper,
rub them onto the paper with a clean spatula until all of the liquid is drawn into the paper.
Carefully scrape the crystals away from the paper.

The Mel-Temp apparatus and Thomas-Hoover apparatus: The dry, crystalline sample is
placed in a capillary tube sealed at one end which will be heated by air or by oil. One
advantage over a Fisher-Johns apparatus is that 3-5 samples can be observed at the same
time allowing you to take the melting point of many samples simultaneously. Also, it is
possible to take the mp of a sample that sublimes prior to its mp (one uses a sealed
capillary). Once again, a rheostat is used to control the heating rate. Ask for a
demonstration.

Preparing the sample in a capillary tube: The sample must be dry or else you cannot get it
into the capillary tube. The capillary tube should be open at one end only. Tap the open
end onto your crystals until a few collect inside the mouth of the capillary. Now tap the
crystals to the closed bottom end by vibration (there is a vibrator on the Thomas-Hoover

18
apparatus) and/or by bouncing the capillary (bottom end down) through the open space of
a large piece of glass tubing. Ask for a demo. Follow the instructions given above for the
rate of heating.

Extraction
Extraction is a method for moving a compound from one medium to another. For example,
if you make coffee from coffee beans, you are extracting some flavorful components of
the bean and some caffeine into the water. The remainder of the beans (grounds) are left
behind and discarded. This is called a solid-liquid extraction. If you are trying to move a
compound from one liquid phase (solvent 1) into another liquid phase (solvent 2), this is
liquid-liquid extraction. The two solvents must be immiscible or insoluble with each other
to the extent that they form two distinct layers (why?). Any solute will distribute itself into
the two solvent layers based on its solubility in the two solvents. These two layers can be
physically separated using a separatory funnel. By measuring the concentrations of a
compound in two immiscible solvents (c1 and c2) we obtain a distribution coefficient, K,
which is a constant for a given compound and given solvents at a fixed temperature. This
is irrespective of the amounts of solvent present (remember concentration units are given
in quantity of solute/unit of solvent). A simple equation illustrates the relationship:
K=c2/c1. Liquid-liquid extractions are common in organic chemistry. Often, water-
insoluble organic solvents, such as ether, methylene chloride and hexane, may contain
some undesirable water soluble components (like acetic acid (aspirin synthesis)). In that
case, we would extract those components out of the organic solvent by using water as the
second solvent. That is often called a water wash. You will have an opportunity to do
several extractions and water washes. A separatory funnel is used for these purposes.

The rule of thumb for liquid-liquid extractions is that several small extractions are
more efficient than one big extraction. Test this out as follows:
If K=6.0, then c2/c1=6.0 or (g/mL2)/(g/mL1)=6. If you have 2.0 g in 200 mL of water and
want to extract with a total of 100 mL of solvent 2, let us determine if it is better to use
one extraction with 100 mL or two 50 mL extractions. The distribution in the former case
is (x/100 mL)/(2.0-x/200 mL)=6.0 where x is grams in solvent 2 and 2-x is grams in
water. Solving for x, gives us x=1.5 g, leaving 0.5 g in the water.

In the latter case, we have (x/50 mL)/(2.0-x/200 mL)=6.0; solving for x gives us
x=1.2 g leaving 0.8 g in the water for the first extraction. The second extraction is (x/50
mL)/(0.8-x/200 mL)=6.0; solving for x=0.48; which means that in the two 50 mL
extractions we obtained 1.2 + 0.48 = 1.68 g, leaving 0.32 g in the water. The two
extractions were clearly more efficient than one extraction. Four 25 mL extractions would
have been even better but there are practical limits.

Separatory Funnel: This glass equipment is very cleverly designed to carry out the task
of separating two immiscible liquids (which form two distinct layers). Work with this
equipment in a proper fashion and it will perform remarkably well. However, read the
instructions first and follow the steps carefully.

19
Use of the Separatory Funnel:
1. First, check that the stopper fits and that the stopcock works properly. Glass
stopcocks must be greased; Teflon stopcocks do not need grease.
2. Close the stopcock and support the funnel using a ring clamp.
3. Place a beaker underneath the funnel to catch any spills or leaks.
4. Fill the funnel with the two solvents. Do not fill the funnel more than 75% of
capacity (so plan ahead in choosing the proper size funnel).
5. Stopper. Remove the funnel from the stand, hold properly and invert
(IMPORTANT, ask for a demo), release any pressure buildup by opening the
stopcock repeatedly BEFORE shaking (and then after shaking).
6. After returning the funnel to the stand where it is secured, loosen the stopper
immediately--shaking builds up pressure
7. When the two phases separate, draw off the lower layer.
8. Pour out the upper layer if necessary. NOTE: The funnel is designed to retain the
last few drops.
9. Save both upper and lower layers until you are certain that you have the compound
you want. If you do not throw it away, it is not lost. Remember, never throw away
a layer unless you are certain that you will never need it.

Drying Agents
When an organic solvent has been exposed to water, it will dissolve a small
amount of it, the amount depending on the solubility of water in the solvent. This water
can be a nuisance for a variety of reasons including the fact that its presence can cause
hydrolysis of many organic compounds and that it can cause cloudiness and separation of
two phases when solutions are concentrated. A general practice in the laboratory is to dry
solutions that have come in contact with water using an appropriate drying agent. A
drying agent is usually an anhydrous inorganic salt which can react with the water present
to form a hydrate. Anhydrous MgSO4, for example, reacts with water to form the
heptahydrate MgSO4·7H2O. Drying agents are characterized by their capacity (the amount
of water they can absorb), the rate at which they absorb the water, and their intensity (or
completeness), which is the amount of water left behind in the solvent at equilibrium.

Some typical drying agents are listed below. They are neutral (as compared to
acidic or basic) and for general use (applicable to most classes of compounds).

DRYING AGENT HYDRATED FORM CAPACITY RATE of INTENSITY


DRYING (COMPLETENESS
)
Magnesium Sulfate MgSO4·7H2O High Rapid Fair
Sodium Sulfate Na2SO4·7H2O High Medium Poor
Calcium Sulfate CaSO4·2H2O Low Rapid Good
Molecular Sieves _____ High Slow Very Good

20
The organic solution may appear milky when wet; it will clear up when the drying
agent is added. Anhydrous magnesium sulfate is a free-flowing powder which cakes and
sticks to the bottom of the flask as it becomes hydrated. Add just enough so when you
swirl the flask a few crystals will circulate. Never add a large amount of drying agent
because your product can become absorbed on the surface, reducing the yield. Enough
should be added to provide a thin layer over most of the bottom of the flask.

Extraction Solvents
Selection of a good extraction solvent is similar to selection of a recrystallization
solvent in several respects. Carbon tetrachloride and chloroform are excellent extraction
solvents for many solutes but are too toxic to use in the undergraduate lab.
Dichloromethane possesses a good balance of properties and is used frequently in our
laboratory. Diethyl ether is an excellent solvent but constitutes a severe fire hazard. Ethyl
acetate is a good selection for the extraction of moderately polar solutes from aqueous
solution. Benzene is too toxic to use but toluene and hexane are often used to extract non-
polar solutes. The solvent chosen must be a good solvent for the solute in question, be
moderately non-toxic, must not react with the solute, must not be miscible with the other
solvent, and should be non-flammable or at least have a relatively low vapor pressure at
room temperature.

Problems
1. In a multiple synthesis, the first reaction produces 1.0 g of product in 85% yield; in the second reaction
you use 0.50 g of the first product to obtain the second product in 40% yield. The overall % yield for these
two steps is 100 times: a. 0.85 × (0.50/1.0) × 0.40; b. 0.85 × 0.40; c. 0.50 × 0.85 × 0.40.

2. If 0.345 g of product is added to 0.0214 g of product, how much product do you have total in mg?

3. If 5.6 mL of a compound with d=0.854 g/mL is 13 mmoles (0.013 moles), what is the MW of the
compound. Use significant figures.

4. How many mL of 2.0 M H2SO4 is needed to neutralize 1.0 g of NaOH (MW=40 g/mol)?

5. If compound A is soluble in water (2.0g in 100 mL at 20 ºC and 5.0 g in 100 mL at 40 ºC), what is the
minimum amount of water necessary at 40 ºC to dissolve 1.4 g of compound A? Cooling this solution to
20oC should precipitate how much solid A?

6. If A + B =C; the MW of A=100, B=200, C=300 (can it be any other value?); if you start with 1 g of A
and 2 g of B and obtain 2.4 g of C, what is the % yield?

7. If A + B = 2C (MW in g/mole of A=100, B=200, C=150), and in your experiment you used 1.0 g of A
and 1.0 g of B and you obtained a yield of 1.0 g of C, what is the limiting reagent, A or B? Calculate the
theoretical yield (in grams) and your % yield.

8. List four properties of a good solvent.

9. The distribution coefficient (ether/water) for compound Z is 10. If 10 mL of ether is used to extract 1.0 g
of Z from 100 mL of water, will there be more compound Z in the ether or the water or an equal amount in
both? How much Z will be left in the water if there are two 10 mL ether extractions?

21
EXPERIMENT 1. SYNTHESIS OF ASPIRIN
Pharmacologically active compounds
Over history, many compounds obtained from nature have been used to cure ills or to
produce an effect in humans. These natural products have been obtained from plants,
minerals, and animals. In addition, various transformations of these and other compounds
have led to even more medically useful compounds. During this semester, you will have
an opportunity to isolate some pharmacologically active natural products and to synthesize
other active compounds from suitable starting materials.

Analgesics are compounds used to reduce pain, antipyretics are compounds used to
reduce fever. One popular drug that does both is aspirin. The Merck Index, which is an
encyclopedia of chemicals, drugs and botanicals, lists the following information under
aspirin: acetylsalicylic acid; monoclinic tablets or needle-like crystals; mp 135 (rapid
heating); is odorless, but in moist air it is gradually hydrolyzed into salicylic and acetic
acids; one gram dissolves in 300 mL of water at 25 °C, in 100 mL of water at 37 °C, in 5
mL alcohol, in 17 mL chloroform.

SYNTHESIS OF ASPIRIN (acetylsalicylic acid)

O
OH O CCH3
CO2H O O CO2H O
H2SO4
+ CH3C O CCH3 + CH3C O H

1. Place 2.0 g (0.015 mole) of salicylic acid in a 125-mL Erlenmeyer flask.

2. Add 5 mL (0.05 mole) of acetic anhydride, followed by 5 drops of conc. H2SO4 (use
a dropper, H2SO4 is highly corrosive) and swirl the flask gently until the salicylic
acid dissolves.

3. Heat the flask gently on the steam bath for at least 10 minutes.

4. Allow the flask to cool to room temperature. If acetylsalicylic acid does not begin to
crystallize out, scratch the walls of the flask with a glass rod. Cool the mixture slightly
in an ice bath until crystallization is completed. The product will appear as a solid
mass when crystallization is completed.

5. Add 50 mL of water and cool the mixture in an ice bath. Do not add the water until
crystal formation is complete.

6. Vacuum filter the product using a Buchner funnel. You can use some of the filtrate to
rinse the Erlenmeyer flask if necessary.

22
7. Rinse the crystals several times with small portions (~5 mL) of cold water and air dry
the crystals on a Buchner funnel by suction until the crystals appear to be free of
solvent. Test this crude product for the presence of unreacted salicylic acid using the
ferric chloride test.

8. Stir the crude solid with 25 mL of a saturated aqueous sodium bicarbonate solution in
a 150 mL beaker until all signs of reaction have ceased (evolution of CO2 ceases).

9. Filter the solution through a Buchner funnel to remove any insoluble impurities or
polymers that may have been formed. Wash the beaker and the funnel with 5 to 10 mL
of water.

10. Carefully pour the filtrate with stirring, a small amount at a time, into an ice cold HCl
solution (ca 3.5 mL of conc. HCl in 10 mL of water) in a 150-mL beaker and cool the
mixture in an ice bath. Make sure that the resulting solution is acidic (pH ~2, use
Hydrion paper) and that the aspirin has completely precipitated out.

11. Filter the solid by suction and wash the crystals 3 times with 5 mL of cold water each.
Remove all the liquid from the crystals by pressing with a clean stopper or cork. Air
dry the crystals and transfer them to a watch glass to dry. Test a small amount of the
product for the presence of unreacted salicylic acid using the ferric chloride solution.

12. When the product is completely dry, weigh the product, determine its melting point
(literature mp 135-136 °C) and calculate the percentage yield.

13. Dissolve the final product in a minimum amount (no more than 2-3 mL) of hot ethyl
acetate in a 25 mL Erlenmeyer flask. Make sure that the product is completely
dissolved while gently and continuously heating on a steam bath.

14. Allow the solution to cool to room temperature and then cool in a ice-bath. Collect
the product by vacuum filtration and rinse out of the flask with a few milliliters of
cold petroleum ether.

15. When the product is completely dry, weigh it, determine its melting point (lit mp 135
°C) and calculate the percentage yield of this recrystallized product. Calculate the %
recovery of recrystallized material from crude material. Take an IR of the product.
Submit the crystalline sample in a small vial with proper labeling to your instructor.

FERRIC CHLORIDE TEST FOR SALICYLIC ACID


Add 10 drops of aqueous 1% ferric chloride solution to a test tube containing a few
crystals of the compound to be tested dissolved in 5 mL water and note the color. Do this
test with 1) phenol, 2) salicylic acid, and 3) your crude product. Formation of an
iron-phenol complex with Fe(III) gives a definite color ranging from red to violet,
depending upon the particular phenol present.

23
Questions:

1. Normally, you measure reactants to at least two significant figures. Why is it not
necessary to measure the volume of acetic anhydride to two significant figures?
What is the theoretical yield of aspirin (in two significant figures).
2. Why is the aspirin washed with cold water?
3. According to the data in the Merck Index, if 1.0 g of aspirin is dissolved in 100 mL
of water at 37 °C, how much aspirin will precipitate out of solution when it is
cooled to 25 °C?
4. A polymeric material, which is a polyester, is formed in this reaction. Polyesters
are often made from dicarboxylic acids and diols. In this case, one molecule
(salicylic acid) provides both the “alcohol” and the carboxylic acid. Write a
structure for the ester formed from acetic acid and ethanol. If you have difficulty,
look at the banana oil experiment.

24
EXPERIMENT 2. SYNTHESIS OF
ACETAMINOPHEN
Acetaminophen-- Another pharmacologically active compound

Analgesics are compounds used to reduce pain, antipyretics are compounds used to reduce
fever. One popular drug that does both is aspirin, another is acetaminophen which is often
used by people who have unwanted, harmful side effects to aspirin. Acetaminophen,
which can be synthesized from p-aminophenol, is probably best recognized under the
trade name Tylenol. The Merck Index lists the following information under
acetaminophen: large monoclinic prisms from water, mp 169-170.5 °C, very slightly
soluble in cold water, considerably more sol in hot water. Soluble in methanol, ethanol,
dimethylformamide, acetone, ethyl acetate. Practically insoluble in petroleum ether,
pentane.

O
NH2 H N CCH3
O O O
+ CH3C O CCH3 + CH3C O H

OH OH

Acetaminophen (CAUTION: p-aminophenol is hazardous to skin, concentrated


HCl is corrosive). Practical grade p-aminophenol contains impurities that must be
removed at the beginning of the synthesis. We will use decolorizing charcoal (Norit) and
water for that purpose. The sequence involves first solubilizing the water insoluble amine
by converting it into a water soluble amine hydrochloride salt, then decolorizing if
necessary. In order to acylate the amine, it is necessary first to partially neutralize the
amine hydrochloride. This is accomplished with a sodium acetate buffer, immediately
followed by addition of the acylating agent. The buffer partially converts the amine
hydrochloride to the free amine which remains in solution and can react with acetic
anhydride. As it reacts, more of the hydrochloride is converted back to the free amine by
the buffer until most of it reacts with the acetic anhydride.

Procedure: The quality of p-aminophenol can vary greatly depending on the source. If
good quality starting material is used it may be possible to skip the charcoal treatment
step. Weigh 2.1 g of p-aminophenol into a 125-mL Erlenmeyer flask and add 35 mL of
water followed by 1.5 mL of concentrated hydrochloric acid. Swirl the flask in an attempt
to dissolve the amine hydrochloride (it may be difficult to determine this since the solution
may be very dark). Add 0.3-0.4 g of decolorizing charcoal (Norit) to the solution (this is
much more than usual but necessary if the crude p-aminophenol contains a lot of

25
polymeric material), swirl the solution on a steam bath for 4-8 minutes and periodically
check to see if the solution is decolorizing (it will be difficult to determine since the
solution may be dark). Remove the charcoal by gravity filtration into another secured 125-
mL Erlenmeyer flask using fluted filter paper while the solution is warm. The flask is
secured to prevent tipping. Rinse the filter paper with 1 mL of water. If the charcoal
comes through the filter paper it may be necessary to refilter or to use a filter aid, Celite.
The filtrate may be clear or, more likely, a tea color. If the solution is a dark brown, add
0.1 g of Norit, heat on the steam bath for a few minutes and filter. The filtrate will
darken with time!

While decolorizing the p-aminophenol, prepare a buffer solution by dissolving 2.5


g of sodium acetate trihydrate in 7.5 mL of water which will give 8.8 mL of solution.

Warm the filtered aqueous p-aminophenol hydrochloride solution on a steam bath,


then add the buffer solution in one portion with swirling. Immediately add 2.0 mL of
acetic anhydride while continuing to swirl the solution. Continue heating on the steam
bath while swirling vigorously for 10 minutes.

Cool the solution in an ice-water bath, stirring with a glass rod until the crude
acetaminophen begins to crystallize. A little bit of rubbing/scratching with a glass rod near
the surface often stimulates the crystallization. After crystallization begins, allow the
solution to sit in the ice bath for almost an hour. Filter your product using a Buchner
funnel and the water aspirator or house vacuum line. Wash (rinse) the crystals once with a
minimum amount of cold water (a few mL should suffice).

Recrystallize all but a few mg of your crude acetaminophen from water by first
dissolving the solid in the minimum amount of hot (boiling) water. Do this carefully
adding small amounts of hot water. You do not want to have excess water. Work on a hot
plate to keep the solution hot. Add another 2 mL of hot water. If there are no insoluble
particles in the solution, you can allow it to cool slowly without having to first filter. If
not, decant the hot solution or try to remove the particles with a spatula or Pasteur pipette
while keeping the solution hot. Alternatively, you may have to gravity filter the hot
solution. After crystallization begins and the solution reaches room temperature, cool the
solution using an ice bath. When crystallization ceases (15 minutes), collect the crystals as
before, rinsing once with a few mL of cold water, and air drying. Record the weight of the
dry, recrystallized acetaminophen. Record the percent yield of dry recrystallized product.

Take a mp of your crude and recrystallized acetaminophen (literature mp 169-170.5 °C).

26
EXPERIMENT 3. CAFFEINE EXPERIMENT
Caffeine is a minor constituent of tea, coffee, and other natural plant materials. The major
constituent of tea is cellulose which is not water soluble. Caffeine is water soluble but so
are some tannins and gallic acid which is formed in the process of boiling tea leaves. The
latter two components can be converted to their calcium salts which are insoluble in
water. The caffeine can then be extracted from the water by methylene chloride in almost
pure form. Some chlorophyll is often extracted at the same time.

Procedure: Place 10 g of tea leaves, 4 g of calcium carbonate powder and 200 mL (to the
nearest mL) of water into a 600 mL beaker. Boil the solution on a hot plate for 10
minutes with occasional stirring. Cool the solution but, while it is still warm, vacuum
filter through a Buchner funnel using a fast filter paper, if available. Normally, hot
solutions are not vacuum filtered. Do not rinse the leaves with any additional water.
Carefully press out as much filtrate as possible since the caffeine is in the aqueous layer.
Measure and record the volume of the filtrate recovered.

Cool the solution to room temperature and pour it into a 250 mL separatory funnel.
Extract with 35mL of methylene chloride. Carefully shake the separatory funnel in order
to extract the caffeine but try not to form an emulsion. Relieve the pressure buildup as
soon as you mix the two liquids. Then shake carefully for 10 seconds and relieve
pressure, repeat the shaking two more times. An emulsion may form. To break the
emulsion formed in the methylene chloride layer, slowly drain the methylene chloride
layer through a small amount of anhydrous magnesium sulphate in a powder funnel with
a loose cotton plug (a tight plug will prevent drainage). Extract the aqueous solution
twice again with a 35 mL of methylene chloride, repeating the steps above to collect the
lower layer. Combine the methylene chloride extracts and, if necessary, dry the
methylene chloride layer further with additional anhydrous magnesium sulphate.

The methylene chloride solution will be removed on a rotary evaporator. Weigh a


150-mL round bottom flask and transfer the dried methylene chloride solution to it. Be
certain that there is no magnesium sulphate in the solution. Removing the methylene
chloride should take less than 5 minutes. You will be left with a small amount of residue
with a greenish tinge. Obtain the weight of crude caffeine by difference. Add 5-8 mL of
hot acetone to dissolve the crude caffeine and transfer the solution to a 50 mL
Erlenmeyer flask for recrystallization. Add a few drops of petroleum ether until you
reach the cloud point (caffeine is less soluble in this mixed solvent) and then cool the
solution. If you do not get a precipitate, you may have used too much acetone. Carefully
boil off the excess on a steam bath or hot plate (in the hood) using a boiling stick or stone
to avoid bumping. Suction filter the caffeine using a small Hirsch funnel and petroleum
ether as a transfer/rinse solvent. A second crop of caffeine may form in the filtrate as the
solvent evaporates. This second crop can also be collected by vacuum filtration but keep
it separate from the first crop. After air drying, weigh each crop and record the total
caffeine recovered from tea. Use the total caffeine recovered and the volume of filtrate
recovered to determine the total amount of caffeine that was dissolved in the original 200

27
mL of water. Assume that no water was lost during the 10 min boil, the extraction was
complete and that the caffeine was uniformly distributed in the original 200 mL water.
What % of of tea by weight is caffeine?

The sublimation will be performed as described by your instructor. The isolated


caffeine from several students will be combined and sublimed. The sublimed material
will then be redistributed to the students. Since caffeine is quite volatile, save all the
caffeine in sealed vials. You will use some of this material for TLC analysis next week.
You will not take a mp of the purified caffeine which would require a sealed capillary to
prevent sublimation near the melting point.

Preparation of Caffeine Salicylate (Optional). Caffeine is a base which can react with
acids to form salts. A well characterized salt of caffeine is caffeine salicylate formed by
using salicylic acid. This derivative of caffeine has a well-defined melting point which
can be used to confirm that the substance you have isolated from tea is indeed caffeine.
Later this semester, you will be required to make solid derivatives of other compounds
for similar purposes.

Procedure. Using an analytical balance (there are several top loader balances in the lab
across the hall which will quickly weigh to 0.0001 mg), weigh 50 mg of caffeine and 37
mg of salicylic acid (both can be plus or minus 1-2 mg) and dissolve them in 4 mL of
toluene in a small 25 mL Erlenmeyer flask by warming on a steam bath. Add 1 mL
(dropwise) of petroleum ether and allow the mixture to cool and crystallize. If necessary,
cool in an ice-water bath. Collect the crystals by vacuum filtration, air dry, weigh, record
the yield, and take a mp (lit mp 137 °C).

Sublimation of Caffeine. Sublimation is a process by which a solid is transferred from


one location to another by virtue of the fact that both solids (and liquids) have vapor
pressures that can become appreciable at elevated temperatures. If the vapor in
equilibrium with a solid at a higher temperature hits a surface that is much cooler, the
vapor will condense on the cooler surface because the equilibrium vapor pressure of the
substance off of the surface will be exceeded. This can lead to a transfer of solid from a
hotter zone to a cooler zone. Since different substances exhibit different vapor pressures
at any given temperature, it is possible to separate compounds by sublimation on the
basis of these differences in vapor pressure. Caffeine for example is much more volatile
than chlorophyll. It is possible to separate these compounds by sublimation.

caffeine

Apparatus. The sublimation procedure we will use for this experiment


will be demonstrated in class.
28
EXPERIMENT 4. THIN LAYER
CHROMATOGRAPHY (TLC)
In this experiment you will be asked to examine the TLC behavior a series of
commercially available analgesics. You will compare their elution behavior with the
behavior of the materials you have either synthesized, isolated in this laboratory or are
given. In addition you will be given an unknown analgesic and will be asked to identify
the active ingredient(s) in the product. The ingredients you will be analyzing are
acetylsalicylic acid, acetaminophen, iburofen and caffeine (not an analgesic). Two
relatively new drugs are naproxene and ketoprofen. These two materials are often used to
relieve headaches, backaches, menstrual cramps, and fever. You will not be testing for

O O
COOH NH CCH3 NH CCH3

O CCH3 CH3CH2O HO
O
Aspirin Phenacetin Aetaminophen

CH3 O
O CH3
CH3 CHCO2H C NH
N N 2

N CH3CHCH2
O N OH
CH3
CH3

Caffeine Ibuprofen Salicylamide

CH3 O CH3
CHCO2H CHCO2H

CH3O

Naproxen Ketoprofen
these two materials.

29
In addition, most commercial drug samples also contain a binder. Frequently this
binder is starch which is very polar and not very soluble in the solvents you will be using
in the chromatography. The binder will not move up the plate much in these experiments.

Principles of Thin Layer Chromatography (TLC)

Rf = the ratio of the distance traveled by the solute to that traveled by the solvent; always
<1.

Chromatography is based on some very simple principles. Most useful solid


adsorbents are characterized by several distinct properties which include large irregular
surface areas and polar sites at various locations on the surface. It is the polar sites that
are responsible for adsorbing solutes as they pass by. Depending on the polarity of the
solute and adsorbent, some solutes are adsorbed more strongly than others. Adsorption is
a competitive phenomenon. The solvent as well as the solute is in competition for these
active sites. If the solvent used is too polar, it will compete for these sites and
preferentially bind to them. The net result will be that the solutes will elute too quickly
and will not separate Rf>0.7. Alternatively if the solvent is not sufficiently polar, the
solutes will bind to these sites and the solvent will not be able to pull them off the
adsorbent. This leads to low Rf’s (>0.2). The challenge is to find a solvent system which
is adsorbed sufficiently to compete for the sites but is not too polar to overwhelm the
solute in the competition. A good way to wash off all solutes off the adsorbent is to
change the eluting solvent to a polar solvent such as methanol or acetic acid.

Relative Polarity of Some Solvents (in terms of decreasing polarity)

water
acids (acetic acid, propionic acid)
amides (N,N-dimethylformamide)
alcohols (methanol, ethanol)
amines (pyridine, trimethylamine)
ketones (acetone, 2-butanone, diethylketone)
esters (methyl acetate, ethyl acetate)
halides (chloroform, methylene chloride, carbon tetrachloride)
ethers (tetrahydrofuran, dioxane, diethyl ether)
aromatic hydrocarbons (benzene, toluene, xylenes)
alkanes (hexane, heptane)

Typical Adsorbents

The typical kinds of adsorbents used in an organic laboratory include paper and related
materials (cellulose, starch, etc.) alumina (aluminum oxide), silica gel (and related silicon
oxides), Florisil® and activated charcoal. There are many grades of these supports
available for chromatography and the proper choice of adsorbent and solvent system is
critical for success. Paper, silica gel and alumina are commonly used in tlc. Typical ratios
of adsorbent to solute used varies anywhere from 30/1 to 100/1 (by weight).

30
Visualization of the Solute

Since most organic compounds are colorless, the ability to detect the presence of small
amounts of solute and the distance they have eluted is an essential component in
successfully using TLC. A number of very sensitive methods are routinely used for such
a purpose. Both destructive and non-destructive methods are available. Most non-
destructive methods usually rely on the use of visible and ultraviolet light to determining
the location of the solute. The process by which the solute is visualized is complex and
beyond the scope of this treatment. However, all organic molecules absorb light in the
ultraviolet or vacuum ultraviolet region and their presence can usually be detected either

31
as a result of differential absorption, fluorescence, or phosphorescence of either the solute
or of a fluorescence probe that can be attached to the solid adsorbent. Destructive
methods include the use of an iodine chamber, and rely on the adsorption of iodine by the
adsorbent. The presence of the solute usually results in a differential color change either
as a result of reaction of the solute with iodine or as a result of its presence on the
adsorbent. Concentrated sulfuric acid usually reacts with most organic compounds often
resulting in color changes. A fine mist of this acid sprayed on the plates is another
destructive method of visualizing a developed TLC plate.

Procedure

You will be using prepared TLC plates which are research quality. Each large TLC sheet
will be carefully cut to provide mini TLC plates. Do not touch the adsorbent with your
fingers, hold the plates at the sides. Do not use ink on the plates, write in pencil lightly.

For a developing chamber, you will use your 4 oz. wide-mouth jar with a piece of
filter paper ca. 3 inches in diameter (trim if necessary). The half TLC plate you receive
can be cut width-wise into thirds to make three mini-TLC plates about 4 cm wide. Cut
evenly and carefully using sharp scissors. These mini TLC plates will fit into the 4oz.
jars. Ethyl acetate, 2.5-3 mL, should be used as the developing solvent to obtain the
proper depth with a baseline 0.5 cm above the bottom of the mini TLC plate. The solvent
front should be allowed to run as close to the top of the plate as possible but not to the
top, stop at least 5 mm from the top. These plates take about 5 min. to develop.

Prepare at least 10 micropipettes for spotting (the procedure will be demonstrated


in class).

A) Investigation of the impact of the solvent polarity on the Rf value

1. Prepare samples of aspirin, acetaminophen, and ibuprofen by crushing a pill of each


using a mortar and pistil. Dissolve each of the resulting powders in 1 mL of methylene
chloride in a stoppered vial. The samples might not be completely soluble; let the solids
decant to the bottom of the vial and work with the solution.
2. Practice your spotting technique on a sheet of filter paper until you are satisfied that
you can apply narrow diameter samples and that your pipettes are reliable.
3. Draw a very light pencil line at 0.5 cm from the bottom of your mini TLC plate (it is
very important that the pencil does not gouge out the adsorbent).
4. Spot the three samples on the line in order and code them in pencil at the top of the
TLC plate.
5. Observe your spotted samples under ultraviolet light to see if they are in appropriate
positions on the starting line of the TLC plate.
6. Carefully place the TLC plate into three different developing chambers so that the
plate sits almost perpendicular in the chamber and the bottom of the plate is nearly flat in
the developing solvent.
One developing chamber contains pure hexanes, the second one a 1:1 hexanes /
acetone mixture and the third one pure acetone.

32
Don’t let the plate touch the filter paper. Close the developing chambers carefully
without disturbing your TLC plate. The solvent front should be rising along the TLC
plate evenly.
7. When the solvent fronts are 0.5 cm from the top, remove the TLC plates from the
developing tanks and immediately mark the solvent front on the plates with a pencil. Let
the TLC plates dry in the hood then observe them under UV light. Circle the spots with a
pencil. Measure and record the distances each of the solutes and the solvent have
traveled.
8. Visualize the TLC plate in an iodine chamber. Record any differences you observe.

Discuss in the laboratory notebook the impact of the solvent polarity on the Rf
values.

B) Determination of the purity of your Aspirin and Acetaminophene samples

Following the procedure under A), spot your aspirin and acetaminophene samples on one
TLC plate and code them in pencil at the top of the TLC plate (prepare solutions of your
samples by dissolving 10 mg of the sample in 1 mL CH2Cl2). Immerse the TLC plate into
a developing chamber containing a 1:1 mixture of acetone / hexanes as a solvent as
described above.

Discuss in the laboratory notebook the purity of your sample. A pure sample should
exhibit only one spot on the TLC plate.

C) Determination of an unknown by TLC

1. Prepare samples of your aspirin, acetaminophen, and caffeine by dissolving ca. 10 mg


of compound in 1.0 mL of 1:1 methylene chloride/ethanol in a stoppered vial.
2. Practice your spotting technique on a sheet of filter paper until you are satisfied that
you can apply narrow diameter samples and that your pipettes are reliable.
3. Draw a very light pencil line at 0.5 cm from the bottom of your mini TLC plate (it is
very important that the pencil does not gouge out the adsorbent).
4. Spot samples on the line in order and code them in pencil at the top of the TLC plate.
Include a sample of the unknown you have chosen.
5. Observe your spotted samples under ultraviolet light to see if they are in appropriate
positions on the starting line of the TLC plate.
6. Carefully place the TLC plate into the developing chamber so that the plate almost
perpendicular in the chamber and the bottom of the plate is nearly flat in the ethyl acetate
developing solvent. Don’t let the plate touch the filter paper. Close the developing
chamber carefully without disturbing your TLC plate. The solvent front should be rising
along the TLC plate evenly.
7. When the solvent front is 0.5 cm from the top, remove the TLC plate from the
developing tank and immediately mark the solvent front on the plate with a pencil. Let
the TLC plate dry in the hood then observe it under UV light. Circle the spots with a
pencil. Measure and record the distances each of the solutes and the solvent have
traveled.

33
8. Visualize the TLC plate in an iodine chamber. Record any differences you observe.

The Rf value for each compound is unique but does depend on the conditions used in the
experiment. A number of variables affect chromatographic Rf values. Therefore, relative
rather than absolute Rf values are used to identify compounds. It is usual to perform TLC
experiments by spotting both knowns and unknowns on the same TLC plate. This
guarantees that many of these variables will remain constant during the course of the
experiment. Compounds with the same Rf values are likely to be the same compound.
Obtaining the same Rf value when performing the experiment using different solvents is
usually accepted as a valid criteria of identity. However, in order for these experiments to
be valid, the two compounds must exhibit Rf values between 0.3 and 0.7. Rf values below
0.3 usually implies that the solvent used is not polar enough while values in excess of 0.7
implies that the solvent is too polar for the adsorbent used.

Two Dimensional Thin Layer Chromatography

In some instances you may find that one solvent system is effective in separating two or
more components but that it is not capable of separating all of them. Another solvent
system may also result in partial separations. An effective way of combining both partial
separations is to perform the tlc in two dimensions. Select a square TLC plate and spot
your mixture so it is situated in the lower right hand corner of the developing chamber.
Develop the TLC plate in one of the solvent systems. Remove the plate from the
chamber, allow the solvent(s) to evaporate and place the plate in a second chamber with a
different solvent system. The plate should be oriented in such a manner that the original
spot should now be located in the lower left hand corner of the developing chamber.
Develop the plate a second time in a similar manner. The resulting plate should now have
each of solutes located along the entire surface of the plate.

Uses of Thin Layer Chromatography


Thin layer chromatography has many uses. Since the analysis takes only a few minutes to
perform, it is easy to follow the course of a reaction using TLC. It is frequently used to
confirm identity or the presence of a particular substance in a mixture. Although TLC is
used primarily as a qualitative test, it has been used as a criterion for purity.

Other Chromatographic Techniques


Thin layer chromatography is only one example of the different techniques used to
separate and purify organic compounds. Other chromatographic techniques include gas
chromatography which will be used later in this course, column chromatography, and
high performance liquid chromatography (HPLC). HPLC is covered in more detail in
analytical chemistry. While all of these techniques are quite different in practice, the
separations achieved by these techniques are based on very similar principles to those
discussed above for this layer chromatography.

34
EXPERIMENT 5. SYNTHESIS OF
ISOPENTYL ACETATE (BANANA OIL)
O O
H2SO4
CH3C O H + HOCH2CH2CH(CH3)2 CH3C O CH2CH2CH(CH3)2 + H2O

The reaction of a carboxylic acid and an alcohol produces an ester and water. The
reaction is catalyzed with acid. Esters often have a fruity taste or odor. Octyl acetate,
prepared using octanol as the alcohol, will remind you of oranges. You will prepare a
compound that smells like bananas.

Procedure:
1. Place 15 mL (12.2 g, 0.138 mole) of isopentyl alcohol in a 100-mL round-bottom
flask and add 20 mL (21 g, 0.35 mole) of glacial acetic acid.
2. Swirl the flask and carefully add 4 mL of conc. H2SO4 (CAUTION: highly
corrosive).
3. Attach a reflux condenser as shown in Figure 2 and, using a heating mantel,
reflux the mixture for 1 hour (do not forget to add a boiling stone). Cool to room
temperature.
4. Place the reaction mixture in a separatory funnel and add 55 mL of cold water
(remember to rinse the reaction flask with 10 mL of cold water and add it to the
separatory funnel). Separate the lower aqueous layer.
5. Extract the organic layer (upper layer) with 25 mL of 5% sodium bicarbonate
solution twice (test to be certain that the aqueous layer is basic to pH paper;
otherwise wash again). CAUTION: sodium bicarbonate will react with acetic acid
resulting in formation of carbon dioxide. Be sure to keep your separatory funnel
unstoppered.
6. Extract the organic layer with 25 mL of water. Finally, add 5 mL of saturated
aqueous sodium chloride to aid in layer separation (it removes traces of water
from the organic layer). Do not shake this solution but simply swirl. Draw off the
lower aqueous layer. Pour the top organic layer into an Erlenmeyer flask and dry
with ~2 g of anhydrous magnesium sulfate.
7. Decant the ester into the distilling flask (make sure that the drying agent is
excluded and all glassware completely dry). Set up the distillation apparatus as
described by your instructor. Be certain that the adapter is open to the atmosphere
and that your thermometer is placed correctly in the distilling head (do not forget
the boiling stone). Collect all distilled material but collect the fraction between
134 and 143 °C in a separate weighed flask. Keep the receiver flask cold to
reduce the vapor escaping into the lab environment. Never distill to dryness.
Record the barometric pressure in the laboratory.
8. Weigh the product and calculate the percentage yield.
9. Obtain an infrared spectrum of your sample.

Discussion

35
The yield in this experiment could be improved significantly by using a chaser solvent.
Every time you transfer you ester from one flask to another, you lose the material that
adheres to the wall of the container and this can be significant for liquids. A simple way
of avoiding this problem is to use a chaser solvent, some solvent that is easily separated
from your ester (by distillation). Washing the glassware with this solvent can allow
quantitative transfer of your ester. The only drawback in using a chaser solvent is the
extra step necessary to remove it. Whether you use a chaser or not depends on the scale
of the synthesis and the cost of the reagents.

Questions:

1. The specific gravity of glacial acetic acid is 1.05 g/mL. Ten milliliters of glacial acetic acid would
provide you with how many moles of acid? How many millimoles (mmol)?
2. What is the purpose of extracting the organic layer with 5% aqueous sodium bicarbonate?
Remember, you used an excess of glacial acetic acid.
3. What is the purpose of extracting with 25 mL of water?
4. Why is it necessary to record the barometric pressure when distilling at atmospheric
pressure?
5. The size of the distilling apparatus affects the amount of material that cannot be distilled (the
holdup). Assume that you could have used a distilling flask with a volume of 0.05 L or one with a
volume of 0.10 L. Using the ideal gas law, PV = nRT, calculate the number of moles of vapor
remaining in the flask at 127oC at 1 atm pressure for both cases.
6. Can you name another ester you synthesized this semester? What was the “alcohol” used?

36
Water
out

Clamp

Water
in

Clamp

Heat
source to
be added

Figure 2. A setup for total reflux

37
DISTILLATION
Distillation is an important commercial process that is used in the purification of a large
variety of materials. However, before we begin a discussion of distillation, it would
probably be beneficial to define the terms that describe the process and related properties.
Many of these are terms that you are familiar with but the exact definitions may not be
known to you. Let us begin by describing the process by which a substance is
transformed from the condensed phase to the gas phase. For a liquid, this process is
called vaporization and for a solid it is called sublimation. Both processes require heat.
This is why even on a hot day at the beach, if there is a strong breeze blowing, it may feel
cool or cold after you come out of the water. The wind facilitates the evaporation process
and you supply some of the heat that is required. All substances regardless of whether
they are liquids or solids are characterized by a vapor pressure. The vapor pressure of a
pure substance is the pressure exerted by the substance against the external pressure
which is usually atmospheric pressure. Vapor pressure is a measure of the tendency of a
condensed substance to escape the condensed phase. The larger the vapor pressure, the
greater the tendency to escape. When the vapor pressure of a liquid substance reaches the
external pressure, the substance is observed to boil. If the external pressure is
atmospheric pressure, the temperature at which a pure substance boils is called the
normal boiling point. Solid substances are not characterized by similar phenomena as
boiling. They simply vaporize directly into the atmosphere. Many of you may have
noticed that even on a day in which the temperature stays below freezing, the volume of
snow and ice will appear to decrease, particularly from dark pavements on the streets.
This is a consequence of the process of sublimation. Both vaporization and sublimation
are processes that can be used to purify compounds. In order to understand how to take
advantage of these processes in purifying organic materials, we first need to learn how
pure compounds behave when they are vaporized or sublimed.

Let's begin by discussing the vapor pressure of a pure substance and how it varies
with temperature. Vapor pressure is an equilibrium property. If we return to that hot
windy day at the beach and consider the relative humidity in the air, the cooling effect of
the wind would be most effective if the relative humidity was low. If the air contained a
great deal of water vapor, its cooling effect would be greatly diminished and if the
relative humidity was 100%, there would be no cooling effect. Everyone in St. Louis has
experienced how long it takes to dry off on a hot humid day. At equilibrium, the process
of vaporization is compensated by an equal amount of condensation. Incidentally, if
vaporization is an endothermic process (i.e. heat is absorbed), condensation must be an
exothermic process (i.e. heat is liberated). Now consider how vapor pressure varies with
temperature. Figure 3 illustrates that vapor pressure is a very sensitive function of
temperature. It does not increase linearly but in fact increases exponentially with
temperature. A useful "rule of thumb" is that the vapor pressure of a substance roughly
doubles for every increase in 10 °C. If we follow the temperature dependence of vapor
pressure for a substance like water left out in an open container, we would find that the
equilibrium vapor pressure of water would increase until it reached 1 atmosphere or
101325 Pa (101.3 kPa, 760 mm Hg). At this temperature and pressure, the water would
begin to boil and would continue to do so until all of the water distilled or boiled off. It is

38
not possible to achieve a vapor pressure greater than 1 atmosphere in a container left
open to the atmosphere. Of course, if we put a lid on the container, the vapor pressure of
water or any other substance for that matter would continue to rise with

Vapor Pressure vs Temperature of Water

250

200
Vapor Pressure (kPa)

150

100

50

0 20 40 60 80 100 120 140


Temperature, °C

Figure 3. Vapor pressure dependence on temperature for water.

temperature until the container ruptured. Elevation of the boiling point with increase in
external pressure is the principle behind the use of a pressure cooker.

Vacuum Distillation

Elevation of the boiling point with an increase in external pressure, while important in
cooking and sterilizing food or utensils, is less important in distillation. However, it
illustrates an important principle that is used in the distillation of many materials. If the
boiling point of water is increased when the external pressure is increased, then
decreasing the external pressure should decrease the boiling point. While this is not
particularly important for the purification of water, this principle is used in the process of
freeze drying, an important commercial process. In addition, many compounds cannot be
distilled at atmospheric pressure because their boiling points are so high. At their normal

39
boiling points, the compounds decompose. Some of these materials can be distilled under
reduced pressure however, because the required temperature to boil the substance can be
lowered significantly. Rewording the "rule of thumb" described above so that it is
applicable here suggests that the boiling point will be lowered by 10 °C each time the
external pressure is halved. For example, if the external pressure above a substance is
reduced to 1/16 of an atmosphere by mean of a mechanical pump, the boiling point will
have been reduced four times by 10 °C for a total reduction of 40 °C (1 atm ×
(1/2)(1/2)(1/2)(1/2) =1/16 atm).

A nomograph is a useful device that can be used to estimate the boiling point of a
liquid under reduced pressure under any conditions provide either the normal boiling
point or the

Pressure-Temperature Nomograph for Vacuum Distillations

Boiling Point Pressure (mm)


at pressure P Boiling Point 1 atm=760 mm;
(mm) at 1 atm
(760 mm)
°C °F

°C °F

(a) (b) (c)

Figure 4. A nomograph used to estimate boiling points at reduced pressures. To use, place a straight edge
on two of the three known properties and read out the third. Column c is in mm of mercury. An atmosphere
is also equivalent to 101.3 kPa and will support a column of mercury, 76 cm (760 mm).

boiling point at a some given pressure is available. To use the nomograph given the
normal boiling point, simply place a straight edge at on the temperature in the central
column of the nomograph (b). Rotating the straight edge about this temperature will

40
afford the expected boiling point for any number of external pressures. Simply read the
temperature and the corresponding pressure from where the straight edge intersects the
first and third columns. As an example lets choose a normal boiling point of 400 °C.
Using the nomograph in Figure 4 and this temperature for reference, rotating the straight
edge about this temperature will afford a continuous range of expected boiling points and
the required external pressures necessary to achieve the desired boiling point. At a
pressure of 6 mm, the expected boiling point would be 200 °C. Likewise, our compound
boiling at 400 °C at 1 atm would be expected to boil at 145 °C at 0.1 mm external
pressure.

Simple Distillation

Although all of us have brought water to a boil many times, some of us may have not
realized that the temperature of pure boiling water does not change as it distills. This is
why vigorous boiling does not cook food any faster than a slow gentle boil. The
observation that the boiling point of a pure material does not change during the course of
distillation is an important property of a pure material. The boiling point and boiling
point range have been used as criteria in confirming both the identity and purity of a
substance. For example, if we synthesized a known liquid that boiled at 120-122 °C, this
value could be used to confirm that we prepared what we were interested in and that our
substance was reasonably pure. Of course, additional criteria must also be satisfied before
the identity and purity of the liquid are known with certainty. In general, a boiling point
range of 1-2 °C is usually taken as an indication of a pure material. You will use both of
these properties later in the semester to identity an unknown liquid.

Occasionally, mixtures of liquids called azeotropes can be encountered that


mimic the boiling behavior of pure liquids. These mixtures when present at specific
concentrations usually distill at a constant boiling temperature and can not be separated
by distillation. Examples of such mixtures are 95% ethanol-5% water (bp 78.1 °C), 20%
acetone-80% chloroform (bp 64.7 °C), 74.1% benzene, 7.4% water, 18.5 % ethanol (bp
64.9 °C). The azeotropic composition sometimes boils lower the than boiling point of its
components and sometimes higher. Mixtures of these substances at compositions other
than those given above behave as mixtures.

Returning to our discussion of boiling water, if we were making a syrup by the


addition of sugar to boiling water, we would find that the boiling point of the syrup
would increase as the syrup begins to thicken and the sugar concentration becomes
significant. Unlike pure materials, the boiling point of an impure liquid will change and
this change is a reflection of the change in the composition of the liquid. In fact it is this
dependence of boiling point on composition that forms the basis of using distillation for
purifying liquids. We will begin our discussion of distillation by introducing Raoult's
Law, which treats liquids in a simple and ideal, but extremely useful manner.

41
Figure 5. The apparatus used in a simple distillation. Note the position of the thermometer bulb in the
distillation head and the arrangement of the flow of the cooling water.

PAobs =  PAo , where PAobs is the observed vapor pressure of A 1


 = nA/(nA + nB + nC +...)
PAo = vapor pressure of pure A
nA, nB, nC ... : number of moles of A, B, C, ...

This relationship as defined is capable of describing the boiling point behavior of


compound A in a mixture of compounds under a variety of different circumstances.
Although this equation treats mixtures of compounds in an oversimplified fashion and is
not applicable to azeotropic compositions, it does give a good representation of the
behavior of many mixtures.

Let's first consider a binary system (2 components) in which only one of the two
components is appreciably volatile. Our previous discussion of sugar + water is a good
example. Raoult's law states that the observed vapor pressure of water is simply equal to
the product of the mole fraction of the water present and the vapor pressure of pure water
at the temperature of the mixture. Once the sugar-water mixture starts to boil, and
continues to boil, we know that the observed vapor pressure of the water must equal one
atmosphere. Water is the only component that is distilling. Since the mole fraction of
water in a mixture of sugar-water must be less than 1, in order for the observed vapor
pressure of water ( PAobs ) to equal one atmosphere, PAo must be greater than one
atmosphere. This can only happen according to Figure 3 if the temperature of the mixture

42
is greater than 100 °C. As the distillation of water continues, the mole fraction of the
water continues to decrease thereby causing the temperature of the mixture to increase.
Remember, heat is constantly being added. If at some point the composition of the
solution becomes saturated with regards to sugar and the sugar begins to crystallize out of
solution, the composition of the solution will become constant; removal of any additional
water will simply result in the deposit of more sugar crystals. Under these set of
circumstances, the composition of the solution will remain constant and so will the
temperature of the solution although it will exceed 100 °C.

During the course of the distillation, the water vapor which distilled was initially
at the temperature of the solution. Suspending a thermometer above this solution will
record the temperature of the escaping vapor. As it departs from the solution, the
temperature of the vapor will cool by collisions with the surface of vessel until it reaches
100 °C. Cooling below this temperature will cause most of the vapor to condense to a
liquid. If cooling to 20 °C occurs in the condenser of a distillation apparatus, then by
using the appropriate geometry as shown in Figure 5, it would be possible to collect
nearly all of the liquid. The only vapor that would be lost to the environment would be
that small amount associated with the vapor pressure of water at 20 °C. Since the vapor
pressure of water at 20 °C is roughly 2.3 kPa, then 2.3/101.325 or 0.023 would be the
fraction of water that would not condense and would pass out of the condenser. This is
why the distillate is frequently chilled in an ice bath during the distillation.

The distillation of a volatile material from non-volatile is one practical use of


distillation which works very well. However, often there may be other components
present that although they may differ in relative volatility, are nevertheless volatile
themselves. Let's now consider the two component system you will be using in the
distillations you will perform in the laboratory, cyclohexane and methylcyclohexane. The
vapor pressures of these two materials in pure form are given in Table 1. As you can see
from this table, although cyclohexane is more volatile than methylcyclohexane, the
difference in volatility between the two at a given temperature is not very great. This
means that both materials will contribute substantially to the total vapor pressure
exhibited by the solution if the distillation is carried out at 1 atmosphere. The total
pressure, PT, exerted by the solution against the atmosphere according to Dalton's Law of
partial pressures, equation 2, is simply the sum of the observed vapor pressures of
cyclohexane, Pcobs , and methylcyclohexane, Pmobs :
obs obs
PT = Pc + Pm . 2

As before, boiling will occur when the total pressure, PT, equals an atmosphere.
However since we have two components contributing to the total pressure, we need to
determine the

43
Table 1. Vapor pressures of cyclohexane and methylcyclohexane as a function of temperature.

cyclohexane methylcyclohexane

T/K P/kPa T/K P/kPa

300 14.1 300 6.7


305 17.6 305 8.5
310 21.7 310 10.6
315 26.5 315 13.2
320 32.2 320 16.2
325 38.8 325 19.8
330 46.5 330 24.0
335 55.3 335 28.9
340 65.4 340 34.6
345 77.0 345 41.2
350 90.0 350 48.7
354 101.3 354 55.4
360 121.3 360 66.9
362 128.5 362 71.1
365 139.9 365 77.9
370 160.5 370 90.2
373 174.0 373 101.3
380 208.8 380 119.3
385 236.7 385 136.4
390 267.3 390 155.3
395 300.9 395 176.2
400 337.5 400 199.1

relative contributions of each. Again we can use Raoult's Law but we need more
information about the system before we can do so. In particular we need to know the
composition of the solution of cyclohexane and methylcyclohexane. For ease of
calculation, let's assume that our original solution has equal molar amounts of the two
components. What we would like to determine is whether it would be possible to
separate cyclohexane from methylcyclohexane by distillation. By separation, we would
like to determine if it would be possible to end up with two receiver flasks at the end of
the experiment that would contain mainly cyclohexane in one and mainly
methylcyclohexane in the other. It is clear that at some point we will need to intervene
in this experiment. Otherwise, if we were to collect the entire contents of the original
distilling flask, called the pot, into one receiver flask, we would end up with the same
composition as we started. Initially the mole fractions of both cyclohexane and
methylcyclohexane are 0.5. From Raoult's Law (equation 1), Dalton's Law (equation 2)
and the information in Table 1, we can estimate that boiling will occur at
approximately 362 K when the total pressure of the two components equals one
atmosphere or 101.3 kPa.:
obs obs
PT = Pc + Pm = 0.5(128.5 kPa) + 0.5(71.1 kPa)  101.3 kPa

The first thing that we should note is that the initial boiling point is higher than the lowest
boiling component and lower than the highest boiling component. Next, we should
inquire about the composition of the vapor. Is the composition of the vapor the same as

44
the initial composition of the pot or is it enriched in the more volatile component? If the
composition of the vapor is the same as that of the original mixture, then distillation will
not be successful in separating the two components. However, we should ask, "What is
the composition of the vapor?" Before the vapor is cooled and condenses on the
condenser, we can treat the vapor as an ideal gas. Recalling that: PV = nRT, where P is
the pressure of the gas or vapor, V is the volume it occupies, n is the number of moles of
-1 -1
gas, R is the Gas Constant (0.0821 L. atm.K .mol ) and T is the temperature, we can
determine the composition of the vapor by taking advantage of the following factors.
First we note that:

PTV = nTRT so that Pcobs V = ncRT and Pmobs V = nmRT where nT refers to the total number
obs obs
of moles, since ( Pc + Pm ) = (nc + nm)RT.

If the total vapor can be treated as an ideal gas, then according to Dalton's Law, so can
each of the components. Since the two components are in thermal contact and are
distilling together, we can expect them to be at the same temperature. We don't
necessarily know the volume of the container, but since it is assumed that the volumes of
the molecules are very small in comparison to the total volume the gas occupies,
whatever the value of V, it is the same for both components. This means we can establish
the following ratio:

Pcobs / Pmobs = nc/nm

which in turn allows us to determine the composition of the vapor from the observed
partial pressures of the two components. If we use the experimental values found in Table
1, we conclude that the composition of the vapor is 1.8/1, and is indeed enriched in the
more volatile component.

This simple treatment allows us to understand the principles behind distillation.


However it is important to point out that distillation is far more complex than our simple
calculation indicates. For example, we just calculated the composition of the vapor as
soon as the solution begins to boil and we have correctly determined that the vapor will
be enriched in the more volatile component. This means that as the distillation proceeds,
the pot will be enriched in the less volatile component. Since the composition of the pot
will change from the initial 1:1 mole ratio and become enriched in the less volatile
component; the new composition in the pot will introduce changes in the composition of
the vapor. The composition of the vapor will also change from the initial ratio we just
calculated to a new ratio to reflect the new composition of the pot. The consequences of
these changes are that the temperature of both the pot and the distillate will slowly
increase from the initial value to a value approaching the boiling point and composition
of the less volatile component. If we are interested in separating our mixture into
components, we are left with the task of deciding how much material to collect in each
receiver and how many receivers to use. Obviously this will depend on the quality of
separation we are interested in achieving. Generally, the more receivers we use, the less
material we will have in each. It is possible to combine fractions that differ very little in
composition but this requires us to analyze each mixture. While it is possible to do this,

45
in general, we really want to end with three receivers, one each enriched in the two
components of our mixture and a third that contains a composition close to the initial
composition.

It is difficult to describe how much material to collect in each receiver since the
volume collected will depend on the differences in the boiling points of the components.
As a general rule, the receiver should be changed for every 10 °C change in boiling point.
Each fraction collected can be analyzed and those with compositions similar to the initial
composition can be combined. The main fractions collected can then be fractionated a
second time if necessary.

The experiment we have just discussed is called a simple distillation. It is an


experiment that involves a single equilibration between the liquid and vapor. This
distillation is referred to as involving one theoretical plate. As you will see, it is possible
to design more efficient distillation columns that provide separations on the basis of
many theoretical plates. Before discussing these columns and the advantages offered by
such fractionating columns, it is important to understand the basis of the advantages
offered by columns with many theoretical plates. The following is a simplified discussion
of the process just described involving a column with more than one theoretical plate.

Fractional Distillation

We have just seen that starting with a composition of 1:1, cyclohexane:


methylcyclohexane, the composition of the vapor was enriched in the more volatile
component. Suppose we were to collect and condense the vapor and then allow the
resulting liquid to reach equilibrium with its vapor. Let’s call this liquid, liquid 2. The
properties of liquid 2 will differ from the original composition in two ways. First, since
the composition of liquid 2 is higher in cyclohexane than the initial one; the temperature
at which liquid 2 will boil will be lower than before (what is the approximate boiling
point of a 1.8/1 mixture of cyclohexane/methylcyclohexane? see Table 1). In addition,
the composition of the vapor, vapor 2, in equilibrium with liquid 2 will again be enriched
in the more volatile component. This is exactly what happened in the first equilibration
(first theoretical plate) and this process will be repeated with each new equilibration. If
this process is repeated many times, the vapor will approach the composition of the most
volatile component, in this case pure cyclohexane, and the liquid in the pot will begin to
approach the composition of the less volatile component, methylcyclohexane. In order for
this distillation to be successful, it is important to allow the condensed liquid which is
enriched in the less volatile component relative to its vapor, to return to the pot. In a
fractional distillation, the best separation is achieved when the system is kept as close to
equilibrium as possible. This means that the cyclohexane should be removed from the
distillation apparatus very slowly. Most fractional distillation apparatus is designed in
such a way as to permit control of the amount of distillate that is removed from the
system. Initially the apparatus is set up for total reflux, (i.e. all the distillate is returned
back to the system). Once the distillation system reaches equilibrium, a reflux to takeoff
ratio of about 100:1 is often used (about 1 out of every 100 drops reaching the condenser
is collected in the receiver).

46
A column which allows for multiple equilibrations is called a fractionating
column and the process is called fractional distillation. An example of a fractionating
column is shown in Figure 6. Each theoretical plate is easy to visualize in this column.
The column contains a total of 4 theoretical plates and including the first equilibration
between the pot and chamber 1 accounts for a total of 5 from pot to receiver. As you
might expect, a problem with this column is the amount of liquid that is retained by the
column. Many other column designs have been developed that offer the advantages of
multiple theoretical plates with low solvent retention. Typical spinning band columns
often used in research laboratories offer fractionating capabilities in the thousands of
theoretical plates with solute retention of less than one mL. Commercial distillation
columns have been designed for gasoline refineries that are multiple stories high and are
capable of separating compounds with boiling points that differ by only a few degrees.

In addition to performing a fractional distillation at one atmosphere pressure, it is


also possible to conduct fractional distillations at other pressures. This is often avoided
when possible because of the increased difficulty and expense in maintaining the vacuum
system leak free.

Figure 6. A fractionating column which contains four chambers, each with a center opening into the
chamber directly above. The vapor entering the first chamber cools slightly and condenses, filling the
lower chamber with liquid. At equilibrium, all chambers are filled with distillate. A portion of the liquid
condensing in the first chamber is allowed to return to the pot. The remaining liquid will volatilize and
travel up the column condensing in the second chamber and so on. As discussed in the text, the
composition of the vapor at each equilibration is enriched in the more volatile component. The heat
necessary to volatilize the liquid in each chamber is obtained from the heat released from the condensing
vapors replacing the liquid that has been removed. The vacuum jacket that surrounds the column ensures a
minimum of heat loss.

47
EXPERIMENT 6. FRACTIONAL AND
SIMPLE DISTILLATION OF A BINARY
MIXTURE
In this experiment, you will be conducting both a simple distillation and a fractional
distillation of a mixture of cyclohexane (bp = 79 ºC) and methylcyclohexane (bp = 101
ºC) to compare the two techniques. The initial mixture will be approx. 1:1
cyclohexane/methylcyclohexane (v/v), your instructor will give you the actual amounts
the day of the experiment.
To maximize the efficiency of the fractional distilling column you will want to
control the rate at which the material distills. To assess the efficiency of the column you
will analyze the distillate using gas chromatography (GC). You will compare these
results with those from the simple distillation.
For this experiment you will work with a partner. Each student pair will be
responsible for collecting a total of 6 ml of distillate in 2 ml increments from both simple
and fractional distillation setups--3 samples from each. You will be required to record the
following data for each fraction in your notebook and on the blackboard:

1. Distillation rate measured for each fraction (ml collected/minute)


2. Temperature range during distillation of each fraction
3. Composition of the distillate (based on GC results).

After collecting your fractions, you will analyze them by gas chromatography (GC). In
your notebook, record the chromatography conditions used for the analyses. You will
also record the data reported by the other groups of students (which will be provided) and
with this data you will construct two graphs for each type of distillation. The first graph
will be a determination of temperature vs fraction number (or volume); the second graph
will be an analysis of composition of each component versus fraction number (a total of
four graphs).

You should discuss your data in writing, especially any discontinuities in the
graphs. What parameters affected the efficiency of your distillation? How might it have
been improved?

Students will sign up at the blackboard to take fractions in the following order:

Fraction # Technician
1-3 Smith, Jones
4-6
7-9
10-12
13-15
16-18
19-21

48
(You will be taking fractions from both setups simultaneously; be certain to label the
fractions immediately so they are not confused; e.g., SD1, SD2, SD3, FD1, FD2, FD3).
Observe the activity of the students ahead of you to become familiar with
conditions, behavior of the column, etc. Record your data on the sheet (or chalkboard) in
the lab. Tightly cap your samples and, if you need to keep them until next week, wrap the
vial with Parafilm (ask the TA).

The two distillation apparatus will be set up as shown in Figures 5 and 6. In this
series of experiments we will use a fractionation column called a Vigreux column. There
are no additional setups for this experiment.

The composition of each mixture will be analyzed by gas chromatography

Gas Chromatography

A gas chromatograph is a very useful instrument that is used in the separation and
analysis of complex mixtures. The instrument consists of three major components. These
are the injector, column and detector. A fine capillary tube that contains a flowing inert
gas, usually He, is used. This capillary first passes through the injector compartment
where the sample is injected into the capillary and immediately volatilized by rapid
heating. The injected solute is then carried on to the capillary column by the carrier gas
and cooled. The interior of the capillary column contains various organic groups
chemically bound to the glass column. These groups selectively adsorb the solute such
that an "equilibrium" is established between the solute adsorbed on the column and that
in the gas phase. As a gas, the solute moves with the carrier; adsorbed on the column, its
motion is retarded. The retention time or time it takes a particular solute to emerge from
the column depends on two factors, its volatility (vapor pressure) and solubility on the
stationary phase of the column (the various organic groups bound on the column). For a
given flow rate, different materials elute differently, depending on these two factors.
Once the solute emerges from the column, it passes into the detector where a signal is
generated. The detector we will be using is called a flame ionization detector or fid. This
detector burns the sample; its output is proportional to the amount of sample burned over
a significant range of sample size. The detector is remarkably sensitive. Typical analyses
require sample sizes in the microgram range. The output produced by these ions is
amplified and sent to an integrator.

The instrument will be set up for your analysis. You will inject your samples
according to the technique demonstrated in the laboratory. The analysis of your results
provided by the integrator should be then made available to your colleagues.

49
EXPERIMENT 7. ISOLATION OF A
NATURAL PRODUCT BY STEAM
DISTILLATION
The concentration and isolation of an essential oil from a natural product has had a
dramatic impact on the development of medicine and food chemistry. The ability to
characterize the structure of the active ingredient from a natural product has permitted
synthesis of this material from other chemicals, resulting in a reliable and often cheaper
source of the essential oil. The process often used in this isolation is called steam
distillation. Steam distillation is an important technique that has significant commercial
applications. Many compounds, both solids and liquids, are separated from otherwise
complex mixtures by taking advantage of their volatility in steam. A compound must
satisfy three conditions to be successfully separated by steam distillation. It must be
stable and relatively insoluble in boiling water, and it must have a vapor pressure in
boiling water that is of the order of 1 kPa (0.01) atmosphere. If two or more compounds
satisfy these three conditions, they will generally not separate from each other but will be
separated from everything else. The following example, expressed as a problem,
illustrates the application of steam distillation:

Suppose we have 1 g of an organic compound present in 100 g of plant material


composed mainly of macromolecular material such cellulose and related substances. Let's
assume that the volatile organic material has a molecular weight of 150 Daltons, a vapor
pressure of 1 kPa and is not soluble in water to an appreciable extent. Examples of such
materials characterized by these properties include eugenol from cloves, cinnamaldehyde
from cinnamon bark or cuminaldehyde from cumin seeds. How much water must we
collect to be assured we have isolated all of the natural oil from the bulk of the remaining
material?

We can simplify this problem by pointing out that the organic material is not
appreciably soluble in water. We know from previous discussions that boiling will occur
when the total pressure of our system equals atmospheric pressure. We can also simplify
the problem by assuming that the essential oil in not appreciably soluble in the
macromolecular material. While in reality this does not have to be correct, this
assumption simplifies our calculation.
obs obs o o
PT = Pwater + Porg ; PT = water Pwater + org Porg

Boiling of our mixture will occur close to 100°C. Remember that very little oil is
soluble in water which makes the mole fraction of water near unity. Similarly for the our
volatile oil, its mole fraction is also close to one according to our assumption. The total
pressure, PT, is the sum of the vapor pressure of water, 100 kPa, and the essential oil,
o
Porg , 1 kPa. Boiling will occur very close to the boiling point of pure water. Treating the
water vapor and the organic vapor which are miscible as ideal, the PV ratio for both
vapors is given by the following:

50
PwaterV/ PorgV = nwaterRT/norgRT;
Pwater/ Porg = nwater/norg and
nwater = wtwater/18; norg = 1/150; rearranging:
wtwater = (100/1)(18/150) = 12 g water or 12 mL

Our calculation suggests that we can be assured that most of the 1 g of the organic
matter has been transferred by the steam if we condense and collect 120 mL of water.
The basis of the separation by steam distillation is that while the water and organic
condensed phases are immiscible, the vapors of both are miscible. Once condensed, the
two separate again allowing for an easy separation. As noted above, both liquids and
solids can be distilled by steam.

In this experiment you will isolate a natural product by steam distillation and
characterize it by its infrared spectrum. The spice you will be given will contain one of
these products shown below. You will be asked to identify which of the compounds you
isolated by examining the infrared spectrum of the oil you isolate.

CH3 CH3
CO2CH3 CHO

OH CH3CH HO
CH3 CHCH3 CH3
CH3 CH2

CH2 CH3
O
HO
OCH3

CH2 CH3

Procedure

Weigh approximately 5.0 g of spice in a 250 mL round bottom flask and add 150 mL of
water. Assemble a simple distillation apparatus as shown in Figure 3 on page 35. Careful
heating is important or frothing may be a problem. Distill about 50 mL of liquid. The
distillate may be cloudy. Allow the mixture in your round bottom flask to cool and
suction filter the residue. The spice should be discarded as solid trash.

The distillate will be extracted with methylene chloride. Add 10 mL of methylene


chloride to the distillate and transfer the contents to a separatory funnel. Gently swirl to
achieve good agitation and separate the lower layer (CH2Cl2 layer). Repeat the extraction

51
again using a second 10 mL portion of methylene chloride. Combine the methylene
chloride extracts, add ~2 g of anhydrous magnesium or sodium sulfate to the methylene
chloride and swirl until the liquid is no longer cloudy and there is no aqueous layer on
top. Filter off the drying agent (gravity) into a tared round bottom flask and remove the
methylene chloride on a rotary evaporator. Record the weight of the natural product
isolated and run an infrared spectrum on it. Identify the structure of your natural product
on the basis of its infrared spectrum and comparison of its gas chromatographic retention
time to an authentic sample.

EXPERIMENT 8. PREPARATION OF
METHYL SALICYLATE - VACUUM
DISTILLATION

In this experiment you will prepare methyl salicylate using a procedure similar to the one
used to prepare isoamyl acetate. The apparatus you will use is shown in Figure 2 on page
30. Weigh out 10 g of salicylic acid and combine it with 25 mL of anhydrous methanol
(density 0.8 g mL–1) and a boiling chip. Swirl the flask and warm the mixture to help
dissolve the acid. Carefully add 10 mL of concentrated sulfuric acid slowly. A precipitate
may form that will redissolve upon further heating. Reflux the mixture for an hour. Cool
the mixture and pour the contents onto ~50 g of ice. Add 25 mL of methylene chloride
and pour the contents into your separatory funnel after all the ice has melted. Separate the
layers and extract the upper layer once more with an additional 25 mL of methylene
chloride. Combine the methylene chloride layers. Empty the contents of the separatory
funnel in a beaker and return the methylene chloride layer to the separatory funnel. Wash
the methylene chloride layer with 25 mL of water, and then with 25 mL portions of 5%
aqueous sodium bicarbonate until there is no further evolution of carbon dioxide. In each
of these operations, the methylene chloride will be in the lower layer. Consequently it
will be necessary to drain the contents of the separatory each time and return the
methylene chloride back to the funnel.

Transfer the methylene chloride layer to a 125 mL Erlenmeyer flask and add a
few grams of anhydrous magnesium or sodium sulfate. Cork the mixture and swirl it to
dry the organic layer (this is a convenient place to stop if you’re running out of time).
Gravity filter the solution through a piece of fluted filter paper into a tared appropriately
sized round bottom flask (the flask should be less than half full). Remove the methylene
chloride using a rotary evaporator. Record the weight of your crude ester by difference.

Vacuum Distillation

52
Methyl salicylate will be purified by vacuum distillation. In this part of the experiment,
crude ester from three students will be combined and distilled at on time. The apparatus
that will be used will be set up in the laboratory for you. The apparatus that will be used
is illustrated in Figure 5 (page 35). The major difference between the apparatus shown in
this figure and the one you will be using is that a vacuum pump will be connected to it.
Once the vacuum distillation is complete, you will be required to run an infrared
spectrum of the recovered material.

53
EXPERIMENT 9. IDENTIFICATION OF
UNKNOWNS
Unknowns:
Each of you will receive two vials: one containing a liquid unknown and one a solid
unknown. You will perform a number of experimental procedures on these compounds to
gather data. The compounds are part of a finite number of compounds that are listed for
you in order of increasing mp and bp. You will determine the actual structure of your
unknowns by applying your experimental data to these lists, obtaining a shorter list of
possible compounds and performing further experiments to make the final determination.

Since each of you has a different compound, you will have to work more
independently than before and you will have to evaluate your data and the reliability of
your data.

Solubility tests: For these tests, you should use approx. 30 mg of your compound in 1
mL of solution to give you a 3% solution. The words soluble and insoluble are
qualitative, not quantitative. We will use 3% as a determination of solubility. If your
compound is not completely soluble as a 3% solution, you can call it insoluble or
partially soluble. Moreover, some compounds which are not soluble at room temperature,
may be soluble at higher temperature. Therefore, after obtaining data at room
temperature, heat your solution in a water bath at 70°C or higher and observe any further
changes. The results you record may have more meaning later when you know more
about your unknown compound. You will test the solubility/reactivity of your compound
in 4 different solutions.

Procedure: Into 4 labeled test tubes, add 1 mL of : water; 5% aq HCl; 5% aq NaHCO3;


and 5% aq. NaOH. To each tube, add approx. 30 mg of the unknown. Shake or stir for a
few seconds. Record your observations. It may take several minutes for the unknown to
appear to dissolve or react. After suitable time, heat the four test tubes in a water bath and
observe. Look for color changes, evolution of gas, any evidence of reaction such as
precipitates, and enhanced solubility. Record your observations. Repeat the solubility test
for the other unknown.

Most organic compounds are not water soluble. Only low molecular weight
and/or highly polar compounds will be water soluble to 3%. Salts of acids or bases will
often be soluble.
If the compound is not soluble in water, solubility in acid suggests that it is basic
(e.g., amines), solubility in strong base suggests that it is at least weakly acidic (e.g.,
phenols) and solubility in weak base (bicarbonate) suggests that it is a stronger acid (e.g.,
carboxylic acids).
Whatever determination you make from this solubility test, you will corroborate
that suggestion with other data including class tests and analysis of the compound's
infrared spectrum.

54
SOLUBILITY CLASSES
AA
ether insol
Litmus red
sol CA
Litmus blue
AM
Litmus unchanged
sol
N

water
SA
sol
5% NaHCO3
sol

insol WA
insol

5% NaOH
sol
B
5% HCl
insol
sol
N
96% H2SO4
insol

insol I

AA amino acids
CA carboxylic acids (5 carbons or less)
AM amines (5 carbons or less)
N alcohols, aldehydes, ketones, nitriles, amides (5 carbons or more)
SA carboxylic acids, phenols with strong electron withdrawing groups (5
carbons or more)
WA weak acids, phenols, enols, oximes, imides, 1,3-diones, nitro
compounds
B aliphatic amines, anilines
N alcohols, aldehydes esters, ketones, ethers, alkynes, akenes
I saturated hydrocarbons, haloalkanes, aryl halides

55
Compounds which are insoluble in all of the above liquids can be tested for solubility and
color change in concentrated sulfuric acid. Compounds giving a color change or showing
solubility include neutral compounds such as alkenes, alkynes, alcohols, ketones,
aldehydes, esters, ethers, amides and nitro compounds. Compounds that appear inert
include alkanes, alkyl halides, and simple aromatic hydrocarbons.

Acids are corrosive material and must be discarded in special containers, which
will be provided by the instructor.

Classification tests
All classification tests should be done in triplicate: known, unknown, and blank. The
known gives you a positive test for reference and tells you if the reagents or procedure
are faulty. The blank gives you the other extreme, a negative test. The results you obtain
from your unknown can then be put into perspective. The triplicate tests should be done
simultaneously thereby taking little extra time. For each test, use clean test tubes. For
certain tests, the presence of even trace amounts of acetone can give false positives.

Do only those tests that you think will be useful based upon solubility results
and your analysis of the infrared spectrum of the compound. If that analysis has not
been done and you wish to do some class tests, do them in the order shown below.

Beilstein test for halogenated compounds. This test is very sensitive. Take a piece of
copper wire with a loop on the end and heat it in the flame of a Bunsen burner until it
glows red. Allow the wire to cool but avoid contaminating it especially if you put it down
on the hood surface. Dip the cool loop into the known compound, and place it in the
flame. You should observe a green flame after the first few seconds when your known
compound burns (yellow flame). Burn off all remaining halogenated known compound,
heat the loop until it glows red, then let the loop cool and run the test on you unknown
solid or liquid. For knowns use bromobenzene or chlorbenzene for the liquid and a
halogenated benzoic acid for the solid. This test should be done in a hood.

Ignition Test for Aromaticity. Place a small amount of compound on the end of a spatula
or on a porcelain lid and apply the flame from a Bunsen burner. Highly unsaturated
compounds such as aromatic compounds burn with a yellow, sooty flame. Use
naphthalene for the solid unknown and toluene or xylene for the liquid known. This test
should be done in a hood.

Bromine Test for Unsaturation: Alkenes and alkynes will readily add bromine across the
multiple bond unless there are electron withdrawing groups on the multiple bond. One
observes the rapid disappearance of the red-brown bromine color. Aromatic compounds
can react with bromine more slowly to give bromine substitution and the formation of
HBr, which can sometimes be observed by placing a piece of wet litmus paper over the
mouth of the test tube. Warning: the reagent deteriorates with the formation of HBr so
compare the results with a blank.

56
Into a dry, clean test tube, dissolve 0.1 mL of a liquid (or 50 mg of a solid) in 1
mL of methylene chloride. Add a 2% solution of bromine in a dichloromethane dropwise
with agitation until the bromine color persists. A positive test requires five or more drops
of bromine solution to reach a persistent red-brown color. Use cyclohexene, octene, or
another simple alkene as the known. Use toluene as a known to test for aromaticity.

Permanganate Test for Unsaturation (Baeyer Test): Aqueous permanganate rapidly


oxidizes double and triple bonds while being reduced to MnO2, a brown precipitate.
Therefore, disappearance of the purple color and formation of a brown precipitate in
minutes is a positive test. However, other compounds react slowly with the reagent
including alcohols, aldehydes, phenols, and aromatic amines so interpret your results
carefully and look for corroboration from the other tests.

Into a clean test tube, dissolve 0.1 mL of a liquid (or 50 mg of a solid) in 1 mL of


95% ethanol or 1,2-dimethyoxyethane. Add a 1% solution of aqueous potassium
permanganate dropwise with agitation.

Use cyclohexene, octene or another simple alkene as the known.

Aldehydes and Ketones--2,4-DNP. Hydrazines such as 2,4-dinitrophenylhydrazine react


with the carbonyl group of aldehydes and ketones to give colored precipitates. Normally
the reaction is fast but heating may be necessary. The test solution is prepared using
sulfuric acid and 95% ethanol. Later, if you wish to make a derivative of your compound,
you can use a different 2,4-DNP solution prepared with HCl and methanol. This usually
gives a slower forming precipitate which often provides a derivative of higher purity (and
higher mp). However, the slow formation of the precipitate is not desirable when looking
for a qualitative test signal. The 2,4-DNPs are usually yellow, orange, or red with the
deeper color often signifying higher conjugation via double bonds or aromatic rings.
Dissolve only 1 drop of your liquid compound (or 10 mg of your solid) in a minimum
number of drops of 95% ethanol in a test tube. Add 1 mL of the 2,4-DNP test solution
and agitate. If a precipitate does not form in 10 minutes, heat on a water or steam bath for
a few minutes. For knowns, use cyclohexanone and benzaldehyde. Warning: the trace of
acetone used to clean your glassware will give a positive test.

Tollens Test for Aldehydes and other easily oxidized functional groups. In this test, a
stabilized silver ion is reduced to elemental silver by an easily oxidized compound, such
as an aldehyde. The aldehyde is oxidized to a carboxylic acid. A positive test is the
formation of a silver mirror as the elemental silver adheres to the wall of the glass tube.
Prepare the Tollens reagent immediately before you plan to use it. Mix 3 mL of Tollens
solution A (aqueous silver nitrate) with 3 mL of Tollens solution B (10% aq. NaOH)
resulting in the formation of solid silver oxide. Add 10% ammonium hydroxide solution
dropwise, with agitation, until the silver oxide just dissolves. This produces a silver
ammonium complex and is the Tollens solution you will use for the test.

Into each of 3 clean, dry test tubes, add 2 mL of the Tollens reagent which is
freshly prepared as above. Dissolve 10 mg of a solid (or 1 drop of a liquid) unknown in

57
the minimum amount of bis(2-ethoxyethyl)ether required to give a clear solution (less
than 1 mL). Add the unknown solution dropwise, with agitation, to the first test tube. Mix
vigorously and allow the solution to stand. Do the same for the known compound. Do not
use benzaldehyde for the known. For the blank, simply add 0.5 mL of bis(2-ethoxyethyl)-
ether.

Warning: Tollens reagent forms silver fulminate which is very explosive. The test
solutions must be acidified with dilute acid disposed of in a jar properly labeled for that
purpose, provided by the instructor. The silver mirror can usually be washed clean with
soapy water and a scrub brush. If not, see your instructor.

Chromic Acid Test for Aldehydes and Alcohols. Dissolve 10 mg of a solid (or 1 drop of a
liquid) unknown in reagent grade acetone in a clean, dry test tube. Add a few drops of
chromic acid solution one drop at a time with shaking. Aldehydes and primary and
secondary alcohols are oxidized very quickly. Tertiary alcohols are not oxidized. The
orange color of the Cr(VI) ion is replace by the green color of Cr(III) as the chromium is
reduced. Aliphatic aldehydes are oxidized in less than a minute, aromatic aldehydes take
a bit longer. Since the condition of the acetone is critical, it is wise to carefully run the
blank to be certain that the acetone itself is not giving a false positive. Use benzaldehyde
and an aliphatic aldehyde for aldehyde knowns and 1- and 2-butanol for alcohol knowns.
The chromic acid solution is prepared by dissolving 1.0 g of CrO3 in 1.0 mL of
concentrated sulfuric acid and then carefully diluting with 3 mL of water.

Warning: Cr(VI) compounds are considered suspect carcinogens and should be handled
carefully.

Ferric Hydroxamate Test for Esters. If you have a carbonyl compound which is not an
aldehyde or ketone or carboxylic acid, it could be an ester. One test for esters is the ferric
hydroxamate test whereby the ester is converted to a hydroxamic acid (HOHN-C=O)
which will give a positive ferric chloride test. Since enols can give a positive ferric
chloride test, first test your compound with ferric chloride solution as follows: dissolve
10 mg of solid (or 1 drop of liquid) unknown in 1 mL of 95% ethanol, add 1 mL of 1 N
HCl, and then a 1-2 drops of 5% ferric chloride solution. If you obtain a color other than
yellow, the test cannot be used. Otherwise, the test is conducted as follows: dissolve 50
mg of solid or 2 drops of liquid unknown in 1 mL of 0.5N hydroxylamine hydrochloride
in 95% ethanol and 0.2 mL 6N NaOH. Heat to boiling for 2-3 minutes, then cool and add
2 mL 1N HCl. If the solution becomes cloudy, add 1-2 mL of 95% ethanol to clarify. Add
1 drop of 5% ferric chloride solution. If a color forms and then fades, add additional
drops of 5% ferric chloride until the color persists. The color is due to a complex between
the hydroxamic acid and the ferric ion. A deep burgundy color is positive. Use banana oil
or methyl benzoate as knowns.

Ferric Chloride test for Phenols. Just as enols can form colored complexes with ferric ion,
phenolate ions can as well. Therefore, this test is designed to convert the weakly acidic
phenols to their conjugate base which can then complex with ferric ion. If the phenol is
water soluble, add a few drops of 2.5% aqueous ferric chloride solution to a 3% aqueous

58
solution of the phenol. A deep red, green, or blue color is positive. If the phenol is not
water soluble, dissolve 20 mg of the solid (or 1 drop of the liquid) in 1 mL of methylene
chloride and add 1 drop of pyridine. Add 3 drops of 1% ferric chloride in methylene
chloride. An intense color is a positive test. Use phenol as a known. Not all phenols will
give a positive test.

Iodoform test for methyl ketones. In this test you will convert the methyl ketone to a
triidomethyl ketone which is then cleaved to form iodoform, HCI3, a yellow solid.
Acetone gives a nice positive test so be certain that no traces of acetone are in your
glassware. In a large clean test tube or a vial, place 100 mg of a solid or 5 drops of a
liquid unknown. Add 2 mL bis(2-ethoxyethyl)ether, 5 mL of water, and 1 mL of 10%
NaOH, and mix well. Add a total of 3 mL of iodine-potassium iodide solution in six
equal portions, stopper and shake well after each addition. Caution: seal the tube
carefully and avoid skin contact with the iodine solution. The color of the iodine will
disappear more slowly in the later additions. The solution should be slightly yellow. Heat
if necessary and shake again to force the iodine to react. When the color is slightly
yellow, add water to nearly fill the test tube or container, stopper, and shake vigorously.
After standing for 15 minutes, a pale yellow precipitate of iodoform (mp 119-121oC) is a
positive test for a methyl ketone. Acetone can be used for the known.

The iodine-potassium iodide solution is prepared from 10 g of iodine and 20 g of


potassium iodide in 100 mL of water.

Hinsburg Test for Amines. If you have a basic compound which you believe to be an
amine, you can corroborate your suspicion and determine if you have a primary,
secondary, or tertiary amine using the Hinsberg test. You will react the amine with a
sulfonyl chloride forming an insoluble sulfonamide of a primary or secondary amine or
the soluble salt of a tertiary amine. The insoluble sulfonamide of a primary amine will be
made soluble in base (via removal of the slightly acidic proton on N) but that of a
secondary amine will not (no proton on N to remove).

Add 100 mg of a solid or 0.1 mL of a liquid unknown, 200 mg of p-


toluenesulfonyl chloride, and 5 mL of 10% KOH solution to a clean test tube. Stopper the
tube and shake it for several minutes. Remove the stopper and heat the mixture on a
steam bath for 1 minute. Cool the solution and if it is not basic to pH paper, add
additional KOH solution. If a precipitate has formed, add 5 mL of water and shake
vigorously. If the precipitate does not redissolve in the basic solution, it is indicative of a
sulfonamide of a secondary amine. If there is no precipitate, add 5% HCl until the
solution is just acidic when tested by pH paper. Formation of a precipitate under acidic
conditions suggests that the previously soluble sulfonamide was of a primary amine. If no
precipitate has formed, the initial amine could have been tertiary.

Use aniline, N-methylaniline and N,N-dimethylaniline for knowns.

UNKNOWN REPORTS

59
You should submit an Unknown Preliminary Report as soon as you have completed the
experiments described below. Preliminary and final report forms can be found at the end
of this laboratory manual.

For the solid unknown, you will take an accurate mp, an IR spectrum, perform
thesolubility tests and, based upon an analysis of this information, do specific
classification tests as described in the manual. If you suspect that you have a carboxylic
acid, we will ask you to do a neutralization equivalent. Then submit the preliminary
report. This form is not graded. It gives us a chance to verify that you have not been
misled. We will either OK your data or suggest that you repeat certain procedures. At this
stage, you should have determined the class of your compound (e.g., aldehyde, ketone,
...) and possibly some additional functionality such as unsaturation, presence of a
halogen, etc.

For the liquid unknown, you will measure the boiling point by simple
distillation. In addition you will take an IR spectrum using the sandwich cell NaCl
method, perform the solubility tests and, based upon an analysis of this information,
perform some specific functionality tests as described in the manual. Once completed,
submit the preliminary report form for your liquid.

After the preliminary analysis has been confirmed, try to determine the actual structure of
your unknown. You will need to make a solid derivative of each unknown to verify your
identification of the compounds. For carboxylic acids, the neutralization equivalent may
substitute for the derivative.

Possible compounds and their derivatives are listed in the manual and in other texts
which are available to you. Most of your unknowns are listed in this Manual but you can
consult other texts such as Shriner, Fuson, Curtin, “The Systematic Identification of
Organic Compounds”, Vogel, “Elementary Practical Organic Chemistry, Part 2,
Qualitative Organic Analysis”, “CRC Handbook of Tables for Organic Compound
Identification”--all of these are in the lab and available to you. All tables are listed first
by class of compound (alcohol, acid, ether, etc). Some headings are more specific such as
aliphatic acid, aromatic acid, carboxylic acid, dicarboxylic acid, halogenated acid, etc. Be
certain that you have checked all the tables that pertain to your search. The compounds
are then listed by escalating mp and bp. Therefore, it is critical that you determine the
major functional group or groups in your compound (IR analysis, solubility and
classification tests) so that you know the CLASS of compound. Also, your mp and bp
must be accurate in order to narrow the possible choices within that class. Consider all
possibilities within at least  5 °C of your mp, 10 °C for bp up to 175 °C and a wider
range for mp and bp above 175 °C.

Final report of the unknowns: report the unknown number, your name, appropriate
properties, the results of the solubility tests, the results of the class tests (list the test and
whether it was + or - or could not determine), the neutralization equivalent if taken (must
be reported for all acids and bases using an appropriate number of significant figures), all
possible compounds within that class and close to that mp or bp ( 5 °C for mp,  10 °C),

60
how you distinguished between these possibilities, a thorough analysis of your infrared
spectrum that includes as many features of your compound as you can determine (e.g.,
the major functional group, other functional groups, aromatic units, double or triple
bonds, etc.). List the pages in your notebook where we can find information related to
this particular unknown.

If you made a derivative, you should report the name of the derivative, its
structure, its literature mp, and the mp you observed. If there is a large difference
between lit mp and obs mp, comment on it (e.g., did you recrystallize the derivative?).
List the mp of the corresponding derivative of the other possible compounds. Submit the
derivative in a properly labeled vial which includes your name, name of product, mp of
product, lit mp of product, date prepared. Also return unused unknown compound.

When you submit this final report you should attach your properly labeled
infrared spectrum; e.g., your name, the compound name and/or unknown number. Having
the correct structure is insufficient! You must list other possibilities and explain how
you differentiated between the “correct” compound and those other possibilities.

Determination of Equivalent Weight (or Neutralization Equivalent)


Molecular Weight Determination

Generally, any acid (or base) can be titrated using standard solutions of base (or
acid). The neutralization equivalent obtained is usually a simple fraction of the molecular
weight (1, 1/2, 1/3, etc). In the titration of an acid with standard base, the endpoint is
reached when all the acid is neutralized and a drop of excess base is added. If
phenolphthalein is used as the indicator, it will turn red at this instant. For a dicarboxylic
acid such as malonic acid, the endpoint is reached when the last of the acid is converted
to the carboxylate anion. The neutralization equivalent will be one-half its molecular
weight.

CH2 (CO2H) 2 + 2 NaOH CH2(CO2Na) 2 + 2 H2O

Equivalent weights must be done in duplicate and the values obtained should
agree within a few percent. If not, do a third determination.

Procedure for acids. Into a 125 ml Erlenmeyer flask place approx. 150 mg* of the
unknown, accurately weighed to the nearest 1 mg or 0.1 mg for at least three place
accuracy. Add approx. 5-10 ml of 95% ethanol to dissolve the unknown and add an equal
amount of water. Add a drop or two of phenolphthalein solution, a magnetic stirring bar,
and titrate with standardized aqueous base to the first permanent pink color that lasts
about 60 seconds.

The equivalent weight (neutralization equivalent) can be calculated as follows:


V (ml) x M (mmol/ml)= wt (mg)/equivalent weight (mg/mequiv); V = volume of
standard base used measured accurately to at least 3 significant figures; M= molarity of
standard base as long as the base is monobasic, probably listed as 0.100M NaOH or KOH

61
(note three significant figures); wt= weight of unknown used. An equivalent weight of
120 mg/mequiv means that the molecular weight is some whole number multiple of 120;
for example, 120 (if monoacid there is 1 milliequivalent/millimol) or 240 (if diacid there
are 2 mequiv/mmol) or 360 (if triacid), etc.

Procedure for Bases. Into a 125 ml Erlenmeyer flask place approx. 150 mg* of the
unknown, accurately weighed to the nearest 0.1 mg for at least three place accuracy. Add
approx. 5-10 ml of 95% ethanol to dissolve the unknown and add an equal amount of
water. Add a drop or two of an appropriate indicator solution, a magnetic stirring bar, and
titrate with standardized aqueous acid to the first permanent color change that lasts about
60 seconds. The calculation for equivalent weight is the same except wherever it says
acid, substitute the word base, and where it says base, substitute the word acid. Usually
0.100 M standardized HCl is used for the titration.

 larger quantities will be required if 150 mg is less than 1.0 meq since at least 10.0 mL
of 0.100 N of solution is required for three place accuracy.

Derivatives Aldehydes and Ketones

2,4-Dinitrophenylhydrazones:
Place 10 mL of a solution of 2,4-DNP (in methanol/hydrochloric acid) into a test tube
and add approx. 1 mmol of the unknown compound (add approx 150-250 mg if you do
not know the molecular weight). If the unknown is a solid, dissolve the compound in a
minimum amount of 95% ethanol or 1,2-dimethoxyethane first. Crystallization will
probably not occur immediately (the reaction in ethanol/sulfuric acid is faster), so gently
warm the solution for a minute on a steam bath and then set it aside to crystallize. It may
take several hours. As a backup, you can repeat the procedure using the 2,4-DNP solution
prepared in ethanol/sulfuric acid but the methanolic solution will probably give you
better crystals (higher mp).

Semicarbazones: The stock solution is prepared by dissolving 1.11 g of semicarbazide


hydrochloride (MW 111.5) in 5 mL of water. Place 0.5 mL of this solution in a small test
tube and add an estimated 1 mmol (add approx 150-250 mg if you do not know the
molecular weight ) of the unknown compound. If the unknown does not dissolve in the
aqueous solution, or if the solution becomes cloudy, add enough methanol dropwise to
dissolve the solid or produce a clear solution. Add 10 drops of pyridine and heat the
mixture gently on a steam bath for 5 minutes. The product should begin to crystallize.
The product can be recrystallized from ethanol if necessary.

Alternative Method: Dissolve 1 g of semicarbazide hydrochloride and 1.5 g of


sodium acetate in 5 mL of water. Then dissolve 1.0 g of the unkown in 10 mL of ethanol.
Mix the two solutions together in a 50 mL Erlenmeyer flask and heat to boiling for about
5 minutes. Place the flask in a beaker of ice and scratch the sides of the flask to induce
crystallization.

62
EXPERIMENT 10. MULTISTEP SYNTHESIS
When synthesizing complex organic molecules, it is common to have a series of
individual transformations whereby the product of one reaction is then used as the
starting material for the next reaction. You will have an opportunity to perform a
multistep synthesis starting with inexpensive, readily available benzaldehyde. The
sequence you will attempt is first the conversion of benzaldehyde to benzoin using the
vitamin, thiamin, as a catalyst. In the second step, the benzoin is oxidized to benzil
through the use of an oxidizing agent. The third step is a condensation reaction of benzil
with dibenzyl ketone (1,3-diphenyl-2-propanone) to produce
tetraphenylcyclopentadienone.

One problem that becomes apparent is that the yield of the overall final product
will be limited by the lowest yielding individual reaction. Therefore, each reaction in the
sequence must be a high yielding reaction. Second, the overall final product yield is the
product of each individual percentage yield. Therefore, if each step is a 90% yield and
there are 10 steps, the overall final product yield is (0.90)10 or 35%. For a twenty step
reaction, the overall yield would only be 12%. In a two step reaction, if one step had a
yield of 50%, the highest overall yield possible is 50%. It is not necessary, or desirable,
to use all of your material in each step.

Step 1. Synthesis of Benzoin.

O O OH
C H C C
2 thiamine H

This reaction is a classic--the conversion of two molecules of an aldehyde to an


alpha-hydroxy ketone. The reaction is known as a benzoin condensation (“condensation”
because two molecules become condensed to one molecule). This reaction, which
requires a catalyst, if often performed with cyanide ion. We will use thiamine as a
catalyst. It is heat-sensitive and may decompose if heated too vigorously. Instead of
running this reaction at elevated temperatures for a few hours, we will allow the reaction
to proceed closer to room temperature for 24 hours or more. Benzaldehyde is easily
oxidized to benzoic acid which can impede the desired reaction so freshly distilled
benzaldehyde is used. The concentration of reactants and temperatures of solutions are
critical to obtaining a good yield so procedures must be followed carefully. Too much
water will force benzaldehyde out of solution preventing an efficient reaction. Too little
water prevents the thiamine hydrochloride from dissolving. Some of the base reacts with
the thiamine hydrochloride to produce thiamine which is the active catalyst.

63
Procedure: Place 1.5 mL of 5M NaOH (CAUTION: extremely caustic) in a 10 mL
Erlenmeyer flask and cool in an ice bath. In a 50 mL Erlenmeyer flask dissolve 0.80 g of
thiamine hydrochloride (MW=337) in 2.5 mL of water. Add 7.5 mL of 95% ethanol to
the thiamine and cool the solution for several minutes in an ice bath. While keeping both
flasks in the ice bath, add the 1.5 mL of previously cooled 5M sodium hydroxide
dropwise (3-5 minutes) to the thiamine solution with swirling so that the solution stays
below room temp. Remove the 50 mL flask from the ice bath, add 5.0 mL of
benzaldehyde (d=1.044 g/mL) at one time, swirling the flask so that the benzaldehyde
mixes with the yellow, aqueous, basic layer. The solution becomes milky but then
clears*. Seal the flask with Parafilm and place it in your drawer until the next lab period.

*If the mixture does not go to solution (e.g., if two layers are obvious), place the flask in
a warm water bath at approx. 50 oC until the solution clears or for a maximum of 10
minutes. You can use hot water from the faucets at the front of the lab. The mixture
should become homogeneous in the water bath but it may not stay homogeneous once it
cools.

The following lab period: Filter the crystals, wash them free of mother liquor with 10-15
mL of a cold 2:1 mixture of water and 95% ethanol, and air dry the solid for 15 min.
Weigh your crude yield, break up any clumps of solid and recrystallize from hot 95%
ethanol (8 mL per gram). You should not have to filter the hot solution. After cooling, the
recrystallized benzoin should be filtered, washed with a minimum of a cold 2:1 mixture
of water and 95% ethanol, and air dried for 15 min or left until the next lab period.
Obtain the mp of the recrystallized benzoin (lit mp listed as 133 and 137 oC for d,l-
benzoin, most students will observe a mp of 133 oC) .

When you are satisfied that you have the product you want, you may dispose of
the first filtrate by neutralizing with dilute HCl, then flushing the aqueous layer down the
drain with plenty of water. The second filtrate (from recrystallization) can be flushed
down the drain with water.

Step 2. Oxidation of Benzoin to Benzil.

O OH O O
C C C C
HNO 3
H + nitric oxides

CAUTION: Concentrated nitric acid is extremely caustic and will burn exposed skin.
Work in a hood! Into a 125 mL Erlenmeyer flask, place 2.0 g* of benzoin (weighed to the
nearest tenth of a g) and carefully add 7 mL of concentrated nitric acid. Heat the mixture
on a steam bath with occasional slow swirling for 30 minutes or until the brown-red nitric

64
oxide gases are no longer evolved. The fumes are toxic and noxious so be certain that the
fume hood safety shield is pulled down.

Carefully cool the flask and contents using tap water (keep the flask covered with
a plastic seal or a cork), then pour into 35 mL of cool water and swirl to coagulate the
precipitated product. Collect the yellow solid using suction filtration and wash twice with
5 mL of cool water to remove some of the nitric acid present. Press the crystals to remove
more water by placing another piece of filter paper over the crystals and pushing with a
beaker or cork; the suction flask MUST be supported and sitting flat on the desktop. This
crude product can be recrystallized from 95% ethanol while it is still slightly wet (4
mL/g). Dissolve it in hot ethanol, add water dropwise to reach the cloud point, and allow
it to slowly crystallize. Filter, dry, record the yield, and take the mp.

* Do not use all of the benzoin you have synthesized. If you do not have more than 2.0 g,
save 100 mg of the benzoin for a mp, IR, and to hand in and use the rest for this next
step, altering the procedure to scale.

If your instructor requests, run a tlc of the recrystallized product with known
samples of benzoin and benzil for comparison of Rf values.

When you are satisfied that you have the product you want, you may dispose of
the filtrate by first neutralizing with sodium carbonate, diluting with water, and flushing
down the drain. Ethanol from the recrystallization goes to the non-halogenated waste
container.

Step 3. Preparation of Tetraphenylcyclopentadienone.

O
O
Ph CH2 C CH2 Ph
Ph Ph
KOH
O O + 2H2O
abs ethanol
Ph C C Ph Ph Ph

Ph = phenyl =

Cyclopentadienone is a relatively unstable compound which will dimerize even at


low temperature. However, the corresponding tetraphenyl compound is quite stable.

Procedure: Into a 50 mL rbf, place 0.7 g* of benzil, 0.7 g of dibenzyl ketone, and 5 mL of
absolute ethanol. Attach a reflux condenser and heat the mixture on a steam bath, water
bath or sand bath until the solids dissolve. It is critical to prevent water (moisture) from
coming into contact with the reactants. Raise the temperature to provide a slow reflux

65
and add a solution of 0.1 g of potassium hydroxide (CAUSTIC) in 1 mL of absolute
ethanol (this solution may be already prepared) dropwise through the top of the
condenser. The reaction is very fast and a purple color will appear. After addition of the
base, allow the mixture to reflux for 15 minutes while periodically shaking the flask.
Cool the reaction flask to room temperature, then in an ice bath. Filter using a Buchner
funnel, wash twice with 5-mL portions of cold 95% ethanol, and air dry for an hour, if
possible. When the crystals are dry (which may take until the next lab period), weigh,
record the yield and the percentage yield. You may recrystallize a portion of the purple
product using a 1:1 mixture of 95% ethanol and toluene (12 mL/0.5 g). Record the mp;
literature mp 219-220 °C).

When you are satisfied that you have the product you want, the filtrate can be
neutralized with dilute aqueous HCl and flushed down the drain. The recrystallizing
solvent should be placed in the non-halogenated waste container.

*Do not use all of the benzil you have synthesized. If you do not have more than 0.7 g,
save 100 mg of the benzil for a mp, IR, and to hand in and use the rest for this next step,
altering the procedure to scale or obtaining additional benzil from your instructor.

Alternative Step 3. Reduction of Benzil with Sodium Borohydride.


There are a wide variety of hydride reducing agents that convert carbonyl compounds
into alcohols. One of the least reactive of these agents is sodium borohydride. Although it
reduces aldehydes and ketones, it is fairly stable in aqueous and alcoholic solutions. The
more reactive hydride reducing agents can reduce other functional groups such as
carboxylic acids, esters, epoxides, and nitriles. Such hydrides react violently with water
releasing hydrogen gas and must be handled very carefully. You will reduce the diketone,
benzil, using sodium borohydride. Three stereoisomers of hydrobenzoin can be formed in
this reaction, RR, SS, and RS which is the meso isomer. It is the meso isomer that
predominates. The stoichiometry of the typical borohydride reaction is:

4 R2C=O + NaBH4  (R2CHO)4B-Na+

hydrolysis of the borate ester (R2CHO)4B-Na+ + H2O  4 R2CHOH

Your starting compound is a diketone so you need one mmol of borohydride for
every two mmol of ketone .

Procedure: Using a 25 mL or a 50 mL Erlenmeyer flask, dissolve 0.50 g of benzil (weigh


to the nearest hundredth of a gram) in 5 mL of warm 95% ethanol. Cool the solution in a
water bath which will produce a fine suspension of benzil particles. Add 0.10 g of sodium
borohydride (weight to the nearest hundredth of a g) which will cause the solution to
warm and dissolve the suspended benzil. As the reduction reaction proceeds in the next
few minutes, the yellow color of benzil will disappear. After a total of 10 minutes, add 5
mL of water, heat to boiling on a steam bath, filter or decant if the solution is not clear.
When the solution cools, dilute to the saturation point with as much as 10 mL of water

66
and set the solution aside for crystallization to occur. In your discussion, mention that
three stereoisomers are possible and suggest why the meso isomer (lit mp 136-7 °C)
predominates.
When you are satisfied that you have the product you want, you may dispose of
the filtrate by diluting with water, neutralizing the excess, unreacted borohydride with
acetic acid, and flushing down the drain.

Step 4. Dimethyltetraphenylphthalate

O Ph

Ph Ph R C C R Ph R
+ CO
(R = CO2CH3) Ph R
Ph Ph
Ph

Into a 10 mL beaker, place 39 mg of tetraphenylcyclopentadienone (0.10 mmol),


3 drops of dimethyl acetylenedicarboxylate (an excess) and 0.3 mL of triethylene glycol.
Mix the ingredients by swirling, cover the beaker with a thin watch glass or appropriate
microwave safe film (thick watch glasses often break in the microwave) and place it in
the microwave oven. Set the oven at power level=6 for 5 minutes. After irradiating for 5
min, the beaker will be hot. Let it cool for a minute. The reaction mixture should be a
golden color which cools to colorless crystals. It may take a few hours to see crystals. If
necessary, leave the material to crystallize until the next class. Collect the crystals in a
micro Hirsch funnel and wash with a few drops of cold 95% ethanol. Recrystallize using
95% ethanol. Record the product mp; literature mp 255-257 °C.

67
INFRARED SPECTROSCOPY
Most of us are quite familiar with infrared radiation. We have seen infrared lamps keep
food hot and often associate infrared radiation with heat. While the generation of heat is a
probable event following the absorption of infrared radiation, it is important to
distinguish between the two. Infrared is a form of radiation that can travel through a
vacuum while heat is associated with the motion and kinetic energy of molecules. The
concept of heat in a vacuum has no meaning because of the lack of molecules and
molecular motion. Infrared spectroscopy is the study of how molecules absorb infrared
radiation and ultimately convert it to heat. By examining how this occurs, we will not
only learn about how infrared radiation is absorbed, but we will also learn about
molecular structure and how the study of infrared spectroscopy can provide information
about the structure of organic molecules. An infrared spectrum of a chemical substance is
very much like a photograph of a molecule. However, unlike a normal photograph that
would reveal the position of nuclei, the infrared spectrum will only reveal a partial
structure. It is the purpose of this narrative to provide you with the tools necessary to
interpret infrared spectra, successfully. In some respects, this process is similar to reading
an X-ray of the chest. While most of us could easily identify the gross structural features
of the chest such as the ribs, most of us would need some guidance in identifying those
features on the X-ray film associated with disease.

In order to interpret infrared spectra, having some idea or model of the physical
process involved when a molecule interacts with infrared radiation would be useful. You
may recall in introductory chemistry, the discussion of how atoms interact with
electromagnetic radiation led to the development of quantum theory and the introduction of
quantum numbers. The interaction of infrared radiation with molecules requires a similar
treatment. While the use of quantum theory is necessary to explain this interaction, most of
us live in a world that appears continuous to us and we do not have much experience
discussing phenomena that occur in discrete steps. The discussion that follows will attempt to
develop a model of how molecules interact with infrared radiation that is based as much as
possible on classical physics. When necessary, we will insert the modifications required by
quantum mechanics. This model, while perhaps oversimplified, will contain the physical
picture that is useful to understand the phenomena and will be correct from a quantum
mechanical standpoint.

Let's begin first by considering two isolated atoms, a hydrogen and a bromine atom
moving toward each other from a great distance. What do you suppose will happen once the
atoms approach each other and can feel each other’s presence? The potential energy curve
for the H-Br molecule is shown in Figure IR-1. As the two atoms approach each other notice
that the potential energy drops. If we recall that energy must be conserved, what must happen
to the kinetic energy? The two atoms must attract each other and accelerate toward each
other, thereby increasing their kinetic energy. The change in kinetic energy is illustrated by
the dotted line in the figure. At some point they will “collide” as indicated by the part of the
potential energy curve that rises steeply at small interatomic distances and then the atoms

68
will

Potential Energy Diagram for HBr

1
Potential Energy

-1
0 1 2 3 4 5
Internuclear Separation

Figure IR-1. The potential (solid line) and kinetic energy (dotted line) of HBr as
a function of the separation of the two nuclei. The kinetic energy at every point
illustrated by the dotted line is equal to the potential energy plus the small amount of
kinetic energy associated with initial motion of the two nuclei when separated at large
distances.

begin to move away from each other. At this point, we might ask, "Will the molecule of HBr
survive the collision"? Unless some energy from this system is lost, say by emission of a
photon of light or collision by a third body to remove some energy, these are two ships
passing in the night. The kinetic energy resulting from the coulombic attraction of the two
atoms will exactly equal the drop in potential energy and the two atoms will fly apart. The
spontaneous emission of a photon of light is improbable, so this mechanism is unlikely to
drop the HBr molecule into the well. Most probable from a physical perspective, is the
collision of our HBr with a third body that will remove some energy and result in the
trapping of the HBr molecule in the well. Though very excited, this molecule will now
survive until other collisions with less energetic molecules leads to an HBr molecule at the

69
bottom of the well and the generation of heat (kinetic energy) that would be experienced in
the exothermic reaction of hydrogen and bromine atoms to form hydrogen bromide. Let us
now consider a hydrogen bromide molecule that has lost a little kinetic energy by collision
and has been trapped in the potential energy well of Figure IR-1. We might ask, “How would
a molecule that does not have enough kinetic energy to escape the well behave in this well?
A molecule with some kinetic energy below this threshold value (total energy slightly less
than 0 in Fig. IR-1) will be able to move within this well. The internuclear separation will
vary within the limits governed by the available kinetic energy. Since this motion involves a
stretching or compression of the internuclear distance it is usually described as a vibration.
Additional collisions with other molecules will eventually lead to the dissipation of the
energy associated with formation of the hydrogen bromide bond. At this point we might ask
the following question. If we remove all the excess kinetic energy from HBr, what will be its
kinetic and potential energy? Alternatively we might ask, "Will the hydrogen bromide
molecule reside at the very bottom of the well when it is cooled down to absolute zero
Kelvin?" Before we answer this question, let's digress for a little and discuss the relative
motions of the hydrogen and bromine atoms in terms of the physics of everyday objects.
Once we learn how to describe the classical behavior of two objects trapped in a potential
energy well, we will return to the question we have just posed.

One model we can use to describe our hydrogen bromide molecule is to consider our
HBr molecule to be made up of balls of uneven mass connected to each other by means of a
spring. Physicists found many years ago some interesting properties of such a system which
they referred to as a harmonic oscillator. Such a system repeatedly interconverts potential
and kinetic energy, depending on whether the spring is exerting a force on the balls or the
momentum of the balls is causing the spring to be stretched or compressed. The potential
energy of this system (PE) is given by the parabola,

PE = k(x-xo)2 1

where x-xo is the displacement of the balls from their equilibrium condition when the
system is at rest and k is a measure of the stiffness of the spring. While this simple
equation does not apply to molecules, please notice how similar the potential energy
surface of the parabola (Figure IR-3) is to the bottom of the surface of Figure IR-1. The
constant k is used to describe chemical bonds and is referred to as the force constant. As
you might imagine, it is a measure of the stiffness of the chemical bond.

Several other relationships were observed that do carry over in describing


molecular systems. For example, they found that when a ball was suspended on a spring
from a horizontal wall, the frequency of vibration or oscillation,  depended only on the
mass of the ball and the stiffness of the spring. The term A is a constant of the
proportionality. By varying the mass of the ball and the stiffness of the spring, they were
able to uncover the following simple relationship between frequency, mass and force
constant:

70
k
A 2
m

Suspending a ball and spring from a horizontal surface is a special case of the more
general situation when you have two more comparable masses attached to each other.
Under these circumstances, when two similar masses are attached to a spring, the
relationship between frequency of vibration, mass and force constant is given by:
k
A 3

where  represents the product of the masses divided by their sum (m1m2)/(m1+m2).
This latter term is found in other physical relationships and has been given the name, the
reduced mass. It can easily be seen that equation 2 is a special case of the more general
relationship given by equation 3. If we consider m1to be much larger than m2, the sum of
m1+ m2  m1 and substituting this approximation into (m1m2)/(m1+m2)  m2.
Substituting m2 into equation 3 where m2 is the smaller of the two masses gives us
exactly the same relationship as we had above when the ball was suspended from a
horizontal wall. The horizontal wall is much more massive than the ball so that the
vibration of a smaller ball has very little effect on the wall. Despite their simplicity,
equations 2 and 3 play an important role in explaining the behavior of molecular systems.
However, before we discuss the important role these equations play in our understanding
of infrared spectroscopy, we need to review some of the properties of electromagnetic
radiation, particularly radiation in the infrared range.

The electromagnetic spectrum is summarized in Figure IR-2. On the extreme right we


find radiowaves and scan from right to left we encounter terms which have become familiar
to us; microwave, infrared, visible ultraviolet and X-rays. All of these forms of
electromagnetic radiation

Wavenumbers, cm-1

1010 108 106 2×105 1x105 4000 650 12 5×10−2 10−3to 10−6

Gamma X- Ultra Visible Near Infra Far Micro TV Radio


Rays Rays violet Light IR red Infrared wave Waves

10-6 10-4 10-2 5x10-2 10-1 2.5 15.4 830 4x105 107 1010

Wavelength (microns)

Figure IR-2. The electromagnetic spectrum.

71
are related to each other in a simple and obvious way. First let us discuss why we refer to
these different forms of light as electromagnetic radiation. Simply stated, all these forms of
radiation have an electric and magnetic field associated with them that varies as shown for
the standing wave in Figure IR-3. Only the electric field is shown in this figure. If we were to
include the magnetic field it would look exactly as the electric field but would be rotated 90 °
out of the plane of the paper and would oscillate above and below the plane of the paper like
a sin or cos wave. In infrared spectroscopy, only the electric field associated with the
electromagnetic radiation is important and we will limit our present discussion to how this
field varies with time. We called the light wave associated with Figure IR-3 a standing wave
because

Electric Field of Light

1
Electric field

-1

-2
0 2 4 6 8 10 12 14 16 18
distance

Figure IR-3. The electric field of light associated with a standing wave with a fixed
wavelength.

this is how the electric field would vary if we took a picture of the wave. One of the
properties of all electromagnetic radiation is that it travels in a vacuum at the speed of 3 x
1010 cm/sec. Therefore, if we were to turn this standing wave "on" we would observe
this oscillating field rapidly passing us by. If we examine the electric field (or the

72
magnetic field which is not shown), we observe that the field is repetitive, varying as a
cos or sin wave. The length of the repeat unit along the x axis is called the wavelength, ,
and it is this property which varies continuously from 106 cm (1010 microns) for radio
waves down to 10-13 cm (10-6 microns) for cosmic radiation. A unit of length that is
frequently used in infrared spectroscopy is the micron. A micron is equivalent to 10-4
cm. If we were to "stand on the corner and watch all the wavelengths go by", since all
electromagnetic radiation would be traveling at 3 x 1010 cm/sec, the frequency, , at
which the shorter wavelengths would have to pass by would have to increase in order to
keep up with the longer wavelengths. This relationship can be described in the following
mathematical equation:

= c; (c = 3 x 1010 cm/sec). 4

The frequency of the light times the wavelength of the light must equal the speed at which
the light is traveling.

In addition to having wave properties such as the ones we have been discussing,
electromagnetic radiation also has properties we would normally attribute to particles.
These “particle like” properties are often referred to as characteristics of photons. We
can discuss the wave properties of photons by referring to the wavelength (eqn. 4) and
frequency associated with a photon. The energy of a single photon is a measure of a
property we would normally associate with a particle. The relationship which determines
the energy associated with a single photon of light, E, and the total energy incident on a
surface by monochromatic light, ET, is given by:

E = h  (or equivalently, E = h c/  from equation 4 

ET = n h  6

where h is Planck's constant and is numerically equal to 6.6 x 10-27 erg s and n is the
number of photons. Equations 4 and 5 tell us that photons with short wavelengths, in
addition to having higher frequencies associated with them, also carry more punch! The
energy associated with a photon of light is directly proportional to its frequency.

At this point we are ready to return to a discussion of how infrared radiation interacts
with molecules. Following our discussion of balls and springs, you have probably figured
that infrared spectroscopy deals with the vibration of molecules. Actually, both rotation and
vibration of molecules is involved in the absorption of infrared radiation, but since molecular
rotation is not usually resolved in most infrared spectra of large organic molecules, we will
ignore this additional consideration. In order to derive the relationship between vibrational
energy and molecular structure, it is necessary to solve the Schoedinger equation for
vibrational-rotational interactions. Since solution of this equation is beyond the scope of this
treatment, we will simply use the relationship that is derived for a harmonic oscillator from
this equation. As you see, the quantum mechanical solution of a harmonic oscillator,

73
equation 7, is remarkably simple and very similar to the relationship we obtained from
considering the classical model of balls and springs.

h k 1
E = (n  ) 7
2  2

Before discussing the implications of equation 7, let's take a moment to see how
similar it is to equations 3 and 5. From equation 5, we see that substituting equation 3 for
 results in equation 7 except for the (n + 1/2) term. However we should point out that
we have substituted the vibrational frequency of two masses on a spring for a frequency
associated with the number of wave maxima (or minima, null points. etc.) passing a
given point (or street corner) per unit time. We are able to do this because of the presence
of the (n +1/2) term. Let's discuss the significance of the (n + 1/2) term before we
returning to answer this question. The previous time you encountered the Schroedinger
equation was probably when studying atomic spectra in Introductory Chemistry. An
important consequence of this encounter was the introduction of quantum numbers, at
that time the principle quantum number, N, the azimuthal quantum number, l, the
magnetic, ml, and spin quantum number, s. This time is no exception. Meet n, the
vibrational quantum number. These numbers arise in a very similar manner. The
Schroedinger equation is a differential equation which vanishes unless certain terms in it
have very discrete values. For n, the allowed values are 0,1,2,... Let us now consider the
energy of vibration associated with a molecule in its lowest energy of vibration, n = 0.
h k
According to equation 7, the energy of vibration is given by E = , when n = 0,
4 
the zero point energy. This equation allows us to answer the question posed earlier about
what would happen to the vibrational energy of a molecule at absolute zero. According to
quantum theory the molecule would continue to vibrate. From the relationship E =
1 k
hwe can evaluate the vibrational frequency as  = , the same as found by
4 
classical physics for balls and springs. This equation states that the vibrational frequency
of a given bond in a molecule depends only on the stiffness of the chemical bond and the
masses that are attached to that bond. Similarly, according to equation 7, once the
structure of a molecule is defined, the force constants and reduced mass are also defined
by the structure. This also defines the vibrational frequencies and energy of absorption.
Stated in a slightly different manner, a molecule will not absorb vibrational energy in a
continuous fashion but will do so only in discrete steps as determined by the parameters
in equation 7 and illustrated for the HBr molecule in Figure IR-4. We have pointed out
that the vibrational quantum number can have positive integer values including a value of
zero. Upon absorption of vibration energy, this vibrational quantum number can change
by +1 unit. At room temperature, most molecules are in the n = 0 state.

Figure IR-4 illustrates the real vibrational levels for HBr. Notice that equation 7
predicts that the energy level spacings should all be equal. Notice according to Figure 4, the
spacings actually converge to a continuum for large values of n. For small values of n, n = 0,

74
1, 2, equation 7 gives a good approximation of the vibrational energy levels for HBr.
Equation 7 was derived from the approximation that the potential energy surface is like a
parabola. Near the minimum of this surface, around the zero point energy, this is a good
approximation. As you go up from the minimum, the resemblance decreases and the
assumptions made in solving the Schroedinger equation no longer are valid.

Let us now return and question the wisdom of substituting the vibrational
frequency of a molecule for the frequency of electromagnetic radiation in equation 5. I
hope at this point of the discussion this does not seem so absurd. If the vibrational
frequency of the molecule, as determined by the force constant and reduced mass, equals
the frequency of the electromagnetic radiation, then this substitution makes good sense.
In fact, this gives us a mechanism by which we can envision why a molecule will absorb
only distinct frequencies

Potential Energy Surface for HBr

1
Potential Energy

zero point energy

-1
0 1 2 3 4
Internuclear separation

Figure IR-4. The potential energy surface for a HBr molecule illustrating how the
vibrational energy levels vary in energy with increasing vibrational quantum number.

75
of electromagnetic radiation. It is known that symmetrical diatomic molecules like
nitrogen, oxygen and hydrogen, do not absorb infrared radiation, even though their
vibrational frequencies are in the infrared region. These homonuclear diatomic molecules
have no permanent dipole moment and lack a mechanism by which they can interact with
the electric field of the light. Molecules like HBr and HCl which have a permanent
dipole, resulting from an unequal sharing of the bonding electrons, have a dipole which
oscillates as the bond distance between the atoms oscillate. As the frequency of the
electric field of the infrared radiation approaches the frequency of the oscillating bond
dipole and the two oscillate at the same frequency and phase, the chemical bond can
absorb the infrared photon and increase its vibrational quantum number by +1. This is
illustrated in Figure IR-5. Of course, some HBr molecules may not be correctly oriented
toward the light to interact and these molecules will not absorb light. Other factors will
also influence the intensity and shape of the absorption. However, when the frequency of
the electromagnetic radiation equals the vibrational frequency of a molecule, absorption
of light does occur and this leads to an infrared spectrum that is characteristic of the
structure of a molecule.

Electric Field of Light



1
Electric field

-
-1 d

-2
0 2 4 6 8 10 12 14 16 18
distance

Figure IR-5. An HBr molecule interacting with electromagnetic radiation. In order for
this interaction to occur successfully, the frequency of the light must equal the natural
vibrational frequency of the HBr and the electric field must be properly orientated.

76
Up to now we have discussed molecules changing their vibrational quantum number by +1.
A change of -1 is also equally possible under the influence of infrared radiation. This would
lead to emission of infrared radiation. The reason why we have not discussed this possibility
is that most molecules at room temperature are in the ground vibrational level (n=0) and
cannot go any lower. If we could get a lot of molecules, let’s say with n = 1, use of infrared
could be used to stimulate emission. This is how an infrared laser works.
We have previously discussed the infrared region of the electromagnetic spectrum in
terms of the wavelength of the light that is involved, 2.5-15  ((4000-650 cm-1) (Figure
IR-3). According to equation 4, we can also express this region of the electromagnetic
spectrum in terms of the frequency of the light. There is an advantage to discussing the
absorption of infrared radiation in frequency units. According to equation 5, energy is
directly proportional to frequency. The energy associated with an absorption occurring at
twice the frequency of another can be said to require twice the energy. Occasionally,
weak bands occur at twice the frequency of more intense bands. These are called
overtones and they result when the vibrational quantum number changes by +2. While
these transitions are weak and are theoretically forbidden (i.e. they occur with an
intensity of less than 5 % of the same transition that involves a change of +1 in the
vibrational quantum number) they are easy to identify when units of frequency are used.
Sometimes absorption bands involving a combination of frequencies occur. There is no
physical significance to adding together wavelengths - there is a physical significance to
the addition of frequencies since they are directly proportional to energy. To convert
wavelength to frequency according to equation 4, we need to multiply the speed of light
by the reciprocal of wavelength. Since the speed of light is a universal constant, the
curious convention of simply using the reciprocal of wavelength has evolved. Thus a
peak at 5 would be expressed as 1/(5x10–4 cm) or 2000 cm–1. You will note that 2000
cm–1 is not a true frequency. A true frequency would have units of cycles/sec. To convert
2000 cm–1 to a true frequency one would need to multiply by the speed of light (cm/sec).
However, 2000 cm–1 is proportional to frequency and this is how frequency units in
infrared spectroscopy are expressed. What would be the frequency of light with a
wavelength of 10 

Analysis of IR Spectra
At this point we are ready to leave diatomic molecules and start talking about complex
organic molecules. Before doing so, it should be pointed out that the discussion that
follows is an oversimplification of the true vibrational behavior of molecules. Many
vibrational motions of molecules are motions that involve the entire molecule. Analysis
of such motions can be very difficult if you are dealing with substances of unknown
structure. Fortunately, the infrared spectrum can be divided into two regions, one called
the functional group region and the other the fingerprint region. The functional group
region is generally considered to range from 4000 to approximately 1500 cm–1 and all
frequencies below 1500 cm–1 are considered characteristic of the fingerprint region. The
fingerprint region involves molecular vibrations, usually bending motions, that are
characteristic of the entire molecule or large fragments of the molecule. Hence the origin
of the term. Used together, both regions are very useful for confirming the identity of a

77
chemical substance. This is generally accomplished by a comparison of the spectrum of
an authentic sample. As you become more proficient in analyzing infrared spectra, you
may begin to assign bands in this region. However, if you are just beginning to interpret
spectra of organic molecules, it is best to focus on identifying the characteristic features
in the functional group region. The functional group region tends to include motions,
generally stretching vibrations, that are more localized and characteristic of the typical
functional groups found in organic molecules. While these bands are not very useful in
confirming identity, they do provide some very useful information about the nature of the
components that make up the molecule. Perhaps most importantly, the frequency of these
bands are reliable and their presence or absence can be used confidently by both the
novice and expert interpreter of infrared spectra. The discussion which follows focuses
primarily on the functional group region of the spectrum. Some functional groups are
discussed in more detail than others. You will find that all this information is
summarized in Table 1 which should prove useful to you when you try to interpret an
unknown spectrum. Finally, you should bear in mind that although we have developed a
model that can help us understand the fundamental processes taking place in infrared
spectroscopy, interpretation of spectra is to a large extent an empirical science.
Information about the nature of a compound can be extracted not only from the
frequencies that are present but also by peak shape and intensity. It is very difficult to
convey this information in Table form. Only by examining real spectra will you develop
the expertise to accurately interpret the information contained within. Be sure to examine
the spectra contained in this handout carefully. Whenever you interpret a spectrum and
extract structural information, check your assignments by examining the spectrum of a
known substance that has similar structural features.

Factors Influencing the Location and Number of Peaks

Before beginning a detailed analysis of the various peaks observed in the functional
group region, it might be useful to mentioned some of the factors that can influence the
location and number of peaks we observe in infrared spectroscopy. Theoretically, the
number of fundamental vibrations or normal modes available to a polyatomic molecule
made up of N atoms is given by 3N-5 for a totally linear molecule and 3N-6 for all
others. By a normal mode or fundamental vibration, we mean the simple independent
bending or stretching motions of two or more atoms, which when combined with all of
normal modes associated with the remainder of the molecule will reproduce the complex
vibrational dynamics associated with the real molecules. Normal modes are determined
by a normal coordinate analysis (which will not be discussed in this presentation). If each
of these fundamental vibrations were to be observed, we would expect either 3N-5 or
3N-6 infrared bands. There are some factors that decrease the number of bands observed
and others that cause an increase in this number. Let’s discuss the latter first.

78
Table 1. A summary of the principle infrared bands and their assignments.
R is an aliphatic group.

Funct. Type Frequencies Peak Examples


Group cm-1 IntensityFigure No.
C-H sp3 hybridized R3C-H 2850-3000 M(sh) 6, 18, 22
sp2 hybridized =CR-H 3000-3250 M(sh) 7, 13, 42
sp hybridized C-H 3300 M-S(sh) 13
aldehyde C-H H-(C=O)R 2750, 2850 M(sh) 14, 15
N-H primary amine, amide RN-H2, RCONH2 3300, 3340 S,S(br) 18, 19
secondary amine, amide RNR-H, RCONHR 3300-3500 S(br) 20, 21
tertiary amine, amide RN(R3), RCONR2 none 22, 23
O-H alcohols, phenols free O-H 3620-3580 W(sh) 17, 24, 25
hydrogen bonded 3600-3650 S(br) 24, 25, 28
carboxylic acids R(C=O)O-H 3500-2400 S(br) 26, 27, 29, 30
CN nitriles RCN 2280-2200 S(sh) 31
CC acetylenes R-CC-R 2260-2180 W(sh) 32
R-CC-H 2160-2100 M(sh) 13
C=O aldehydes R(C=O)H 1740-1720 S(sh) 14
ketones R(C=O)R 1730-1710 S(sh) 35
esters R(CO2)R 1750-1735 S(sh) 33, 34
carboxylic acids RCO2H 1720-1680 S(sh) 26,27,30
amides (Amide I) RCONH2, RCONHR 1670-1640 S(sh) 19, 21
(Amide II) RCONH2 1650-1620 S(sh) 19, 21
(Amide II) RCONHR 1550 S(sh) 19, 21
amide RCONR2 1650-1620 S(sh) 23
anhydrides R(CO2CO)R 1820, 1750 S, S(sh) 36
- +
carboxylates R(CO2) , M 1600, 1400 S,S(sh) 42
C=C olefins R2C=CR2 1680-1640 W(sh) 10, 39, 40
R2C=CH2 1600-1675 M(sh) 9, 35
R2C=C(OR)R 1600-1630 S(sh) 41
-NO2 nitro groups RNO2 1550, 1370 S,S(sh) 28

We have already mentioned overtones, which are absorption of energy caused by


a change of 2 rather than 1 in the vibrational quantum number. While overtones are
usually forbidden transitions and therefore are weakly absorbing, they do give rise to
more bands than expected. Overtones are easily identified by the presence of a strongly
absorbing fundamental transition at slightly more than half the frequency of the overtone.
On occasion, combination bands are also observed in the infrared. These bands, as their
name implies, are absorption bands observed at frequencies such as 1 + 2 or 1 - 2,
where 1 and 2 refer to fundamental frequencies. Other combinations of frequencies are
possible. The symmetry properties of the fundamentals play a role in determining which
combinations are observed. Fortunately, combination bands are seldom observed in the

79
functional group region of most polyatomic molecules and the presence of these bands
seldom cause a problem in identification. Another cause of splitting of bands in infrared
is due to a phenomena called Fermi Resonance. While a discussion of Fermi Resonance
is beyond the scope of this presentation, this splitting can be observed whenever two
fundamental motions or a fundamental and combination band have nearly the same
energy (i.e. 1 and 22 or 1 and 2 + 3). In this case, the two levels split each other. One
level increases while the other decreases in energy. In order to observe Fermi Resonance,
in addition to the requirement that a near coincidence of energy levels occurs, other
symmetry properties of these vibrations must also be satisfied. As a consequence, Fermi
Resonance bands are not frequently encountered.

There are also several factors which decrease the number of infrared bands observed.
Symmetry is one of the factors that can significantly reduce the number of bands observed in
the infrared. If stretching a bond does not cause a change in the dipole moment, the vibration
will not be able to interact with the infrared radiation and the vibration will be infrared
inactive. Other factors include the near coincidence of peaks that are not resolved by the
spectrometer and the fact that only a portion of the infrared spectrum is usually accessed by
most commercial infrared spectrometers.

This concludes the general discussion of infrared spectroscopy. At this point we are
ready to start discussing some real spectra.

Carbon-Hydrogen Stretching Frequencies

Let's take one more look at equation 7 and consider the carbon-hydrogen stretching
frequencies. Since k and mH are the only two variables in this equation, if we assume that
all C-H stretching force constants are similar in magnitude, we would expect the
stretching frequencies of all C-H bonds to be similar. This expectation is based on the
fact that the mass of a carbon atom and whatever else is attached to the carbon is much
larger the mass of a hydrogen. The reduced mass for vibration of a hydrogen atom would
be approximately the mass of the hydrogen atom which is independent of structure. All
C-H stretching frequencies are observed at approximately 3000 cm–1, exactly as
expected. Fortunately, force constants do vary some with structure in a fairly predictable
manner and therefor it is possible to differentiate between different types of C-H bonds.
You may recall in your study of organic chemistry, that the C-H bond strength increased
as the s character of the C-H bond increased. Some typical values are given below in
Table 2 for various hybridization states of carbon. Bond strength and bond stiffness
measure different properties. Bond strength measures the depth of the potential energy
well associated with a C-H. Bond stiffness is a measure of how much energy it takes to
compress or stretch a bond. While these are different properties, the stiffer bond is
usually associated with a deeper potential energy surface. You will note in Table 2 that
increasing the bond strength also increases the C-H bond stretching frequency.

80
Table 2. Carbon Hydrogen Bond Strengths as a Function of Hybridization
Bond Strength IR Frequency
Type of C-H bond
kcal/mol cm-1
sp3 hybridized C-H CH3CH2CH2-H 99 <3000
2 CH2=CH-H
sp hybridized C-H 108 >3000
sp hybridized C-H HCC-H 128 3300

C-H sp3 hybridization

Methyl groups, methylene groups and methine hydrogens on sp3 carbon atoms all absorb
between 2850 and 3000 cm–1. While it is sometimes possible to differentiate between
these types of hydrogen, the beginning student should probably avoid this type of
interpretation. It should be pointed out however, that molecules that have local
symmetry, will usually show symmetric and asymmetric stretching frequencies. Take, for
example, a CH2 group. It is not possible to isolate an individual frequency for each
hydrogen atom. These two hydrogens will couple and will show two stretching
frequencies, a symmetric stretching frequency in which stretching and compression of
both hydrogens occurs simultaneously, and an asymmetric stretching frequency in which
stretching of one hydrogen atom is accompanied by compression of the other. While
these two motions will occur at different frequencies, both will be found between the
2850-3000 cm–1 envelope. Similarly for a CH3 group, symmetric and asymmetric
vibrations are observed. This behavior is found whenever this type of local symmetry is
present. We will find other similar examples in the functional groups we will be
discussing. Some examples of spectra containing only sp3 hybridization can be found in
Figures IR-5,6, and located at the end of this discussion. These peaks are usually sharp
and of medium intensity. Considerable overlap of several of these bands usually results
in absorption that is fairly intense and broad in this region.

C-H sp2 hybridization

Hydrogens attached to sp2 carbons absorb at 3000-3250 cm–1. Both aromatic and vinylic
carbon hydrogen bonds are found in this region. An example of a molecule that contains
only sp2 hybridization can be found in Figure IR-7. Other examples of molecules that
contain sp2 C-H bonds along with other functional groups include Figures IR-13, 25 and
37. Examples of hydrocarbons that contain both sp2and sp3 hybridization can be found in
Figures IR-8-12. These peaks are usually sharp and of low to medium intensity.

C-H sp hybridization

Hydrogens attached to sp carbons absorb at 3300 cm–1. An example of a spectrum that


contains sp hybridization can be found in Figure IR-13. These peaks are usually sharp
and of medium to strong intensity.

C-H aldehydes

81
Before concluding the discussion of the carbon hydrogen bond, one additional type of C-
H stretch can be distinguished, the C-H bond of an aldehyde. The C-H stretching
frequency appears as a doublet, at 2750 and 2850 cm–1. Examples of spectra that contain
a C-H stretch of an aldehyde can be found in Figures IR-14 and 15. You may (should)
question why the stretching of a single C-H bond in an aldehyde leads to the two bands
just described. The splitting of C-H stretching frequency into a doublet in aldehydes is
due to the phenomena we called “Fermi Resonance”. It is believed that the aldehyde C-H
stretch is in Fermi resonance with the first overtone of the C-H bending motion of the
aldehyde. The normal frequency of the C-H bending motion of an aldehyde is at 1390
cm–1. As a result of this interaction, one energy level drops to ca. 2750 and the other
increases to ca. 2850 cm–1. Only one C-H stretch is observed for aldehydes that have the
C-H bending motion of an aldehyde significantly shifted from 1390 cm–1.

C-H exceptions

In summary, it is possible to identify the type of hydrogen based on hybridization by


examining the infrared spectra in the 3300 to 2750 cm–1 region. Before concluding, we
should also mention some exceptions to the rules we just outlined. Cyclopropyl
hydrogens which are formally classified as sp3 hybridized actually have more s character
than 25 %. Carbon-hydrogen frequencies greater than 3000 cm–1 are observed for these
stretching vibrations. Halogen substitution can also affect the C-H stretching frequency.
The C-H stretching frequencies of hydrogens attached to a carbon also bearing halogen
substitution can also be shifted above 3000 cm–1. This is illustrated in Figure IR-16. The
last exception we will mention is an interesting case in which the force constant is
increased because of steric interactions. The infrared spectrum of tri-t-butylcarbinol is
given in Figure IR-17. In this case, the hydrogens are sp3 hybridized but stretching the C-
H bonds leads to increased crowding and bumping, and this is manifested by a steeper
potential energy surface and an increase in k, the force constant in equation 6.

Nitrogen Hydrogen Stretching Frequencies

Much of what we have discussed regarding C-H stretching frequencies is also applicable
here. There are three major differences between the C-H and N-H stretching frequencies.
First, the force constant for N-H stretching is stronger, there is a larger dipole moment
associated with the N-H bond, and finally, the N-H bond is usually involved in hydrogen
bonding. The stronger force constant leads to a higher frequency for absorption. The N-H
stretching frequency is usually observed from 3500-3200 cm–1. The larger dipole moment
leads to a stronger absorption and the presence of hydrogen bonding has a definite influence
on the band shape and frequency position. The presence of hydrogen bonding has two major
influences on spectra. First, its presence causes a shift toward lower frequency of all
functional groups that are involved in hydrogen bonding and second, the peaks are generally
broadened. Keep these two factors in mind as you examine the following spectra, regardless
of what atoms and functional groups are involved in the hydrogen bonding.
The N-H stretching frequency is most frequently encountered in amines and amides.
The following examples will illustrate the behavior of this functional group in a variety of
circumstances.

82
Primary amines and amides derived from ammonia

The N-H stretching frequency in primary amines and in amides derived from ammonia
have the same local symmetry as observed in CH2. Two bands, a symmetric and an
asymmetric stretch are observed. It is not possible to assign the symmetric and
asymmetric stretches by inspection but their presence at approximately 3300 and 3340
cm–1 are suggestive of a primary amine or amide. These bands are generally broad and a
third peak at frequencies lower than 3300 cm–1, presumably due to hydrogen bonding, is
also observed. This is illustrated by the spectra in Figures IR-18 and 19 for n-butyl amine
and benzamide.

Secondary amines and amides

Secondary amines and amides show only one peak in the infrared. This peak is generally in
the vicinity of 3300 cm–1. This is illustrated in Figures IR-20 and 21. Again notice the effect
of hydrogen bonding on the broadness of the N-H peak.

Tertiary amines and amides

Tertiary amines and amides from secondary amines have no observable N-H stretching band
as is illustrated in Figures IR-22 and 23.

N-H bending motions

You may recall that we will be ignoring most bending motions because these occur in the
fingerprint region of the spectrum. One exception is the N-H bend that occurs at about
1600 cm–1. This band is generally very broad and relatively weak. Since many other
important bands occur in this region it is important to note the occurrence of this
absorption lest it be mistakenly interpreted as another functional group. Figure IR-18
illustrates the shape and general intensity of the bending motion. Most other functional
groups absorbing in this region are either sharper or more intense.

Hydroxyl Stretch

The hydroxyl stretch is similar to the N-H stretch in that it hydrogen bonds but does so
more strongly. As a result it is often broader than the N-H group. In those rare instances
when it is not possible to hydrogen bond, the stretch is found as a relative weak to
moderate absorption at 3600-3650 cm–1. In tri-t-butylmethanol where steric hindrance
prevents hydrogen bonding, a peak at 3600 cm–1 is observed as shown in Figure 17.
Similarly for hexanol, phenol, and hexanoic acid, Figures IR-24, 25, and 26, gas phase
and liquid phase spectra illustrate the effect of hydrogen bonding on both the O-H stretch
and on the rest of the spectrum. In should be pointed out that, in general, while gas phase
spectra are usually very similar, frequencies are generally shifted to slightly higher
values in comparison to condensed phase spectra. Gas phase spectra that differ

83
significantly from condensed phase spectra are usually taken as evidence for the presence
of some sort of molecular association in the condensed phase.

The hydroxyl group in phenols and alcohols usually is found as a broad peak
centered at about 3300 cm–1 in the condensed phase as noted above and in the additional
examples of Figures IR-24, 28, and 29. The O-H of a carboxylic acid, so strongly
associated that the O-H absorption in these materials, is often extended to approximately
2500 cm–1. This extended absorption is clearly observed in Figures IR-26, 27, and 29 and
serves to differentiate the O-H stretch of a carboxylic acid from that of an alcohol or
phenol. In fact, carboxylic acids associate to form intermolecular hydrogen bonded dimers
both in the solid and liquid phases.

The nitrile group

The nitrile group is another reliable functional group that generally is easy to identify.
There is a significant dipole moment associated with the CN bond which leads to a
significant change when it interacts with infrared radiation usually leading to an intense
sharp peak at 2200-2280 cm–1. Very few other groups absorb at this region with this
intensity. The spectrum in Figure IR-31 illustrates the typical behavior of this functional
group. If another electronegative atom such as a halogen is attached to the same carbon
as the nitrile group, the intensity of this is markedly reduced.

The carbon-carbon triple bond

The CC bond is not considered to be a very reliable functional group. This stems in part
by considering that the reduced mass in equation 7 is likely to vary. However it is
characterized by a strong force constant and because this stretching frequency falls in a
region where very little else absorbs, 2100-2260 cm–1, it can provide useful information.
The terminal carbon triple bond (CC-H) is the most reliable and easiest to identify. We
have previously discussed the C-H stretching frequency; coupled with a band at 3300
cm–1, the presence of a band at approximately 2100 cm–1 is a strong indication of the -C
C-H group. The spectrum in Figure IR-13 illustrates the presence of this group.

An internal -CC- is more difficult to identify and is often missed. Unless an


electronegative atom such as nitrogen or oxygen is directly attached to the sp hybridized
carbon, the dipole moment associated with this bond is small; stretching this bond also
leads to a very small change. In cases where symmetry is involved, such as in diethyl
acetylenedicarboxylate, Figure IR-32, there is no change in dipole moment and this
absorption peak is completely absent. In cases where this peak is observed, it is often
weak and difficult to identify with a high degree of certainty.

The carbonyl group

The carbonyl group is probably the most ubiquitous group in organic chemistry. It comes in
various disguises. The carbonyl is a polar functional group that frequently is the most intense
peak in the spectrum. We will begin by discussing some of the typical acyclic aliphatic

84
molecules that contain a carbonyl group. We will then consider the effect of including a
carbonyl as part of a ring and finally we will make some comments of the effect of
conjugation on the carbonyl frequency.

Acyclic aliphatic carbonyl groups

Esters, aldehydes, and ketones

Esters, aldehydes, and ketones are frequently encountered examples of molecules exhibiting
a C=O stretching frequency. The frequencies, 1735, 1725, 1715 cm-1 respectively, are too
close to allow a clear distinction between them. However, aldehydes can be distinguished by
examining both the presence of the C-H of an aldehyde (2750, 2850 cm–1) and the presence
of a carbonyl group. Examples of an aliphatic aldehyde, ester, and ketone are given in
Figures IR-14, 34, 36, and 35, respectively.

Carboxylic acids, amides and carboxylic acid anhydrides

Carboxylic acids, amides and carboxylic acid anhydrides round out the remaining
carbonyl groups frequently found in aliphatic molecules. The carbonyl frequencies of
these molecules, 1700-1730 cm–1 (carboxylic acid), 1640-1670 cm–1 (amide) and 1800-
30, 1740-1775 cm–1 (anhydride), allow for an easy differentiation when the following
factors are also taken into consideration.

A carboxylic acid can easily be distinguished from all the carbonyl containing
functional groups by noting that the carbonyl at 1700-1730 cm–1 is strongly hydrogen
bonded and broadened as a result. In addition it contains an O-H stretch which shows
similar hydrogen bonding as noted above. Spectra which illustrate the effect of hydrogen
bonding include Figures IR-27, and 29.

Amides are distinguished by their characteristic frequency which is the lowest


carbonyl frequency observed for an uncharged molecule, 1640-1670 cm–1 (Amide I). In
addition, amides from ammonia and primary amines exhibit a weaker second band
(Amide II) at 1620-1650 cm–1 and 1550 cm–1 respectively, when the spectra are run on
the solids. Amides from secondary amines do not have a hydrogen attached at nitrogen
and do not show an Amide II band. The Amide I band is mainly attributed to the
carbonyl stretch. The Amide II involves several atoms including the N-H bond. We will
return to the frequency of the amide carbonyl when we discuss the importance of
conjugation and the effect of resonance on carbonyl frequencies. The spectra of
benzamide, a conjugated amide (Figure IR-19), and N-methyl acetamide (Figure IR-21)
clearly identify the Amide I and II bands. The spectrum of N,N dimethyl acetamide
(Figure IR-23) illustrates an example of an amide from a secondary amine.

Anhydrides can be distinguished from other simple carbonyl containing


compounds in that they contain and exhibit two carbonyl frequencies. However, these
frequencies are not characteristic of each carbonyl. Rather they are another example of
the effects of local symmetry similar to what we have seen for the CH2 and NH2 groups.

85
The motions involved here encompass the entire anhydride (-(C=O)-O-(O=C-) in a
symmetric and asymmetric stretching motion of the two carbonyls. The two carbonyl
frequencies often differ in intensity. It is not possible to assign the peaks to the
symmetric or asymmetric stretching motion by inspection nor to predict the more intense
peak. However, the presence of two carbonyl frequencies and the magnitude of the
higher frequency (1800 cm–1) are a good indication of an anhydride. Figure IR-36
contains a spectrum of an aliphatic anhydride.

Cyclic aliphatic carbonyl containing compounds

The effect on the carbonyl frequency as a result of including a carbonyl group as part of a
ring is usually attributed to ring strain. Generally ring strain is believed to be relieved in large
rings because of the variety of conformations available. However as the size of the ring gets
smaller, this option is not available and a noticeable effect is observed. The effect of
increasing ring stain is to increase the carbonyl frequency, independent of whether the
carbonyl is a ketone, part of a lactone(cyclic ester), anhydride or lactam (cyclic amide). The
carbonyl frequencies for a series of cyclic compounds is summarized in Table 3.

Table 3. The Effect of Ring Strain on the Carbonyl Frequencies of Some Cyclic
Molecules

Ring ketone: cm-1 lactones: cm-1 lactams: cm-1


Size
3 cyclopropanone: 1800
4 cyclobutanone: 1775 -propiolactone: 1840
5 cyclopentanone: 1751 -butyrolactone: 1750 -butyrolactam: 1690
6 cyclohexanone: 1715 -valerolactone: 1740 -valerolactam: 1668
7 cycloheptanone: 1702 caprolactone: 1730 caprolactam: 1658

Carbon carbon double bond

Like the CC bond, the C=C bond stretch is not a very reliable functional group.
However, it is also characterized by a strong force constant and because of this and
because the effects of conjugation which we will see can enhance the intensity of this
stretching frequency, this absorption can provide useful and reliable information.

Terminal C=CH2

In simple systems, the terminal carbon carbon double bond (C=CH2) is the most reliable
and easiest to identify since the absorption is of moderate intensity at 1600-1675 cm-1.
We have previously discussed the C-H stretching frequency of an sp2 hybridized C-H.
The spectrum in Figure IR-9 illustrates the presence of this group. In addition the
terminal C=CH2 is also characterized by a strong band at approximately 900 cm–1. Since
this band falls in the fingerprint region, some caution should be exercised in its
identification.

86
Internal C=C

An internal non-conjugated C=C is difficult to identify and can be missed. The dipole
moment associated with this bond is small; stretching this bond also leads to a very small
change. In cases where symmetry is involved, such as in 4-octene, Figure IR-10, there is no
change in dipole moment and this absorption peak is completely absent. In cases where this
peak is observed, it is often weak. In 2,5-dihydrofuran, Figure IR-39, it is difficult to assign
the C=C stretch because of the presence of other weak peaks in the vicinity. The band at
approximately 1670 cm–1 may be the C=C stretch. In 2,5-dimethoxy-2,5-dihydrofuran,
Figure IR-40, the assignment at 1630 cm-1 is easier but the band is weak.

There is one circumstance that can have a significant effect on the intensity of
both internal and terminal olefins and acetylenes. Substitution of a heteroatom directly on
the unsaturated carbon to produce, for example, a vinyl or acetylenic ether, or amine
leads to a significant change in the polarity of the C=C or CC bond and a substantial
increase in intensity is observed. The C=C in 2,3-dihydrofuran is observed at 1617.5 cm–
1
and is one of the most intense bands in the spectrum (Figure IR-41). Moving the C=C
bond over one carbon gives 2,5-dihydrofuran attenuates the effect and results in a weak
absorption (Figure IR-39).

Aromatic ring breathing motions

Benzene rings are encountered frequently in organic chemistry. Although we may write
benzene as a six membered ring with three double bonds, most are aware that this is not a
good representation of the structure of the molecule. The vibrational motions of a
benzene ring are not isolated but involve the entire molecule. To describe one of the
fundamental motions of benzene, consider imaginary lines passing through the center of
the molecule and extending out through each carbon atom and beyond. A symmetric
stretching and compression of all the carbon atoms of benzene along each line is one
example of what we might describe as a ring breathing motion. Simultaneous expansions
and compressions of these six carbon atoms lead to other ring breathing motions. These
vibrations are usually observed between 1450 and 1600 cm–1 and often lead to four
observable absorptions of variable intensity. As a result of symmetry, benzene, Figure
IR-7, does not exhibit these bands. However most benzene derivatives do and usually 2
or 3 of these bands are sufficiently separate from other absorptions that they can be
identified with a reasonable degree of confidence. The least reliable of these bands are
those observed at approximately 1450 cm–1 where C-H bending motions are observed.
Since all organic molecules that contain hydrogen are likely to have a C-H bond,
absorption observed at 1450 cm–1 is not very meaningful and should usually be ignored.
Two of the four bands around 1600 cm–1 are observed in ortho and meta xylene,
identified by the greek letter  and a third band at about 1500 cm–1 is assigned (Figure
IR-11 and 12). We will return to a discussion of these bands when we discuss the effects
of conjugation on the intensities of these motions.

Nitro group

87
The final functional group we will include in this discussion is the nitro group. In addition to
being an important functional group in organic chemistry, it will also begin our discussion of
the importance of using resonance to predict effects in infrared spectroscopy. Let's begin by
drawing a Kekule or Lewis structure for the nitro group. You will find that no matter what
you do, it will be necessary to involve all 5 valence

+ O- + O
N N

O O -
electrons of nitrogen and use them to form the requisite number of bonds to oxygen. This
will lead to a positive charge on nitrogen and a negative charge on one oxygen. As a
result of resonance, we will delocalize the negative charge on both oxygens and as
shown, this leads to an identical structure. Since the structures are identical, we would
expect the correct structure to be a resonance hybrid of the two. In terms of geometry, we
would expect the structure to be a static average of the two geometric structures both in
terms of bond distances and bond angles. Based on what we observed for the CH2 and
NH2 stretch, we would expect a symmetric and an asymmetric stretch for the N-O bond
in the nitro group halfway between the N=O and N-O stretches. Since both of those
functional groups are not covered in this discussion, we will need to assume for the
present that this is correct. Two strong bands are observed, one between 1500 and 1600
cm–1 and a second between 1300 and 1390 cm–1, Figure IR-28.

Effect of resonance and conjugation on infrared frequencies

Let's continue our discussion of the importance of resonance but shift from the nitro group to
the carboxylate anion. The carboxylate anion is represented as a resonance hybrid by the
following figure:
O- O
C C

O O -
Unlike the nitro group which contained functional groups we will not be discussing, the
carboxyl group is made up of a resonance hybrid between a carbon oxygen single bond
and a carbon oxygen double bond. According to resonance, we would expect the C-O
bond to be an average between a single and double bond or approximately equal to a
bond and a half. We can use the carbonyl frequency of an ester of 1735 cm–1 to describe
the force constant of the double bond. We have not discussed the stretching frequency of
a C-O single bond for the simple reason that it is quite variable and because it falls in the
fingerprint region. However the band is known to vary from 1000 to 1400 cm–1. For
purposes of this discussion, we will use an average value of 1200 cm-1. The carbonyl
frequency for a bond and a half would be expected to fall halfway between 1735 and
1200 or at approximately 1465 cm–1. The carboxyl group has the same symmetry as the

88
nitro and CH2 groups. Both a symmetric and asymmetric stretch should be observed. The
infrared spectrum of sodium benzoate is given in Figure IR-42. An asymmetric and
symmetric stretch at 1410 and 1560 cm–1 is observed that averages to 1480 cm–1, in good
agreement with the average frequency predicted for a carbon oxygen bond with a bond
order of 1.5. While this is a qualitative argument, it is important to realize that the
carboxylate anion does not show the normal carbonyl and normal C-O single bond
stretches (at approximately 1700 and 1200 cm–1) suggested by each of the static
structures above.

In the cases of the nitro group and the carboxylate anion, both resonance forms contribute
equally to describing the ground state of the molecule. We will now look at instances where
two or more resonance forms contribute unequally to describing the ground state and how
these resonance forms can effect the various stretching frequencies.

Carbonyl frequencies

Most carbonyl stretching frequencies are found at approximately 1700 cm–1. A notable
exception is the amide carbonyl which is observed at approximately 1600 cm–1. This
suggests that the following resonance form makes a significant contribution to describing
the ground state of amides:
+
NR2 NR2
C C

O O -
You may recall that resonance forms that lead to charge separation are not considered to
be very important. However the following information support the importance of
resonance in amides. X-ray crystal structures of amides show that in the solid state the
amide functional group is planar. This suggests sp2 hybridization at nitrogen rather than
sp3. In addition the barrier to rotation about the carbon nitrogen bond has been measured.
Unlike the barrier of rotation of most aliphatic C-N bonds which are of the order of a few
kcal/mol, the barrier to rotation about the carbon nitrogen bond in dimethyl formamide is
approximately 18 kcal/mol. This suggests an important contribution of the dipolar
structure to the ground state of the molecule and the observed frequency of 1600 cm–1,
according to the arguments given above for the carboxylate anion, is consistent with
more C-O single bond character than would be expected otherwise.

Conjugation of a carbonyl with a C=C bond is thought to lead to an increase in


resonance interaction. Again the resonance forms lead to charge separation which clearly de-
emphasizes their importance.

O O-

89
However this conjugative interaction is useful in interpreting several features of the
spectrum. First it predicts the small but consistent shift of approximately 10 cm-1 to lower
frequency, observed when carbonyls are conjugated to double bonds or aromatic rings.
This feature is summarized in Table 4 for a variety of carbonyl groups. Next, the dipolar
resonance form suggests a more polar C=C than that predicted for an unconjugated C=C.
In terms of the change in dipole moment, contributions from this structure suggest that
the intensity of infrared absorption of a C=C double bond would increase relative to an
unconjugated system. Comparison of Figures IR-9, 10 and 35 with Figures IR-43, and
44-47 shows this to be the case. Conjugation is associated with an increase in intensity of
the C=C stretching frequency. Finally, examination of Figures IR-43-46 reveals an
intricacy not previously observed with simple non-conjugated carbonyls. The carbonyls
of Figures 43-46 which are all conjugated appear as multiplets while those unconjugated
carbonyls such as those in Figures IR-14 and 35 appear as single frequencies. Note
however that not all conjugated carbonyls appear as multiplets (Figures IR-15 and 47.
Resolution of this additional complicating feature can be achieved if we consider that
conjugation requires a fixed conformation. For most conjugated carbonyls, two or more
conformations are possible. The s-cis form is shown below on the right and the s-trans
form is shown on the left.
+
+

O- O-

If the resonance interaction in these two forms differ, the effect of resonance on the
carbonyl will differ leading to similar but different frequencies. The presence of multiple
carbonyl frequencies is a good indication of a conjugated carbonyl. In some conjugated
systems such as benzaldehyde and benzyl 4-hydroxyphenylketone (Figures IR-15 and
47), only one conformation by symmetry is possible and conjugation does not lead to any
additional carbonyl frequencies.

90
Table 4. The effect of conjugation on carbonyl frequencies.

Non-conjugated Frequency Conjugated Frequency Frequency


Compound cm-1 Compound cm-1 cm-1

butanal 1725 2-butenal 1691 benzaldehyde 1702

2-butanone 1717 methyl vinyl 1700, acetophenone 1685


ketone 1681

propanoic acid 1715 propenoic acid 1702 benzoic acid 1688

ethyl propionate 1740 ethyl acrylate 1727 ethyl benzoate 1718

butanoic anhydride 1819, 2-butenoic 1782, benzoic 1786,


1750 anhydride 1722 anhydride 1726

cis-cyclohexane- 1857, 1-cyclohexene- 1844, phthalic 1852,


1,2-dicarboxylic 1786 1,2-dicarboxylic 1767 anhydride 1762
anhydride anhydride

It should also be noted that in many of the examples given above, cis-trans
isomerization about the carbon-carbon double bond is also possible. Some of the
observed bands may also be due to the presence of these additional isomers. Since the
intensity of the peak is determined by the change in dipole moment, the presence of a
small amount of geometric isomer can still lead to a detectable peak.

Experimental infrared spectra

Up to now we have been focusing in on theory and interpretation of infrared spectra. At this
point we should spend some time discussing the practical aspects of how infrared spectra
have been obtained and how they are currently obtained. The method used plays an important
factor that must be considered when trying to interpret the results. Infrared spectroscopy has
been an important tool for the organic chemist since the 1960's. During this period of time
spectra have been recorded using various techniques which have changed with time. Since
the method used to obtain a spectrum influences the spectrum, some familiarity with the
different techniques is necessary if your are interested in using existing IR spectra to confirm
the identity of a pure chemical substance.

Let's first start by considering gas phase spectra.

Cells and gas phase spectra

These type of spectra were more a curiosity and of theoretical interest until the introduction
of the combined techniques of gas chromatography-Fourier transfer infrared spectroscopy
(GC-FTIR). The major advantages of this method is that spectra can be obtained on

91
micrograms of material and the spectra do not show the effects of interactions between
molecules characteristic of condensed phase spectra. These spectra are usually obtained at
elevated temperatures. Condensed phase spectra however will continue to be important
because of the fact that many compounds do not survive injection into a gas chromatograph.
Currently, most frequency correlations for various functional groups are reported for the
condensed phase. Frequencies observed in the gas phase are usually slightly higher than
those observed for the same functional group in the condensed phase.

Gas phase spectra can also be taken at room temperature. All that is needed is a
sample with a vapor pressure of several millimeters and a pathlength of about a
decimeter (10 cm). Cells with NaCl or KBr windows are commercially available or can
be built easily. Crystals of KBr are transparent from 4000-250 cm–1 and are perfectly
acceptable for most uses. KBr has the disadvantage of being hydroscopic and must be
stored in a desiccator. Cells of sodium chloride are transparent from 4000-600 cm–1, less
expensive and less hydroscopic. These cells are also acceptable for routine spectra.

Cells and condensed phase spectra

Condensed phase spectra can be taken as a solid or as a liquid. Comparison of the same
sample in the liquid and solid phase will differ. However the major differences observed will
be in the fingerprint region. In cases where infrared spectroscopy is used as a criteria of
identity, the spectra under comparison should be obtained under identical experimental
conditions. Liquid phase spectra are the easiest to obtain. All that is needed are two polished
disks of NaCl or KBr, both commercially available. A thin film is prepared by depositing a
drop of the liquid between the two plates and mounting them in the beam of the
spectrometer. This is referred to as a neat liquid. Glass is not a useful material in infrared
spectroscopy because of the strong absorptions due to the Si-O group. The infrared spectrum
of quartz is shown in Figure 49.

Spectra of solids can be obtained in a variety of ways. The method of choice varies
depending on the physical properties of the material under consideration. We will list several
methods that can be used satisfactorily along with the limitations and advantages of each.

Neat Spectra (thin film)

In order to obtain an infrared spectrum of a solid, it is necessary to get light, mainly


infrared, through the sample. This can be achieved in various ways and we will outline
some that have proven successful in the past. A thin layer of a solid deposited as a
solution on an infrared cell and the solvent allowed to evaporate has proven successful
with many compounds. Solvents such as CHCl3, CH2Cl2 and CCl4 have been frequently
been used. The solid sample should have an appreciable solubility in one of these
solvents. A drop of a solution left to evaporate will deposit a thin film of crystal that will
often transmit sufficient light to provide an acceptable infrared spectrum. This method
suffers from the disadvantage that a spectrum of the solvent must also be run to
determine whether all of the solvent has evaporated.

92
Nujol mull

A mull is a suspension of a solid in a liquid. Under these conditions, light can be transmitted
through the sample to afford an acceptable infrared spectrum. The commercial sample of
Nujol, or mineral oil, which is a long chain hydrocarbon is often used for this purpose. Most
solids do not dissolve in this medium but can be ground up in its presence. A small mortar
and pestle is used for this purpose. If the grinding process gives rise to small particles of
solid with diameters roughly the same as the wavelength of the infrared radiation being used,
2-5 microns, these particles will scatter rather than transmit the light. The effect of poor
grinding is illustrated in Figures IR-29 and 30 for a sample of benzoic acid. If you find this
type of band distortion with either a Nujol mull or a KBr pellet (discussed below), simply
continue grinding the sample up until the particles become finer.

The major disadvantage of using a Nujol mull is that the information in the C-H
stretching region is lost because of the absorptions of the mulling agent. A spectrum of Nujol
is shown in Figure IR-5. To eliminate this problem, it may be necessary to run a second
spectrum in a different mulling agent that does not contain any C-H bonds. Typical mulling
agents that are used for this purpose are perfluoro- or perchlorohydrocarbons. Examples
include perchlorobutadiene, perfluorokerosene or a perfluorohydrocarbon oil (Figure IR-48).

The use of Nujol mulls has decreased in recent years. This has been due to the
introduction of a technique referred to at attenuated total reflectance or ATR. This technique
which you will be using for solids is described below.

KBr pellets

A KBr pellet is a dilute suspension of a solid in a solid. It is usually obtained by first


grinding the sample in anhydrous KBr at a ratio of approximately 1 part sample to 100 parts
KBr. Although it is best to weigh the sample (1 mg) in the KBr (100mg), with some
experience it is possible to use your judgment in assigning proportions of sample to KBr. The
mixture is then ground up in an apparatus called a Wiggle-Bug, frequently used by dentists to
prepare amalgams. The ground up sample mixture is then placed on a steel plate containing a
paper card with a hole punched in it. The sample is placed in the hole, making sure that some
sample also overlaps the paper card. Paper the thickness and consistency of a postcard is
usually used and the hole is positioned on the card so that it will lie in the infrared beam
when placed on the spectrometer. A second steel plate is placed over the sample and card and
the steel sandwich is placed in a hydraulic press and subjected to pressures of 15000 psi for
about 20 seconds. Removal of the paper card following decompression usually results in a
KBr pellet that is reasonably transparent both to visible light and infrared radiation. Some
trial and error may be necessary before quality pellets can be obtained routinely. Samples
that are not highly crystalline sometimes prove difficult and do not produce quality pellets.
However good quality spectra can be obtained on most samples. The only limitation of KBr
is that it is hydroscopic. Because of this, it is usually a good idea to obtain a spectrum run as
a Nujol mull on your sample as well. The two spectra should be very similar and since Nujol

93
is a hydrocarbon and has no affinity for water, any absorption in Nujol between 3400-3600
cm–1 can be attributed to the sample and not to the absorption of water by KBr.

The use of KBr pellets has been the method of choice for preparing solids for IR
analysis. This method has been used a long period of time and the infrared spectra of many
solids are available in the literature as KBr pellets. You should be aware that because KBr is
hydroscopic, an hydrogen bonded OH stretch is usually detectable in these spectra. At
present, ATR has largely replaced KBr.

ATR

Attenuated total reflectance is the technique used presently for obtaining spectra of solids.
The ATR apparatus is remarkably easy to use and the spectra obtained are almost as good as
the best spectra obtained by KBr. The technique operates on a very simple but perhaps
surprising observation. When light is reflected off a surface, a portion of the wave actually
penetrates about a wavelength into the medium it is being reflected off of. For IR, this is in
the micron range, 10-6 m. The IR that penetrates before being reflected can be absorbed by
the medium or not, depending on its wavelength. Since present day detectors are sufficiently
sensitive, and because a FTIR normally averages a variable number of scans, spectra of
remarkable quality can be obtained by ATR. The ATR attachment uses a zinc selenide
crystal to focus the light and a diamond anvil on which the solid sample is compressed to
obtain sufficient loading. Diamond does absorb in the IR and this is the reason why ATR is
not usually used for liquids. Sodium chloride or KBr give better quality spectra since they do
not absorb in the regions of interest. A background scan that includes absorption by water,
CO2 and the diamond anvil is illustrated in Figure IR-50

ATR Correction

As noted above, a portion of the light actually penetrates about a wavelength into the
medium it is being reflected off. This means that light of longer wavelength penetrates
deeper than light of shorter wavelength. IR absorption follows Beer's law (i.e. absorbance
depends on path length). Therefore, it is important to correct for the wavelength dependence
of absorption.

Nicolet AVATAR 360 FT Infrared Spectrometer

The operation of the Avatar FTIR, the instrument that you will be using in this laboratory,
will be demonstrated. It is remarkably easy to use. However before using it you should
familiarize yourself with some of the general operating features of an FTIR instrument and
its capabilities and limitation. A discussion of the performance of a Fourier Transfer infrared
spectrometer is beyond the scope of this publication. However the following will summarize
some of the essential features of an FTIR. To begin with the AVATAR 360 is a single beam
instrument. Unlike a double beam instrument that simultaneously corrects for absorptions
due to atmospheric water vapor and carbon dioxide, most FTIR spectrometers correct for the
background absorption by storing an interferogram and background spectrum before
recording your spectrum. An interferogram contains the same information as a regular

94
spectrum, frequency vs. intensity, but this information is contained in the form of intensity
vs. time. The Fourier Tranform is the mathematical process which converts the information
from intensity and time to intensity and frequency. One of the major advantages of a FTIR
instrument is that is takes on the order of a second to record an entire spectrum. This makes it
very convenient to record the same spectrum a number of times and to display an averaged
spectrum. Since the signal to noise ratio varies as the square root of the number of scans
averaged, it is easy to obtain good signal to noise on this type of instrument, even if you have
very little sample.

The instrument is accessed by using a dedicated computer. The instrument is usually


left on. However if it has been turned off, it is usually a good idea to allow the infrared
source to warm up for 5 min. With the sample pathway empty (when using the ATR
accessory, the stainless steel, arm should be down, just as if you were running a regular
spectrum but without any sample), a background should be run. This spectrum is stored
internally so that it is not necessary to do anything else with the spectrum. You do not need
to "add to window". The sample is next placed in the beam (as demonstrated), the mouse is
used to click on collect sample, abbreviated as Col Smp. The instrument will record your
spectrum. Usually 16 scans are averaged. When the task is completed, you will be prompted
to "Add to window". If you wish to have a hard copy of the spectrum you should respond
"yes" with the mouse. The instrument normally reports spectra in adsorption mode. Clicking
the % trans icon will convert your spectrum to transmittance. Clicking on the print icon will
send your spectrum to the printer. If more than 1 spectrum is present on the screen, the active
one will be in red. Clinking on another spectrum will activate that spectrum. Clicking on
Clear will remove the active spectrum.

To correct you spectrum for the ATR wavelength dependence discussed above, while
your spectrum is in adsorption mode, click on "process", "other corrections", "ATR". This
will adjust the wavelength dependence of your spectrum accordingly.

Interpretation of Infrared Spectra

We have just concluded a discussion of a large number of frequencies and the functional
groups that are generally associated with these frequencies. At this point you may be asking
yourself how to begin to interpret these frequencies with regards to obtaining information of
molecular structure. There are a number of different approaches that can be used and often
the best approach to use depends on the nature of the information you would like to obtain
from your infrared spectrum. For example, if you are repeating a synthesis in the laboratory
and you wish to determine whether you have successfully isolated the material you intended
to prepare, you may be able to compare your spectrum to an infrared spectrum of an
authentic sample. In this case, you are using infrared analysis for establishing the identity of
your sample. Assuming that your spectrum has been run under the same conditions, as your
reference, i.e. neat sample, KBr pellet, etc., you should be able to reproduce the spectrum of
the reference material, peak for peak. The presence of some additional peaks in your
spectrum may indicate a contamination with solvent, starting material or an impurity that has
not been removed. The presence of fewer peaks than your reference is of more concern. This
generally indicates a failure to obtain the desired material.

95
If the structure of the material of interest is unknown, then a more systematic analysis
of your spectrum will be necessary. You should be aware that it is not usually possible to
determine molecular structure from the infrared spectrum alone. Usually, some supplemental
spectroscopic and/or structural information (such as molecular formula) is also necessary.
For the unknowns in this course, you will generally be using infrared spectroscopy to
differentiate between a few possible compounds. Frequently, this can be achieved by an
analysis of the functional groups in your spectrum. The discussion which follows, uses a
more generalized approach to analyze spectra. This approach should be applicable in a
variety of different circumstances. If a portion of the discussion is not relevant to you, simply
skip it and continue until it does become relevant.

Information from Molecular Formulas


The Degree of Unsaturation

Once the molecular formula of an unknown is known, it is a simple matter to determine the
degree of unsaturation. The degree of unsaturation is simply the sum of the number of
carbon-carbon double bonds and rings. Each reduces the number of hydrogens or any other
element with a valance of one by two. Although there is a general formula that can be
memorized and used, a much simpler procedure is to note the number of carbon atoms and
any other elements in your molecular formula and simply draw a molecule that contains the
requisite number of carbons atoms and any other elements that are present. Make sure there
are no rings or carbon-carbon double bonds in your structure. Each carbon, nitrogen, sulfur
and oxygen should have four, three, two and two single bonds, respectively. Use as many
hydrogens as you need to make up the appropriate number of bonds for each element. Be
sure to include all halogens and any other elements in your structure as well. Count the
number of hydrogens in your structure and subtract this number from the number in your
original molecular formula. The difference, divided by two equals the degree of unsaturation,
the sum of the number of rings and double bonds.

Consider C6Cl6 as an example: CH2Cl-CHCl-CHCl-CHCl-CHCl-CH2Cl. The number


of hydrogens in my sample molecule is 8; there are none in the original molecular
formula. The difference, divided by two is four. The degree of unsaturation is four. An
unsaturation factor of four is quite common and characteristic of benzene and its
derivatives.

Application of the degree of unsaturation to the interpretation of an infrared spectrum


is quite straightforward. Clearly some functional groups can be eliminated by composition.
Amines, amides, nitriles and nitro groups can be eliminated if the molecule does not contain
any nitrogen. Alternatively everything but amines can be eliminated if the molecular formula
contains nitrogen and no degrees of unsaturation. The following steps should serve as a
general protocol to follow and should prove useful regardless of the structure of your
unknown or whether the degree of unsaturation is known.

1. Examine the C-H stretching frequencies at 3000 cm–1. Absorption bands bands at
frequencies slightly larger than 3000 cm–1 are indicative of vinyl or/and aromatic

96
hydrogens. The presence of these peaks should be consistent with the degree of
unsaturation of your molecule. The absence of absorption above 3000 cm–1 but the
presence of some unsaturation in the molecular formula are consistent with a cyclic
compound. If your degree of unsaturation is 4 or greater, look for 2 to 4 absorption peaks
between 1600-1450 cm–1 and weak peaks at 2000-1667 cm-1. These are characteristic of
aromatic compounds.

2. Next look for a doublet at 2750 and 2850 cm–1 characteristic of an aldehyde. The
presence of these two bands should also be accompanied by a strong absorption at
approximately 1700 cm–1. Most spectra display strong absorption in the 1800-1700 cm–1
region. If your spectrum does, check to see if the carbonyl is a closely spaced doublet or
multiplet. Closely spaced multiplicity in the carbonyl region accompanied by C-H
absorption at 3000-3100 cm–1 is frequently characteristic of an ,- unsaturated
carbonyl compounds. Check to make sure that the carbonyl frequency is consistent with
conjugation.

3. If you unknown contains broad absorption from 3600-3000 cm–1, your molecule could
have an O-H or N-H stretch. Your molecular formula may allow you to differentiate. Check
the multiplicity of this peak. A doublet is characteristic of a primary amine or and amide
derived from ammonia. Check the carbonyl region at around 1650-1600 cm–1. Two bands in
this region are consistent of an amide from ammonia or a primary amine. Remember a broad
and relatively weak band at about 1600 cm–1 is characteristic of N-H bending. Usually you
will only see this band in amines, since that carbonyl group of the amide will interfere. Be
sure to look for the effect of hydrogen bonding which usually results in a general broadening
of the groups involved.

4. If the broad band starting at 3600 expands to nearly 2400 cm–1, look for the presence
of a broad carbonyl at approximately 1700 cm–1. This extremely broad OH band is only
observed in carboxylic acids and enols from -diketones. The presence of a relatively
intense but broad band at approximately 1700 cm–1 is good evidence for a carboxylic
acid.

5. Don’t try to over-interpret your spectrum. Often, it is not possible to arrive at a unique
structure based on infrared analysis alone. You should use your infrared analysis much like
you would use other classification tests. You can learn a great deal about your unknown from
your spectrum but be sure to use other important physical data such as melting point, boiling
point and solubility characteristics of your unknown to assist you in narrowing down the
different structural possibilities.

97
APPENDIX 1. Table of IR Spectra
Figure 5. Nujol
Figure 6. n-Decane
Figure 7. Benzene
Figure 8. Indene
Figure 9. 1-Heptene
Figure 10. trans -4-Octene
Figure 11. ortho-Xylene
Figure 12. meta-Xylene
Figure 13. Phenylacetylene
Figure 14. Butanal
Figure 15. Benzaldehyde
Figure 16. Chloroform
Figure 17. Tri-t-butylmethanol
Figure 18. Butylamine
Figure 19. Benzamide
Figure 20. Diethylamine
Figure 21. N-Methyl acetamide
Figure 22. Triethylamine
Figure 23. N,N-Dimethyl acetamide
Figure 24. n-Hexanol, vapor and liquid
Figure 25. Phenol, vapor and liquid
Figure 26. Hexanoic acid, vapor and liquid
Figure 27. Decanoic acid
Figure 28. 4-Chloro-2-nitrophenol
Figure 29. Benzoic acid
Figure 30. Benzoic acid
Figure 31. trans-2-Phenyl-1-cyanoethene
Figure 32. Diethyl acetylenedicarboxylate
Figure 33. Methyl acetate
Figure 34. Ethyl n-hexanoate, vapor and liquid
Figure 35. 5-Hexene-2-one
Figure 36. Propionic anhydride
Figure 37. Benzoic anhydride
Figure 38. 3-Methylpimelic acid anhydride
Figure 39. 2,5-Dihydrofuran
Figure 40. 2,5-Dimethoxy-2,5-dihydrofuran
Figure 41. 2,3-Dihydrofuran
Figure 42. Sodium benzoate
Figure 43. 5-Methyl-3-hexene-2-one
Figure 44. 3-Nonen-2-one
Figure 45. Ethyl vinyl ketone
Figure 46. 3-Penten-2-one
Figure 47. Benzyl 4-hydroxyphenyl ketone

98
Figure 48. Perfluorohydrocarbon oil
Figure 49. Quartz
Figure 50. ATR background

99
100
90
80
70
% Transmittance

60
50
40
30
20
10
4000 3500 3000 2500 2000 1500 1000 500
-1
Wavenumbers, cm

Figure IR-5. Nujol; thin film (Nujol is a trade name for mineral oil, a long chain alkane)

100

80
% Transmittance

60
(CH2)4
rock
40

CH3, CH2 symmetric CH3, CH2 bending


20
and asymmetric stretches vibrations

0
4000 3500 3000 2500 2000 1500 1000 500
-1
Wavenumbers, cm

Figure IR-6. n-decane, neat liquid, thin film: CH3CH2CH2CH2CH2CH2CH2CH2CH2CH3

100
90

80

70
% Transmittance

60
overtones of
50 the benzene
ring
40

30

20
4000 3500 3000 2500 2000 1500 1000 500
Wavenumbers, cm-1

Figure IR-7. Benzene, neat, thin film

100

80
% Transmittance

60

1410
1020 860
40

20

0
4000 3500 3000 2500 2000 1500 1000 500
-1
Wavenumbers, cm

Figure IR-8. Indene; neat; 0.05 mm cell:

101
100

80

(CH2)4
% Transmittance

HC=C
60 stretches C=C rocking
stretch
40 CH3,CH2
bending
20 vibrations out of plane
CH3,CH2 symmetric H wag of vinyl
and asymmetric stretches group 900, 1000 cm-1)
0
4000 3500 3000 2500 2000 1500 1000 500
-1
Wavenumbers, cm

Figure IR-9. 1-Heptene; neat sample, thin film: H2C=CH-CH2CH2CH2CH2CH3

80
70
60
% Transmittance

50
40
30
20 CH3, CH2 stretching
vibrations CH3, CH2 bending
10 vibrations
0
4000 3500 3000 2500 2000 1500 1000 500

Wavenumbers, cm-1

Figure IR-10. trans -4-Octene, neat liquid, thin


film

102
Figure IR-11. ortho-Xylene, neat, thin film

103
CH3

CH3

90
80
70
% Transmittance

60
50
40 

30
20
10
0
4000 3500 3000 2500 2000 1500 1000 500
Wavenumbers, cm-1

Figure IR-12. meta-Xylene, neat, thin film:

104
100

80
% Transmittance

60
carbon-carbon
triple bond stretch
40

20

0
4000 3500 3000 2500 2000 1500 1000 500
-1
Wavenumbers, cm

Figure IR-13. Phenylacetylene, neat liquid; thin film:

100
90
80
% Transmittance

70
60 H-C=O
50 Impurity peak,
probably O-H
40
30
20
10
4000 3500 3000 2500 2000 1500 1000 500
Wavenumbers, cm-1

Figure IR-14. Butanal, neat liquid, thin film: CH3CH2CH2CHO

105
100

80
% Transmittance

60

40
C-H stretch of -CHO

20

0
4000 3500 3000 2500 2000 1500 1000 500
Wavenumbers, cm-1

Figure IR-15. Benzaldehyde, neat, thin film:

120

100
% Transmittance

80
3035 cm-1
60

40

20

0
4000 3500 3000 2500 2000 1500 1000 500
Wavenumbers, cm-1

Figure IR-16. Chloroform, thin film: CHCl3

106
100
90
80
70
% Transmittance

60
50
40
30
20
10
0
4 0 0 0 3 5 0 0 3 0 0 0 2 5 0 0 2 0 0 0 1 5 0 0 1 0 0 0 5 0 0
-1
W a v e n u m b e rs, c m

Me
Me Me
Me C
Me C OH
Me C
Me
Me Me
Figure IR-17. Tri-t-butylmethanol, KBr pellet:

100
90
80
% Transmittance

70
60 NH2 bend

50
40 symmetric and
asymmetric NH2
30 stretches
20
10
4000 3500 3000 2500 2000 1500 1000 500
Wavenumbers, cm-1

Figure IR-18. Butylamine, neat liquid; thin film: CH3CH2CH2CH2NH2

107
120

100
% Transmittance

80

60

40
symmetric and
20 asymmetric NH2
stretches
0
4000 3500 3000 2500 2000 1500 1000 500
-1
Wavenumber, cm

O
CNH2

Figure IR-19. Benzamide, KBr pellet:

120

100

80
% Transmittance

60

40

20

4000 3500 3000 2500 2000 1500 1000 500


-1
Wavenumbers, cm

Figure IR-20. Diethylamine, neat liquid; thin film: CH3CH2NHCH2CH3

108
100
90
80
% Transmittance

70
60
50
40
30
20 Amide I Amide II

10
4000 3500 3000 2500 2000 1500 1000 500
-1
Wavnumbers, cm

Figure IR-21. N-Methyl acetamide, neat liquid; thin film: CH3CONHCH3

100

80
% Transmittance

60

40

20

0
4000 3500 3000 2500 2000 1500 1000 500
Wavenumbers, cm-1

Figure IR-22. Triethylamine, neat liquid; thin film: (CH3CH2)3N

109
100

80
% Transmittance

60

40

20

0
4000 3500 3000 2500 2000 1500 1000 500
Wavenumbers, cm-1

Figure IR-23. N,N-Dimethylacetamide, neat liquid; thin film: CH3CON(CH3)2

110
100
vapor 102 °C

80 liquid film
% Transmittance

60

40

20

0
4000 3500 3000 2500 2000 1500 1000 500
-1
Wavenumbers, cm

Figure IR-24. The liquid and vapor spectra of n-hexanol.

120
Vapor phase spectrum(135 °C)
100

80
% Transmittance

liquid film
60

40

20

0
4000 3500 3000 2500 2000 1500 1000
Wavenumbers, cm-1

OH

Figure IR-25. The liquid and vapor spectra of phenol.

111
120

100

80 vapor: 134 °C
% Transmittance

60

40
liquid film
20

4000 3500 3000 2500 2000 1500 1000


-1
Wavenumber, cm

Figure IR-26. The liquid and vapor spectra of hexanoic acid: CH3(CH2)4 CO2H

100

80
% Transmittance

60

40

20

0
4000 3500 3000 2500 2000 1500 1000 500
Wavenumbers, cm-1

Figure IR-27. Decanoic acid, neat liquid, thin film: CH3(CH2)8CO2H

112
120

100

80
% Transmittance

60

40

20
symmetric and asymmetric
-NO2 stretch
0
4000 3500 3000 2500 2000 1500 1000 500
Wavenumbers, cm-1

Figure IR-28. 4-Chloro-2-nitrophenol, KBr pellet

100

80
% Transmittance

60

40

20

0
4000 3500 3000 2500 2000 1500 1000 500
-1
Wavenumber cm

COOH

Figure IR-29. Benzoic acid; KBr disk:

113
100
90
80
% Transmittance

70
60
50
40
30
20
4000 3500 3000 2500 2000 1500 1000 500
-1
Wavenumbers, cm

Figure IR-30. Benzoic acid, KBr; band distortions and broadening caused by poor grinding,
compare to Figure 29.

100

80
% Transmittance

60

40

20 Top: CHCl3 vs CHCl3


Bottom: 5% in CHCl3

0
4000 3500 3000 2500 2000 1500 1000 500
-1
Wavenumbers, cm

Figure IR-31. trans-2-Phenyl-1-cyanoethene, in CHCl3 solution: Ph-CH=CH-CN

114
100

80
% Transmittance

60

40

20

0
4000 3500 3000 2500 2000 1500 1000 500
-1
2987 1370 1040 Wavenumbers, cm
1724 1255 857
1467 1095 748
Figure IR-32. Diethyl acetylenedicarboxylate; neat liquid: C2H5OCO-CC-CO2CH2CH3

100

80
% Transmittance

60

40
Top: CHCl3 vs CHCl3
20 Bottom: 2.5 % v/v un CHCl3

0
4000 3500 3000 2500 2000 1500 1000 500
-1
Wavenumbers, cm

Figure IR-33. Methyl acetate in 0.1 mm NaCl cells: CH3CO2CH3

115
120

100

80 vapor: 106 °C
% Transmittance

60

40 liquid film

20

4000 3500 3000 2500 2000 1500 1000


Wavenumber, cm-1

Figure IR-34. Vapor and liquid spectra of ethyl n-hexanoate: CH3CH2CH2CH2CH2CO2CH2CH3

100

80
% Transmittance

60

40

20

0
4000 3500 3000 2500 2000 1500 1000 500
-1
Wavenumbers, cm

Figure IR-35. 5-Hexene-2-one, neat liquid: CH2=CH2CH2CH2COCH3

116
100

80
% Transmittance

60

40

20 1819
1751

0
4500 4000 3500 3000 2500 2000 1500 1000 500
-1
Wavenumber, cm

Figure IR-36. Propionic anhydride, neat liquid: CH3CH2CO2COCH2CH3

100

80
% Transmittance

60

40

20 symmetric and asymmetric


O=C-O-C=O stretch

0
4000 3500 3000 2500 2000 1500 1000 500
-1
Wavenumbers, cm

O O
Ph C-O-C Ph
Figure IR-37. Benzoic anhydride, KBr pellet:

117
100

80
% Transmittance

60

40

20
symmetric and asymmetric
O=C-O-C=O stretch
0
4000 3500 3000 2500 2000 1500 1000 500
Wavenumbers, cm-1

CH3

O O O
Figure IR-38. 3-Methylpimelic acid anhydride, KBr pellet:

100

80
% Transmittance

60

40

20

0
4000 3500 3000 2500 2000 1500 1000 500
2850 1012 798 Wavenumbers, cm
-1

1345 981 736


1079 892 665
O
Figure IR-39. 2,5-Dihydrofuran, neat liquid:

118
100

80
% Transmittance

60

40

20

0
4000 3500 3000 2500 2000 1500 1000 500
-1
Wavenumbers, cm
2934 1091 938
1373 1026 906
1196 976 665
OCH 3
CH 3 O O

Figure IR-40. 2,5-Dimethoxy2,5-dihydrofuran, neat liquid:

100

80
% Transmittance

60

40

20

0
4000 3500 3000 2500 2000 1500 1000 500
-1
Wavenumbers, cm
2963 1186 993
1617 1136 914
1275 1056 709
O
Figure IR-41. 2,3-Dihydrofuran, neat liquid:

119
100

80
% Transmittance

60

40

20

0
4000 3500 3000 2500 2000 1500 1000 500
Wavenumbers, cm-1

CO2Na

Figure IR-42. Sodium benzoate, KBr pellet:

120
100

80
% Transmittance

60

40

20

0
4000 3500 3000 2500 2000 1500 1000 500
-1
Wavenumbers, cm

CH3
Figure IR-43. 5-Methyl-3-hexene-2-one, tech. grade, 80%; neat CH3 O

CH3

100

80
% Transmittance

60

40

20

0
4000 3500 3000 2500 2000 1500 1000 500
-1
Wavenumbers, cm

2930 1361 982


1677 1253 606 O
1467 1167 547

CH3 CH=CH(CH2)4CH3
Figure IR-44. 3-Nonen-2-one, 95%; neat liquid, thin film:

121
100

80
% Transmittance

60

40

20

0
4000 3500 3000 2500 2000 1500 1000 500
2980 1460 1121 Wavenumbers, cm-1
1685 1402 987
1618 1203 960
Figure IR-45. Ethyl vinyl ketone, neat liquid: CH3CH2COCH=CH2

100

80
% Transmittance

60

40

20

0
4000 3500 3000 2500 2000 1500 1000 500
2974 1444 1181 Wavenumbers, cm-1
1675 1361 1127
1620 1254 973

Figure IR-46. 3-Penten-2-one, neat liquid: CH3CH=CHCOCH3

122
100

80
% Transmittance

60

40

20

0
4000 3500 3000 2500 2000 1500 1000 500
-1
Wavenumbers, cm

3368 1338 995


1666 1230 841
1578 1168 728

Figure IR-47. Benzyl 4-hydroxyphenyl ketone, Nujol mull:

O
OH

100

80
% Transmittance

60

40

20

0
4000 3500 3000 2500 2000 1500 1000 500
-1
Wavenumbers, cm

Figure IR-48. Perfluorohydrocarbon oil; thin film

123
100

80

60
% Transmittance

40

20

-20
4000 3500 3000 2500 2000 1500 1000 500
-1
Wavenumbers, cm

Figure IR-49. Quartz; KBr pellet.

124
11

10

8
Single Beam

4000 3000 2000 1000


Wavenumbers (cm-1)
Figure IR-50. IR spectrum of background and of the ATR accessory.

125
APPENDIX 2. Tables of Selected Unknowns and
Derivatives
More extensive tables of unknowns may be found in the following texts: A.I. Vogel,
Elementary Practical Organic Chemistry, Part 2, J. Wiley & Sons, Inc., NY, NY,
1966, CRC Handbook of Tables for Organic Compound Identification. 3rd ed; Zvi
Rappoport, Ph.D., CRC Press, Boca Raton, FL, 1967.

Aldehydes 2 ,4-DINITRO-
SEMI- PHENYL-
COMPOUND BP MP CARBAZONE HYDRAZONE
Note: All temperatures in ºC.

Ethanal (Acetaldehyde) 21 _ 163 168


Propanal (Propionaldehyde) 48 _ 89 155
Propenal (Acrolein) 52 _ 171 165
2-Methylpropanal 64 _ 125 187
(Isobutyraldehyde)
Butanal (Butyraldehyde) 75 _ 104 123
3-Methylbutanal 92 _ 107 123
(Isovaleraldehyde)
Pentanal (Valeraldehyde) 102 _ _ 106
2-Butenal (Crotonaldehyde) 104 _ 199 190
2-Ethylbutanal 117 _ 99 95
(Diethylacetaldehyde)
Hexanal (Caproaldehyde) 130 _ 106 104
Heptanal (Heptaldehyde) 155 _ 109 108
2-Furaldehyde (Furfural) 162 _ 202 212
2-Ethylhexanal 163 _ 254 114
Octanal (Caprylaldehyde) 171 _ 101 106
Benzaldehyde 179 _ 222 237
5-Methylfurfural 187 _ 211 212
Phenylethanal 195 33 155 121
(Phenylacetaldehyde)
2-Hydroxybenzaldehyde 197 _ 231 248
(Salicylaldehyde)
4-Methylbenzaldehyde 204 _ 221 234
(p-Tolualdehyde)
3,7-Dimethyl-6-octenal 207 _ 84 78
(d-Citronellal)
2-Chlorobenzaldehyde 213 11 229 213
4-Methoxybenzaldehyde 248 2.5 210 253
(p-Anisaldehyde)
trans-Cinnamaldehyde 250 d. _ 215 255
 Naphthaldehyde 292 34 221 _
3,4-Methylenedioxybenzaldehyde 263 37 230 266 d.
(Piperonal)
2-Methoxybenzaldehyde 245 38 215 d. 254
(o-Anisaldehyde)
4-Chlorobenzaldehyde 214 48 230 254
2, 3-dimethoxybenzaldehyde 285 54 231 264 d.

126
3-Nitrobenzaldehyde _ 58 246 293
4-Dimethylaminobenzaldehyde _ 74 222 325
Vanillin 285 d. 82 230 271 d.
4-Nitrobenzaldehyde _ 106 221 320 d.
4-Hydroxybenzaldehyde _ 116 224 280 d.
dl-Glyceraldehyde _ 142 160 d. 167

NOTE: d indicates decomposition

Ketones
SEMI- 2,4-DINITROPHENYL
COMPOUND BP MP CARBAZONE HYDRAZONE
Note: all temperatures in °C.

2-Propanone (Acetone) 56 _ 187 126


2-Butanone 80 _ 146 117
(Methyl ethyl ketone)
3-Methyl-2-butanone 94 _ 112 120
(Isopropyl methyl ketone)
2-Pentanone l01 _ 112 143
(Methyl propyl ketone)
3-Pentanone (Diethyl ketone) l02 _ 139 156
(Pinacolone) 106 _158 125
3,3-dimethyl-2-butanone
4-Methyl-2-pentanone 117 _ 132 95
2,4-Dimethyl-3-pentanone 124 _ 160 95
(Diisopropyl ketone)
3-Hexanone 125 _ 113 130
2-Hexanone 128 _ 125 106
4-Methyl-3-penten-2-one 130 _ 164 205
(Mesityl oxide)
Cyclopentanone 131 _ 2l0 146
2,3-Pentanedione 134 _ 122 (mono) 209
209 (di)
2,4-Pentanedione 139 _ _ 122 (mono)
(Acetylacetone) 209 (di)
4-Heptanone 144 _ 132 75
3-Heptanone 99-100 (111)
2-Heptanone l 51 _ 123 89
Cyclohexanone 156 _ 166 162
2,6-Dimethyl-4-heptanone 168 _ 122 92
4-Octanone 170 _ 96 _
2-Octanone 173 _ 122 58
Cycloheptanone 181 _ 163 148
2,5-Hexanedione 194 185 (mono) 257 (di)
(Acetonylacetone) 224 (di)
d-Fenchone 196 _ 184 140
b-Thujone 202 _ 174 114
Acetophenone 202 20 198 240
Isophorone 213 _ 199 _
Phenyl-2-propanone 216 27 198 156
(Phenylacetone)
Propiophenone 218 20 182 l 91
Pulegone 224 _ 174 142

127
2-Undecanone 228 12 122 63
d-Carvone 230 _ _ 191
Ketones (continued)

SEMI- 2,4-DINITROPHENYL-
COMPOUND BP MP CARBAZONE HYDRAZONE
Note: all Temperatures in °C.

4-Chloroacetophenone 232 12 _ 231


4-Phenyl-2-butanone 235 _ 142 127
1-Acetylnaphthalene 302 34 224 _
4-Phenyl-3-buten-2-one 212 37 187 227
4-Methoxyacetophenone 258 38 198 220
1-Indanone 242 42 233 258
Benzophenone 305 48 167 238
4-Bromoacetophenone 225 5l 208 230
2-Acetanaphthalene 301 54 235 262 d
2-Indanone _ 58 218 _
Desoxybenzoin 320 60 148 204
3-Nitroacetophenone 202 81 257 228
9-Fluorenone 342 83 234 283
Benzil 347 95 244 d (di) 189 (di)
d,2- Benzoin 344 l33 206 245
4-Hydroxypropiophenone _ 148 _ 229
Dimedone _ 149 _ _
Xanthone 350 174 _ _
Camphor 205 179 237 177

Carboxylic Acids
bp mp p-Toluidide Anilide Amide
Carboxylic Acid ( oC) (oC) (mp oC) (mp oC) (mp oC)

Formic acid 101 8 53 47 43


Acetic acid 118 17 148 114 82
Acrylic acid 139 13 141 104 85
Propanoic acid 141 124 103 81
2-Methylpropanoic acid 154 104 105 128
Butanoic acid 162 72 95 115
2-Methylpropenoic acid 163 16 87 102
Trimethylacetic acid 164 35 127 178
Pyruvic acid 165 d 14 109 104 124
3-Methylbutanoic acid 176 109 109 135
Pentanoic acid 186 70 63 106
Hexanoic acid 205 75 95 101
2-Bromopropanoic acid 205 24 125 99 123
Octanoic acid 237 16 70 57 107
Nonanoic acid 254 12 84 57 99
Decanoic acid 268 32 78 70 108
3-Phenylpropanoic acid 279 48 135 98 105
Bromoacetic acid 208 50 158 131 91
Chloroacetic acid 189 63 162 137 121
Palmitic acid 63 98 91 106

128
Stearic acid 69 102 95 109
trans-2-Butenoic acid 72 132 118 159

bp mp p-Toluidide Anilide Amide


Carboxylic Acid ( oC) (oC) (mp oC) (mp oC) (mp oC)

Phenylacetic acid 77 136 118 156


2-Methoxybenzoic acid 200 101 131 129
2-Methylbenzoic acid 104 144 125 142
3-Methylbenzoic acid 263 110 118 126 94
Benzoic acid 249 122 158 163 130
Cinnamic acid 300 133 168 153 147
2-Chlorobenzoic acid 140 131 118 139
3-Nitrobenzoic acid 140 162 155 143
2-Aminobenzoic acid 147 d. 151 131 109
2-Bromobenzoic acid 150 141 155
Benzilic acid 150 190 175 154
Adipic acid 154 241 241(di) 220(di)
3-Bromobenzoic acid 156 136 155
2-Hydroxybenzoic acid 158 156 136 142
3-Chlorobenzoic acid 158 125 134
2-Iodobenzoic acid 162 141 184
3,5-Dimethylbenzoic acid 166 133
d-Tartaric acid 171
4-Methylbenzoic acid 180 160 145 160
2-Naphthoic acid 184 192 171 193
4-Methoxybenzoic acid 280 184 186 170 167
Succinic acid 185 260(di)
3-Iodobenzoic acid 187 186
3-Hydroxybenzoic acid 201 163 157 170
4-Hydroxybenzoic acid 214 204 197 162
4-Nitrobenzoic acid 240 204 211 201
4-Chlorobenzoic acid 242 194 179

d: decomposes s: sublimes

Alcohols
3,5-Dinitro- Phenyl-
bp mp benzoate urethane
Alcohol (oC) (oC) (mp oC) (mp oC)

Methanol 65 108 47
Ethanol 78 93 52
2-Propanol 82 123 88
2-Methyl-2-propanol 83 26 142 136
2-Propen-1 -ol 97 49 70
1 -Propanol 97 74 57
2-Methyl-2-butanol 102 116 42
2-Methyl-1-propanol 108 87 86
3-Pentanol 115 101 48
1 -Butanol 118 64 61
2-Pentanol 119 62
3-Methyl-3-pentanol 123 96 43

129
2-Chloroethanol 129 70 31
4-Methyl-2-pentanol 132 65 143

3,5-Dinitro- Phenyl-
bp mp benzoate urethane
Alcohol (oC) (oC) (mp oC) (mp oC)

1-Pentanol 138 46 46
Cyclopentanol 140 115 132
1-Hexanol 157 58 42
Cyclohexanol 160 113 82
1 -Heptanol 176 47 60
2-Octanol 179 32 114
1-Octanol 195 61 74
Benzyl alcohol 204 113 77
1 -Phenylethanol 204 20 92 95
2-Phenylethanol 219 108 78
3-Phenyl-2-propen-1 -ol 250 34 121 90
Menthol 212 41 158 111
Diphenylmethanol 288 68 141 139
Benzoin 133 165
Cholesterol 147 168
Borneol 208 154 138

Dicarboxylic Acids (all T in C)

bp mp p-Toluidide Anilide Amide

l - Malic - 100-1 206-7(di) 197(di) 156-7(di)


Sebacic - 133s 201(di) 202(di) 210(di)
Maleic - 137 142(di) 187(di) 260(di)
2,2-dimethylsuccinic - 139-142
Adipic - 154 241 241(di) 220(di)
d-Tartaric - 169-71 263-4(di) 196d(di)
Succinic - 185 260(di)
Phthalic - 200-6 201(di) 253-5(di) 220(di)
d,l-Tartaric - 206 226d(di)

130
Primary Amines
bp mp Benzamide Picrate Acetamide
Amine (oC) (oC) (mp oC) (mp oC) (mpoC)

t-Butyl amine 46 134 198 101


Propylamine 48 84 135
sec-Butylamine 63 76 139
Isobutylamine 69 57 150
Butylamine 78 42 151
Cyclohexylamine 135 149 104
Benzylamine 184 105 194 60
Aniline 184 163 198 114
2-Methylaniline 200 144 213 110
3-Methylaniline 203 125 200 65
2-Chloroaniline 2O8 99 134 87
2,6-Dimethylaniline 216 11 168 180 177
2-Methoxyaniline 225 6 60 200 85
3-Chloroaniline 230 120 177 74
4-Methylaniline 200 43 158 182 147
2-Ethylaniline 210 47 147 194 111
4-Methoxyaniline 58 154 170 130
4-Chloroaniline 70 192 178 179
2-Nitroaniline 72 110 73 92
Ethyl p-aminobenzoate 89 148 110
3-Nitroaniline 114 157 143 155
4-Nitroaniline 148 199 100 215
2,4-Dinitroaniline 180 202 120

Secondary Amines
bp mp Benzamide Picrate Acetamide
Amine (oC) (oC) (mp oC) (mp oC) (mp oC)

Diethylamine 56 42 155
Diisopropylamine 84 140
Piperidine 106 48 152
Dipropylamine 110 75
Morpholine 129 75 146
Diisobutylamine 139 121 86
Dibutylamine 139 59
Benzylmethylamine 184 117
N-Methylaniline 196 63 145 102
N-Ethylaniline 205 60 132 54
N-Ethyl-m-toluidine 221 72
Dicyclohexylamine 256 153 173 103
N-Benzylaniline 298 37 107 48 58
Diphenylamine 302 52 180 182 101

131
Tertiary Amines (all Temperatures in °C)

Compound bp mp picrate methiodide

Triethylamine 89 _ 173 280


Pyridine ll6 _ 167 117
2-Methylpyridine (-Picoline) 129 _ 169 230
3-Methylpyridine (-Picoline) 143 _ 150 92
4-Methylpyridine (-Picoline) 143 _ 167
Tripropylamine 157 _ 116 207
N,N-Dimethylbenzylamine 183 _ 93 179
N,N-Dimethylaniline 193 _ 163 228 d.
Tributylamine 216 _ 106 180
N,N-Diethylaniline 218 _ 142 102
Quinoline 239 _ 203 133

d. indicates decomposition.

Esters (all Temperatures in °C)

Compound bp
Methyl formate 32
Ethyl formate 54
Ethyl acetate 77
Methyl propionate 80
Methyl acrylate 80
Isopropyl acetate 89
Methyl isobutyrate 93
t-Butyl acetate 98
Ethyl acrylate 101
Methyl methacrylate 100
Propyl acetate 102
Methyl butyrate 102
2-Butyl acetate 111
Methyl 3-methylbutanoate 117
Ethyl butanoate 120
Butyl acetate 127
Methyl pentanoate 128
Ethyl 3-methylbutanoate 132
Pentyl acetate 142
3-Methylbutyl acetate 142
Ethyl lactate 154
Ethyl hexanoate 168
Methyl acetoacetate 170
Dimethyl malonate 180
Ethyl acetoacetate 181
Diethyl oxalate 185
Methyl benzoate 199
Diethyl malonate 199
Ethyl octanoate 207
Ethyl benzoate 212
Diethyl succinate 217
Diethyl fumarate 218

132
Methyl phenylacetate 220

Esters (all Temperatures in °C)

Compound bp
Diethylmaleate 222
Methyl salicylate 224
Ethyl phenylacetate 228
Ethyl salicylate 234
Ethyl cinnamate 271
Dimethyl phthalate 284
Compound mp
Methyl cinnamate 36
Phenyl salicylate 42
Dibenzyl succinate 52
Ethyl p-nitrobenzoate 56
Phenyl benzoate 70
Methyl 3-nitrobenzoate 78
Methyl 4-bromobenzoate 81
Benzoin acetate 83
Methyl 4-nitrobenzoate 96
Dimethyl fumarate 102

Phenols ºC
ºC ºC -NAPHTHYL-
COMPOUND BP MP URETHANE
2-Chlorophenol 176 7 120
3-Methylphenol (m-Cresol) 203 12 128
2-Methylphenol (o-Cresol) 191 32 142
2-Methoxyphenol (Guaiacol) 204 32 118
4-Methylphenol (p-Cresol) 202 34 146
Phenol 182 42 133
4-Chlorophenol 217 43 166
2,4-Dichlorophenol 210 45 __
4-Ethylphenol 219 45 128
2-Nitrophenol 216 45 113
4-Ethylphenol 219 47 128
2-Isopropyl-5-methylphenol 234 51 160
(Thymol)
3,4-Dimethylphenol 225 64 141
4-Bromophenol 238 64 169
3,5-Dimethylphenol 220 68 109
2,5-Dimethylphenol 212 75 173
Vanillin(4-Hydroxy-3-methoxybenzaldehyde) 285 81 ----
l-Naphthol (-Naphthol) 280 94 152
1,2-Dihydroxybenzene (Catechol) 246 l05 175
1,3-Trihydroxybenzene (Resorcinol) 281 109 275
4-Nitrophenol __ 112 150

133
2-Naphthol (-Naphthol) 286 121 157
1,2,3-Trihydroxybenzene 309 133 __
(Pyrogallol)
4-Phenylphenol 305 164 __

134
Department of Chemistry
University of Missouri-St. Louis
Solid Unknown
Chemistry 263 Preliminary Report Form
Name: ________________
Date: ________
Unknown Number ______
Physical Properties
a) appearance:
b) melting point:
c) boiling point:
e) other properties you may have measured:
Solubility Tests (use +/- to indicate positive, negative tests)
H2O: _____; 5%NaOH, _____; 5%NaHCO3, _____; 5%HCl, _____ ; H2SO4,_____.
pH(aqueous solution): _____ .

135
Department of Chemistry
University of Missouri-St. Louis
Liquid Unknown
Chemistry 263 Preliminary Report Form
Name: ________________
Date: ________
Unknown Number ______
Physical Properties
a) appearance:
b) melting point:
c) boiling point:
e) other properties you may have measured:
Solubility Tests (use +/- to indicate positive, negative tests)
H2O: _____; 5%NaOH, _____; 5%NaHCO3, _____; 5%HCl, _____ ; H2SO4,_____.
pH(aqueous solution): _____ .

136
Department of Chemistry
University of Missouri-St. Louis
Chemistry 263 Solid Unknown Final Report Form
Name: ________________
Date: ________
Unknown Number ______
Compound Identification (give structure and proper name, including stereochemistry if
appropriate)

Other Possible Compounds (compounds with very similar properties):

Physical Properties
a) appearance:
b) melting point:
c) boiling point:
e) other properties you may have measured:
Solubility Tests (use +/- to indicate positive, negative tests)
H2O: _____; 5%NaOH, _____; 5%NaHCO3, _____; 5%HCl, _____ ; H2SO4,_____.
pH(aqueous solution): _____ .
Functional Groups Present (based on IR or classification tests, attach IR spectra to this
report)

Frequencies present Likely Functional Group

Derivative
a) structure:

b) melting point:

c) melting point of derivatives of other possible compounds given above:

137
Department of Chemistry
University of Missouri-St. Louis
Chemistry 263 Liquid Unknown Final Report Form
Name: ________________
Date: ________
Unknown Number ______
Compound Identification (give structure and proper name, including stereochemistry if
appropriate)

Other Possible Compounds (compounds with very similar properties):

Physical Properties
a) appearance:
b) melting point:
c) boiling point:
e) other properties you may have measured:
Solubility Tests (use +/- to indicate positive, negative tests)
H2O: _____; 5%NaOH, _____; 5%NaHCO3, _____; 5%HCl, _____ ; H2SO4,_____.
pH(aqueous solution): _____ .
Functional Groups Present (based on IR or classification tests, attach IR spectra to this
report)

Frequencies present Likely Functional Group

Derivative
a) structure:

b) melting point:

c) melting point of derivatives of other possible compounds given above:

138

You might also like