0% found this document useful (0 votes)
178 views171 pages

Chemistry 1 Nya General Chemistry Laboratory Manual: WINTER 2023 NAME

This document provides an overview of safety procedures and guidelines for a chemistry laboratory course. It outlines 10 lab safety rules and 12 laboratory practice and safety rules to protect students. A safety contract is also included that students must sign agreeing to abide by the rules, such as wearing safety goggles and closed-toe shoes at all times in the lab.

Uploaded by

Naira Metayer
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
178 views171 pages

Chemistry 1 Nya General Chemistry Laboratory Manual: WINTER 2023 NAME

This document provides an overview of safety procedures and guidelines for a chemistry laboratory course. It outlines 10 lab safety rules and 12 laboratory practice and safety rules to protect students. A safety contract is also included that students must sign agreeing to abide by the rules, such as wearing safety goggles and closed-toe shoes at all times in the lab.

Uploaded by

Naira Metayer
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 171

CHEMISTRY 1 NYA

General Chemistry
LABORATORY MANUAL

WINTER 2023

NAME: ____________________________________

SECTION: _________________________________
ii
TABLE OF CONTENTS

Foreword to the Student and Lab Safety Rules 2


Lab Safety Contract 4
Scientific Integrity and Academic Dishonesty 6
Contents of Drawers and Cupboards 8
Fundamentals of Analysis 10
Glossary of Analytical Terms 10
I. General Guidelines for Chemical Analysis 12
II. Major Equipment in the Chemistry Lab 14
A. Analytical Balance 14
B. Spectrophotometer 19
III. Liquid Handling Glassware 26
C. Beaker 27
D. Erlenmeyer Flask 29
E. Graduated Cylinder 30
F. Volumetric Pipette 33
G. Mohr Pipette 39
H. Volumetric Flask 46
I. Burette 50
J. Reading the Meniscus 57
K. Miscellaneous Glassware 58
IV. Data Collection and Analysis 59
A. Significant Figures 59
B. Introduction to Basic Statistics for Chemistry Students 64
Experiment 0: Introduction and Fundamentals of Analysis 0-1
Experiment 1: Measurements of Mass and Volume 1-1
Experiment 2: Solutions 2-1
Experiment 3: Titrations I – Standardization of NaOH and Analysis of an
HCl Solution 3-1
Experiment 4: Titrations II – Analysis of Organic Acids 4-1
Experiment 5: Titrations III – Redox Titrations 5-1
Experiment 6: Periodic Trends 6-1
Experiment 7: Molecular Geometry 7-1
Experiment 8: Colligative Properties 8-1

iii
1
Foreword to the Student

Laboratory work is an indispensable part of the chemistry curriculum. It provides the


student with experience in the use of chemicals and chemical apparatus, thereby facilitating the
student's understanding of chemical theory.
The experiments in this manual are introduced with a brief summary of the purpose of the
experiment and of the theory upon which the experiment is based. This introduction should be
read and understood before entering the laboratory. Consultation with a chemical text should
be made in the case of non-understanding of the experiment. For example, make sure that you
understand the definition of new terms, the balancing of equations, the treatment of the data and
the calculation of the results.
For the safety of yourself and your fellow students, strict adherence to the instructions
listed below is insisted upon. Disregard of these instructions will result in the offender being
denied use of the laboratory and a zero for that lab.

Academic Rules in The Laboratory

1. Prepare yourself for the lab; read the lab for understanding and complete the prelab
assignment.
2. Show up to the lab on time. Safety instructions and other information are given at the
beginning of the each lab. If you are more than 10 minutes late, you will not be allowed
to do the lab experiment.
3. There are no make-up sessions for missed labs! Students with serious commitments (i.e.
government or medical appointment) must inform the lab instructor before missing the
lab.
4. Electronic devices must be switched off while in the lab. Headphones or earbuds must
be left in bag, locker, or pocket.
5. Loud talking, whistling, singing or any other disruptive noises will not be tolerated in
the laboratory. Converse in moderate tones.
6. Your work requires your complete attention and, therefore, students should not leave
the laboratory for any extended periods of time without permission of the instructor.
7. Your work area should be kept tidy at all times and any equipment used must be
returned to the drawer or to a specified place, CLEAN and in good condition.
8. A passing grade in the lab is required for a passing grade in the course.
9. You must conduct yourself according to the lab policy on scientific integrity. Failure
to comply with this policy will result in a grade of zero on your lab report. See the
section on scientific integrity and academic dishonesty on Page 5.

(Lab Rules are continued on the following page)

2
Laboratory Practice and Safety Rules

1. Safety goggles are mandatory. Prescription glasses are allowed under safety goggles.
Contact lenses are not allowed. Anyone wearing contacts will be asked to leave the
laboratory and will receive a zero for that lab.
2. Lab coats are mandatory. Lab coats must be worn closed and with sleeves covering
forearm (never rolled up to elbows).
3. Long hair should always be tied back.

4. Before leaving the lab, students must always wash their hands with soap and water.
5. Shorts, sandals and open-toed shoes are not permitted in the lab.
6. Do not bring bags, books or clothing into the laboratory. Bring your lab manual,
calculator and writing utensils.
7. Eating, chewing gum and drinking is NEVER permitted in the laboratory. This
includes tutorial and tests held in the lab space.
8. Reagent bottles must be kept closed and returned to the appropriate storage place
when you are finished with them. Do not leave these bottles open or on your bench.
9. Unused reagents are not to be returned to the bottle. Always use a transfer beaker
instead of placing foreign objects (i.e. stirring rods, pipets) into reagent bottles.
10. Solutions are to be disposed of as instructed in the lab manual. Any solutions poured
down a drain must be flushed with water.
11. Any chemical spills or broken glassware should be brought to the attention of the
instructor or teacher. Broken glass goes in the glass waste box, not the garbage.
12. In case of accidents, know where the emergency eye wash, fire blanket and fire
extinguishers are located. Any incident should be reported to the instructor or
teacher immediately.
Should the fire alarm sound or the electricity go out during the lab, hotplates must be
unplugged. Then immediately follow the instructor to the closest exit, as indicated by the
evacuation instructions posted in the lab. Remember, no one knows if the fire alarm is due to a
real fire or a prank so treat the fire alarm as if there is a real fire. Setting off the fire alarm when
there is no fire is a criminal offence.

3
SAFETY CONTRACT

Complete this form and give it to your laboratory instructor.

I have carefully read the Laboratory Practice and Safety Rules” (pages ii and iii). Whenever I am
in an area where laboratory reagents are being used, I agree to abide by the following rules:
1. Wear safety goggles. Never wear contact lenses
2. Lab coats are mandatory in the labs, as well as wearing proper clothing (closed-toed
shoes, legs must be covered: no shorts, skirts, etc).
3. Do not eat, drink, chew gum or apply cosmetics in the laboratory.
4. Be familiar with the location and use of all safety equipment.
5. Report all incidents to the laboratory instructor.
6. Treat all laboratory reagents as if they are poisonous and corrosive.
7. Use good housekeeping practices.
8. Dispense reagents carefully. Dispose of laboratory reagents as directed.
9. Become familiar with each laboratory assignment before coming to the laboratory.
10. Anticipate the common hazards that may be encountered in laboratory.

Yes, I wear contact lenses/ prescription glasses.


Occasionally Always
No, I do not wear contact lenses/ prescription glasses.
If you are colour-blind, include this information on here: Yes No

For my own safety, I will not wear contact lenses in the laboratory; I agree to wear safety
goggles in the chemistry laboratory at all times (with my prescription glasses underneath, if
I use them). Failure to follow this rule will result in a grade of zero with no make-up for
that lab.
Please give any health information that might help the laboratory instructor and teacher to
provide a safer environment for you, or that could aid the laboratory instructor in responding to
an incident involving you in the lab.

Allergies:
Medical Issues:
Medication being taken:

STUDENT NAME: _______________________________ CHEM SECTION: _______


STUDENT SIGNATURE: _________________________ DATE: ________

4
5
Scientific Integrity and Academic Dishonesty
For all scientists, behaving with scientific integrity is crucial.

Scientific integrity refers to maintaining the quality and objectivity of scientific research and
experimentation such that these activities are sound and worthy of the public’s confidence.

For the public, it is taken to be a given that the data reported in a scientific publication are
accurate, truthful and objective. It is the responsibility of the scientist to behave in accordance
with these expectations.

Manipulated or false data are poison to the process of scientific inquiry, and, when they go
undetected, set back scientific fields for years. Progress is greatly delayed when the scientists can
not trust the work of their peers.

The consequences of academic dishonesty are extremely serious. As one particularly egregious
example of academic dishonesty, consider the case of the famous and award-winning (former)
research physicist Jan Hendrick Schön, who was caught manipulating data in his published
research papers in 2002. After a series in investigations into his misconduct, Schön was
dismissed from his research position. Schön’s 28 research articles have been retracted by their
respective journals. In 2004, Schön’s PhD was revoked. The gravity of these events needs no
exaggeration: Schön’s prospects of working in his chosen field are practically nil, and his name
is forever tarnished as a result of his dishonesty.

The seriousness of academic dishonesty cannot be understated. As a scientist-in-training you are


required to conduct your laboratory research activities with scientific integrity, even at the
college-level. This includes the laboratory components of your Science courses at Marianopolis.
You are thus expected to honestly and accurately report every measurement carried out in the lab.

As a scientist-in-training you are encouraged to question or even doubt the validity of a


measurement but it must be reported as it was observed. If you believe that a measurement is
flawed, you are encouraged to approach your instructors to discuss your suspicions. However,
you must never manipulate or falsify experimental data. Your results are not to be
doctored or adjusted for any reason. A grade zero will be assigned for any labs in this
course bearing falsified or manipulated data.

6
7
GENERAL CHEMISTRY: EQUIPMENT LIST

LARGE DRAWER
VOLUMETRIC FLASKS
1 x 50 mL with yellow cap
1 Liquid Funnel
1 X 100 mL with blue cap
ERLENMEYER FLASKS
GRADUATED CYLINDERS
3 x 125 mL
1 X 10 ml
1 X 250 ml
1 X 25 ml
1 X 100 ml

BEAKERS
- 1 x 30 mL 1 Watch Glass
1 x 50 mL 1 X 50 mL 1 Large Stirring Rod
1 X 100 mL 1 X 100 mL
1 X 150 mL 1 X 150 mL TEST TUBE RACK
1 X 250 mL - 5 Large Test Tubes
1 X 400 mL - 7 Small Test Tubes
1 X 600 mL - 2 Small Stirring Rods

SMALL DRAWER CUPBOARD

1 Stand
10 mL Mohr Pipette
1 Buret Clamp
10 mL Volumetric Pipette
1 Ring Clamp
Green (10 mL) Pipette Helper
1 Hot Plate

8
9
Fundamentals of Analysis
Foreword:

The majority of chemistry, biochemistry and biology research experimentation require liquids
and solutions to be measured, contained, transported and delivered in some form. Many
experimental fields rely extensively on these methods. Developing a mastery of the techniques of
preparing and handling solutions in the laboratory is important for most scientists and important
for understanding many of the concepts taught in this and other chemistry courses.

We will begin by developing a familiarity with liquid handling techniques and the related
equipment before these skills are put into practice in experiments. Having an understanding of
the fundamentals of chemical analysis before beginning experiments will give you the greatest
chance of success in these experiments and give you the most comprehensive learning
experience for this course.

Glossary of Analytical Terms:

Analyte. The molecule or species of interest that is being analyzed. During a titration, the
analyte is usually placed in an Erlenmeyer flask. In spectrophotometric analyses, the analyte is
placed in a cuvette.

Aspirate. To draw liquid by suction.

Balance. A scale; an instrument for measuring the mass of an object or material.

Beaker. A wide-mouthed glass vessel with a spout for pouring liquids into other glassware.

Blank. A solution containing none of the analyte of interest used to calibrate the
spectrophotometer; typically the blank will consist of the solvent used to dissolve the analyte, but
none of the analyte itself.

Burette. Glassware designed to deliver measured volumes of liquids to another vessel with high
accuracy, typically during a titration.

Cuvette. An optically transparent vessel used to contain liquid samples for spectrophotometric
analysis.

Empirical. Measured or determined by experimentation rather than pure theory.

Erlenmeyer Flask. A conical flask with a cylindrical neck.

Graduation. The markings on a measuring instrument that indicate the quantity being measured.

Kimwipe. A lintless towel.

10
Lightpath. The path a beam of light takes from the light source, through a sample, and into a
detector in microscopy, or spectroscopy.

Meniscus. The surface of a liquid inside a vessel. Typically, polar solvents like water will give a
concave (smiling face) meniscus inside vessels with polar surfaces, like clean glass.

Pipet. (verb). To measure and deliver a liquid using a pipette.

Pipette. (noun). Glassware designed to measure and deliver small volumes of liquids with high
accuracy and precision.

Reagent. A material that is added to the reaction mixture to bring about a chemical reaction.

Receiving Vessel. A container that will receive a liquid being delivered from another vessel.

Receptacle. A container that will receive a liquid being delivered from another vessel.

Sheet. To coat the inside surfaces of a vessel with a liquid.

Spectrophotometer. An instrument that measures the amount of light absorbed by a sample at a


given wavelength.

Stock Solution. A concentrated solution, usually prepared in a large batch, that will be sampled
and diluted before it is used in experiments.

Tare. To zero the balance. If an object is placed on the balance and the 0/T button is pressed,
that object is 'tared' and the balance is zeroed with that mass on the platform.

Theoretical. Determined by largely theoretical means, or based on an aggregate of literature


values.

Titrant. A solution delivered to the analyte during titration, usually from a burette.

Titration. An experiment designed to quantitatively determine the end-point of a chemical


reaction by incremental combination of the reagents. The titrant is incrementally delivered with a
burette to the analyte.

Uncertain Digit. The last significant figure in a measured value. Often the uncertain digit must
be estimated.

Vessel. A container that holds liquid.

Volumetric Flask. Glassware designed to contain solutions with precisely measured volumes,
often used during solution preparation.

11
I. General Guidelines for Chemical Analysis

A. Reagent Bottle
-Reagent bottles contain pure chemicals. Avoid contaminating these bottles at all costs.

-Never transfer material into the reagent bottle. e.g. If you have some extra chemical left over,
DO NOT pour it back into the reagent bottle; instead pour it into the appropriate waste container.

-Avoid leaving reagent bottles open. Once you are finished collecting the chemical you need
from the reagent bottle, close the lid immediately. Do not leave the lid open for the next person
to use the bottle.

-Use a clean spatula or spoon to scoop solid chemicals from the reagent bottle into a receiving
vessel whenever possible instead of pouring directly from the reagent bottle.

B. Stock Solution
-A stock solution is a concentrated solution that is meant to be diluted to some lower working
concentration for experiments. Often a single stock solution is prepared in a large batch. Small
amounts of the stock solution can then be conveniently sampled and used for different
experiments. Thus, contamination or alteration of the composition of the stock solution must be
avoided at all costs; this would affect the results of all the experiments that subsequently use the
stock solution.

-Pour the stock solution from the stock bottle into a clean and dry beaker, and then use the
beaker to pour the stock solution as needed. Avoid pouring from the stock bottle directly into a
graduated cylinder. Never pipet directly out of the stock bottle - this risks contaminating the
stock solution.

-Never transfer material into the stock solution. e.g. if you have some extra stock solution
leftover, DO NOT pour it back into the stock bottle under any circumstances. Instead, pour the
leftover stock solution into the appropriate waste container.

C. Tap Water vs. Purified Water


-Tap water is a solution containing many contaminants. Ions such as Na+, Ca2+, Mg2+, Cl- and
trace amounts of Fe2+, K+, CO32-, NO3- and F- are present in tap water. Water also contains
organic compounds and even microorganisms. Altogether, these impurities generally make tap
water an unsuitable solvent for chemical analyses, with a few exceptions.

-There are two major classes of purified water available in the modern chemistry lab:

i. Distilled water has been boiled and the vapours have been collected and condensed. This
greatly reduces the concentration of non-volatile contaminants (ions, intact microorganisms and
many organic compounds) present in the water.

12
ii. Deionized water has been filtered to remove ionic contaminants. Many different methods of
deionization exist with a range of purities.

-Purified water (either distilled water or deionized water) is expensive to produce and should not
be unnecessarily wasted.

-Neither type of purified water is entirely pure, but either is preferable to using tap water as a
solvent for most analyses.

e.g. During solution preparation, the solvent is typically distilled water.


e.g. In a titration, the analyte (the sample) is typically prepared with distilled water.
e.g. In a titration, the titrant (the solution being delivered to the analyte) is typically prepared
with distilled water.

-Tap water is nonetheless entirely appropriate for cleaning purposes (see guidelines for cleaning
glassware below).

-Typically glassware is cleaned with tap water, and then rinsed with distilled water before use or
storage.

-In laboratories at Marianopolis College, the purified water is actually deionized water. For
simplicity and for historical reasons, all the purified water in the lab is referred to as ‘distilled
water’. Henceforth, the terms ‘distilled water’ and ‘deionized water’ are used synonymously in
this manual because the distinctions between these types of water are not of consequence for the
experiments described herein.

13
II. Major Equipment in the General Chemistry Lab

What follows is a detailed description of each major piece of equipment to be used in this
course. The methods for handling the equipment safely and using the equipment accurately
and consistently are explained.

A. Analytical Balance
The analytical balance is an electronic scale designed to measure small masses with high
accuracy and precision. The balances in our chemistry labs measure masses in the range of 0 to
180 g to a precision of 1 mg (0.001 g). Typically the balance will be used to determine the mass
of solids, and occasionally liquids.

The balance is used extensively in research, and correspondingly the balance will feature heavily
in this and other chemistry courses. Mastering the analytical balance is important for accurate
experimental results and crucial for the maintenance of this expensive piece of equipment.

The modern analytical balance is electrically operated and microprocessor-controlled. The


instrument is essentially a flat box with controls and a digital readout facing the front. In the
centre of the instrument is a measuring platform onto which the sample is placed. Typically the
platform is covered with a glass enclosure which prevents air currents from affecting the reading.
When taking measurements, one should ensure that all doors to the enclosure are closed.

The balance is simple to operate, but expensive to replace, so care must be taken not to
contaminate or damage the instrument. Abuse of the balance can easily render the instrument
irreparably damaged so please treat the balance with appropriate care.

A schematic diagram of the analytical balance appears below.

Platform

On/Off 0/T

Power switch Digital readout reports the mass in 0/T switch sets the balance to 0.000 g.
grams on the platform

Front view of the analytical balance. Placement of the buttons may vary depending on
the balance. Not pictured: glass enclosure.

The On/Off button on the balance turns the power on/off, respectively. You must hold the off
button for a few moments to turn the balance off. The 0/T button ‘zeroes’ or ‘tares’ the balance;
this sets the balance at a baseline reading of 0 regardless of what is on the balance platform.

14
General Guidelines:
-The analytical balance is an expensive and delicate instrument. Abusive handling that would
damage the balance must be avoided at all costs. For example, contamination of the balance with
solids or liquids will damage the instrument and must be avoided. Stressing the balance by
placing more than the maximum mass on the platform must likewise be avoided.

-Chemicals are never placed directly on the platform. If it is necessary to determine the mass of a
sample of a material, the sample is placed in a vessel (e.g. a beaker) and this vessel is then placed
on the balance platform. Thus, the likelihood that the instrument becomes contaminated with the
sample is minimized. A list of some appropriate sample vessels for use with the analytical
balance is presented in the table below.

Sample Vessels Description


Weigh Boat A disposable plastic bowl
Weighing Paper A thin sheet of disposable wax paper
Beakers, Erlenmeyer Standard lab glassware, described below
Flasks
Watch Glass A glass saucer

-The outside of the vessel must always be dry before it is placed on the balance platform.
Typically the inside of the vessel should be dry, but there are situations where this is not strictly
necessary. See individual experimental protocols for details.

-Avoid excessive handling of the vessel. Fingerprints and water droplets on the outside of the
vessel will cause error in your measurements. As a rule, the outside of the vessel should be kept
clean and dry before any measurements with the balance are made.

-Importantly, the mass of the vessel must be accounted for. There are two methods of accounting
for the vessel’s mass. Either:

i. The vessel’s mass may be subtracted from the total mass of the vessel and sample. The mass of
the vessel must be determined before it is filled, when it is empty and clean.

ii. The vessel’s can be ‘tared’ before the sample is added. The empty and clean vessel is placed
on the balance, the balance is ‘tared’, and then the sample is added. Thus the mass of the sample
is determined directly.

Depending on the details of the experimental protocol, either method of accounting for the mass
of the vessel may be most appropriate. With experience, you will learn which method to use in a
given situation. For training purposes in this course, we will indicate which method to use.

-Under no circumstances should a student tamper with the balance platform. If there is an issue,
consult your instructor.

-Hot objects should not be placed in the balance case or weighed; all objects must be at room
temperature in order to obtain accurate masses.

15
-The analytical balances in our laboratory report masses to 1 mg (0.001g). For masses measured
in g, this means the instrument has a precision of 3 digits after the decimal. No estimation is
required, because the analytical balances in our lab give a digital reading. All masses measured
with the analytical balance should thus be reported with 3 digits after the decimal (for masses in
g).

General workflow:
-You should always make sure the balance is clean and in good working order before taking any
measurements. If the balance platform is contaminated with solids, a paintbrush may be used to
gently brush the powder onto the countertop to be discarded. The balance should never be
contaminated with liquids, but if this does occur, immediately consult your instructor.

-The power must be on in order to take measurements with the balance. Pressing the 0/T button
will zero the balance and give a reading of 0.000 g.

-The reading should be stable with the doors of the enclosure closed.

-Never push or attempt to move the balance.

-Never add liquid samples to your measuring vessel inside the balance or near the balance.
Liquid handling should occur at your bench, not at the balance table.

-Carefully open the doors of the enclosure. The doors are glass and will shatter if dropped.

-Gently place measuring vessels onto the balance platform. Do not drop the vessel onto the
platform, or press down on the platform. Stressing the platform will damage the balance.

-If you need to make multiple measurements for a given experiment, use the same balance for
each measurement if possible. This will maximize the precision of your measurements by
eliminating variability that sometimes occurs between different balances.

-When taking a measurement, place the measuring vessel on the balance platform, close the
enclosure doors and wait for the reading to stabilize.

-Record the mass reported by the balance including all the digits. All of the digits returned by the
balance are significant.

-Remove your sample vessel from the balance and ensure that the balance is clean. If you have
contaminated the balance with solids, clean the balance with a paintbrush as described above. If
you have contaminated the balance with liquids, consult your instructor.

Example Workflows:

e.g. i. Task: Determine the empty mass of an Erlenmeyer flask. Then determine the mass of
100.0 mL of water poured into the Erlenmeyer flask.

16
Suggested Method:
a. Acquire a clean and dry 150 mL Erlenmeyer flask. Turn on the analytical balance and press
the 0/T button to zero the balance.

b. Make sure the outside of the Erlenmeyer is dry. Place the Erlenmeyer flask on the balance
platform, close the enclosure doors, and wait for the reading to stabilize. Record the reading.

c-d. Remove the flask from the balance and return to your bench. At your bench (not on the
balance table), add 100.0 mL of water to the Erlenmeyer flask.

e. Return to the balance table. Zero the balance by pressing the 0/T button.

f. Make sure the outside of the flask is dry. Place the flask containing the water onto the balance,
close the enclosure door, and wait for the reading to stabilize. Record the reading. In the
flowchart below, a total of 161.904 g - 63.187 g = 98.717 g of water would be reported.

On/ On/ On/ On/


Off 0.000 g 0/T
Off 63.187 g 0/T
Off 0.000 g 0/T
Off 161.904 g 0/T

a b c d e f

Flowchart e.g. i, above.

e.g. ii. Task: Weigh 3 g of NaCl (s) into a beaker for the preparation an NaCl (aq) solution.

Suggested Method:
a. Acquire a clean and dry 150 mL beaker. Make sure the outside of the beaker is dry.

b. Place the beaker on the balance platform. The mass of the beaker will be indicated.

c. Press the 0/T button. The balance should return a reading of 0.000 g. The balance and beaker
are now ‘zeroed’ or ‘tared’.

d. Carefully add some NaCl (s) from the reagent bottle, using a spoon or spatula into the beaker.
Add the solid to the beaker until the reading is 3 g. Record the reading. In the flowchart below, a
3.157 g sample of NaCl (s) was acquired: the reading would be recorded as 3.157 g.

17
On/ On/ On/ On/
Off 0.000 g 0/T
Off 63.187 g 0/T
Off 0.000 g 0/T
Off 3.157 g 0/T

a b c d

Flowchart e.g. ii, above

Many experiments call for a specific mass of a reagent to be added to a vessel, such as in the
example workflow ii, above.

In this situation, it is typically not necessary to acquire exactly 3.000 g of the sample. Notice the
wording of the task said “3 g.” This implies 3 g within reason. So if we acquire a total of 3.157 g
in our vessel, this is likely fine, and we should not attempt to remove sample from the vessel.
Likewise, if we acquire 2.856 g of the sample, this likely fine, and it is not necessary to add more
sample to the vessel. In each case, we should record the mass that was actually acquired. We
should report 3.157 g or 2.856 g as the case may be, not 3.000 g.

With experience, you will be able to determine whether we need to acquire exactly 3.000 g or
just approximately 3 g, both from the scientific language used in the experimental protocol, and
from the objectives of the experiment. For the purposes of this course, we will help guide you
towards what is appropriate.

If we accidentally acquire an unreasonable amount of the sample, we can discard some of the
sample. Remove the vessel from the balance before you do so. Extra sample should be discarded
in the waste; this waste should never be returned to the reagent bottle.

B. Spectrophotometer
Many materials have a property of absorbing light at certain wavelengths, while allowing other
wavelengths of light to be transmitted (to pass through). A spectrophotometer is an instrument
that measures the light at a specific wavelength that passes through a sample. The diagram below
illustrates the main features of a typical spectrophotometer.

605 nm 0.000

18
A basic spectrophotometer. The digital display on the top of the instrument indicates the
wavelength setting (in nm) and the current absorbance reading. The keypad on the left side of
the instrument controls the instrument settings: the +/- buttons adjust the wavelength setting;
the ‘REF’ button zeroes the spectrophotometer. The keypad contains other controls that we do
not need for the purposes of this course. The sample chamber on the right side of the
instrument is indicated with a white arrow. The arrow indicates the direction of the
instrument’s lightpath through the sample. The sample chamber in this case accommodates a
1 cm cuvette. The lid of the sample chamber should be closed whenever measurements are
being made; in this illustration, the lid is shown in the open position.

The amount of light that is absorbed by a sample (i.e. light that does not pass through the sample)
is proportional to the concentration of the material in the sample that absorbs the light at that
wavelength. Thus, for solutions containing coloured solutes, the absorbance of the solution is
proportional to the concentration of the solute, according to the Beer-Lambert law:

A  l C

where
A is the absorbance of the solution, a unitless measure of the light absorbed by the sample;
 is the molar absorptivity of the solute: a measurement of how strongly a chemical species
absorbs light, at a given wavelength, typically expressed in L mol-1 cm-1;
l is the path length: the physical length of the lightpath through the sample, typically expressed in
cm;
C is the concentration of the light-absorbing solute, expressed in moles/L.

For a given solute being measured with a given instrument,  and l are constants, and thus the
Beer-Lambert law can be simplified to:

Ak C

In this case, k is sometimes colloquially referred to as the ‘Beer’s constant’ for that given solute
measured with that instrument (using a constant path length). Thus under conditions of a
constant path length, a linear relationship exists between the concentration of a coloured solute
and the absorbance of the solution because this simplified version of the Beer-Lambert Law is a
linear equation in the form

y = mx + b.

In other words, a plot of absorbance with respect to the concentration of a coloured solute should
intersect the origin (have a y-intercept of 0) and a slope equal to k.

For example, a plot of the absorbance of light at 605 nm with respect to the concentration of
CuSO4 (aq) in solution is presented below. Briefly, a series of nine CuSO4 solutions with different
concentrations was prepared. The absorbance of each solution at 605 nm was determined with a
spectrophotometer. The resulting plot is a ‘standard curve’ of the absorbance for CuSO4 (aq)
solutions with this instrument.

19
0.5
0.45
0.4
0.35
0.3 y = 0.9623x
A605

0.25 R2 = 0.9999
0.2
0.15
0.1
0.05
0
0 0.1 0.2 0.3 0.4 0.5
[CuSO4] (M)
Standard curve for CuSO4 (aq) measured with a
spectrophotometer with a 1 cm path length at 605 nm.

Notice the relationship between absorbance and concentration in the CuSO4 standard curve
above is linear, as the Beer-Lambert law predicts. Notably, regression analysis resulted in a
linear relation that does not intersect the origin as it should according to the Beer-Lambert law;
this is a result of experimental error and limitations of the sensitivity of the spectrophotometer
used to make the measurements.

Importantly, the standard curve above can be used experimentally to determine the concentration
of a CuSO4 (aq) solution. Provided we have a spectrophotometer with the same path length as was
used to generate the standard curve, we only need to measure the absorbance of a CuSO4 (aq)
solution. Knowing the absorbance of a CuSO4 (aq) solution will allow us to extrapolate the
concentration of that solution using the standard curve above.

e.g. We have a CuSO4 (aq) solution of unknown concentration. We measure the absorbance of the
CuSO4 (aq) solution at 605 nm using the same spectrophotometer that was used to generate the
standard curve, above. The absorbance of the solution at 605 nm was 0.372. Extrapolating from
the standard curve, we would estimate that the molar concentration of the CuSO4 (aq) solution
was about 0.387 M.

General Guidelines:
-The spectrophotometer is a delicate and expensive instrument. Treat it with appropriate care.

-Solutions should not be poured into the spectrophotometer. Typically, solutions of interest are
poured into a cuvette, a small vessel made of optically transparent material

20
-Cuvettes may come in shapes and sizes, but are typically small test-tubes with a square cross
section. The cuvettes in our laboratory are optically-transparent plastic, and have a path length of
1 cm. This means the distance between the inner walls of the cuvette is 1 cm. Thus, when the
cuvette is filled with a liquid sample, a beam of light will travel through a path length of 1 cm of
the sample. A typical 1 cm cuvette is pictured below.

1 cm
1 cm

An optically transparent plastic cuvette with 1 cm path length. This cuvette is essentially a test
tube with a 1 cm-square cross section, and will contain about 2 mL of a liquid sample. The
gray line indicates a single graduation: the cuvette should be filled to about this level. The
gray arrowhead indicates the ‘front’ surface of the cuvette. This surface should be aligned
with the lightpath of the spectrophotometer. Light will enter through the ‘back’ surface of the
cuvette, pass through the sample, and then pass through the ‘front’ surface of the cuvette, into
the spectrophotometer’s detector.

-For the most accurate analyses of coloured solutes, the absorbance of the sample must always be
corrected for the background. i.e. The absorbance of the sample – background absorbance will be
taken to be the absorbance of the analyte. This is necessary because even when the solvent
consists of only purified water, the water itself and any trace impurities present may have a small
absorbance at a given wavelength. Moreover, the cuvette itself will often have some
characteristic absorbance at a given wavelength. As a rule, any absorption by the cuvette, solvent
and contaminants should be subtracted from the absorbance of the sample, as a correction to
accurately determine the absorbance of the coloured solute of interest.

-The correction described above can be routinely applied by using a ‘reference’ sample. The
reference, sometimes referred to as a ‘blank’ is a cuvette containing only the solvent used in the
analysis. The absorbance of the reference is taken, and used as a baseline for measurements of
the sample. This baseline should be subtracted from all absorbance measurements of the sample.
Most modern spectrophotometers will apply this correction automatically once the reference
sample is set.

-A spectrophotometer is capable of generating a range of different wavelengths of light. This


allows the operator to select a wavelength most appropriate for the sample being analyzed.
Ideally, the operator would select a wavelength for which the analyte (the coloured solute of
interest) has a high absorption, but the solvent and other components of the sample have
relatively low absorptions.

-Cuvettes are delicate and must be handled with care. Although plastic cuvettes are unlikely to
break, mishandling the cuvette will result in scratches that will affect the results.

21
-Solutions can be carefully poured into cuvettes from a beaker

-The cuvette should be filled with liquid to its graduation. If the level of liquid is significantly
below the graduation, the sample will not intersect the light path. .

-The outside of the cuvette should be clean and dry before it is inserted into the sample chamber
of the spectrophotometer. If the outside of the cuvette is wet with solution or water, dry it with a
Kimwipe before inserting it into the sample chamber.

-The absorbance readings taken with a spectrophotometer are highly sensitive to contamination.
Fingerprints on the cuvette surface will be sufficient to cause significant experimental error. As a
rule, handle the cuvette by its ‘sides’, i.e. the surfaces on either side of the arrowhead on the
‘front’ of the cuvette. Since the lightpath is through the front/back of the cuvette, handling the
sides of the cuvette will not pose a problem for your measurements. See the diagram of the
cuvette above.

-As a precautionary measure, it is typical to wipe the front and back surfaces of the cuvette with
a Kimwipe immediately before inserting the cuvette into the sample chamber. The lightpath will
pass through the front and back surfaces, so these surfaces must be clean and absolutely free of
fingerprints.

See the diagram of the cuvette above.

-When setting the reference, or taking absorbance readings with the spectrophotometer, the lid of
the sample chamber should be closed. This will minimize the introduction of room light into the
sample chamber, which can confuse the instrument’s detector and lead to experimental error.

-The spectrophotometers in our lab will give absorbance readings to three decimals. All of the
digits reported by the spectrophotometer are significant, so record all of the digits for each
measurement.

-Typically the spectrophotometer will be powered on, warmed up and set up for you before you
arrive at the lab. Avoid adjusting settings other than those described in this section. Consult your
instructor if you have any issues with the instrument.

Example Workflow:
e.g. i. Determine the absorbance of a CuSO4 (aq) solution at 605 nm using the
spectrophotometer.

Suggested method:
a. Acquire two clean cuvettes. Rinse the cuvettes with distilled water: fill each cuvette with
distilled water, and discard this rinse into a waste container. Repeat the rinse three times for the
best results.

22
b. Fill the first cuvette with distilled water to the graduation. This cuvette will serve as the
reference (blank). This implies that the solvent for the CuSO4 solution is simply distilled water
and contains no other solutes.

c. Rinse the second cuvette with the CuSO4 (aq) solution to be analyzed; fill the cuvette with the
solution, and discard this solution into the waste container. The cuvette should be rinsed with the
solution at least once. Fill the cuvette with the CuSO4 (aq) solution to the graduation.

d. Make sure the spectrophotometer is set to make measurements at the specified wavelength of
605 nm. Use the +/- keys to adjust the wavelength setting as necessary.

e. Wipe the front and back surfaces of the reference cuvette with a Kimwipe. Insert the reference
cuvette (blank) into the sample chamber of the spectrophotometer. Take care to align the
arrowhead on the cuvette with the white arrow on the spectrophotometer indicating the lightpath.
Close the lid. The instrument may return a non-zero absorbance reading. Press the ‘Ref’ button
on the spectrophotometer. The display should now return an absorbance reading of 0.000.

f. Remove the reference cuvette from the sample chamber. Wipe the front and back surfaces of
the cuvette containing CuSO4 (aq) with a Kimwipe. Insert the cuvette into the sample chamber of
the spectrophotometer. Take care to align the arrowhead on the cuvette with the white arrow on
the spectrophotometer indicating the lightpath. Close the lid. The instrument will now return the
absorbance reading of the sample. DO NOT press the ‘Ref’ button or any other button. The
spectrophotometers in our lab will return the absorbance reading automatically. In this
example\the absorbance of the CuSO4 solution would be reported as 0.540.

A flowchart for e.g. i appears on the following pages.

23
H 2O

a 3x
H 2O

b 'Blank'

c 'Sample'

Flowchart for e.g. i, above. Continued on the following page. . .

24
605 nm 0.000

'Blank' 'Sample'
d

'Blank'

Press

605 nm 0.000 605 nm 0.008 605 nm 0.000

'Blank' 'Sample'

605 nm -0.008 605 nm -0.008 605 nm 0.540

If one does not get stable value on the spectrophotometer, this could be caused by an air bubble inside
cuvette. Remove the air bubble and put cuvette back in spectrophotometer.
Flowchart for e.g. i, above.

25
III. Liquid Handling Glassware
Liquid handling refers to the techniques of measuring, containing, transporting, and delivering
liquids and solutions. Various specialized glassware are used to handle liquids in the lab. What
follows is a set of general guidelines for using glassware in the chemistry laboratory, and a list of
the most important glassware with detailed instructions on the proper use of each.

General cleaning guidelines

-Laboratory glassware must be treated with care. Glassware dropped on the floor will definitely
break; glassware knocked over onto the bench top or into other glassware is likely to chip or
break. Broken glassware poses a safety hazard and is expensive to replace.

-As a rule, glassware should be cleaned with warm tap water and a sparing amount of detergent.
Some glassware can be cleaned with brushes, but other glassware cannot. For specific cleaning
guidelines, see sections A through I below.

-Be very careful when washing glassware. Soapy water is very slippery and may cause you to
drop the glassware.

-Use a small amount of detergent when washing glassware. Excess detergent does not improve
the cleanliness of the glassware, and will take more time and more water to rinse away.

-Detergent used to clean glassware must be thoroughly rinsed away with tap water. Detergent
residue will often affect your results.

-Tap water contains mineral ions such as Ca2+ and Mg2+ which may affect your results and/or
contribute to a residue on the glassware if left to dry. Thus, washed glassware should typically be
rinsed with distilled water before use or storage. Two or three rinses with small amounts of
distilled water on the inside surfaces of the glassware is typically sufficient.

-Do not dry cleaned glassware with paper towel. From an analytical perspective, paper towel is
filthy and this contamination may affect your results. There is only one exception to this rule:
you will need to dry the outside of any flasks, beakers, or other vessels before you place them on
a balance. You may use paper towel on the outside of these glassware for this purpose.

Principals of measuring liquids in graduated glassware

-Graduated glassware is calibrated with a built-in scale to indicate the volume of liquid it
contains. These ‘graduations’ are imprinted on the glassware itself and indicate the volume with
varying degrees of accuracy, depending on the type of glassware.

-Some glassware is crudely manufactured, and the accuracy of graduations is relatively poor. e.g.
plastic graduated cylinders, beakers, Erlenmeyer flasks
-Some glassware is manufactured to narrow tolerances, and is correspondingly much more
expensive. This type of glassware is usually individually calibrated by hand, and thus tends to be

26
highly accurate. This glassware is most suitable for analytical measurements of volume. e.g.
volumetric pipettes, volumetric flasks, Mohr pipettes.

-Some volumetric glassware have only a single graduation. When using these glassware, the goal
is to fill the glassware until the meniscus is exactly centered on the graduation. See Reading The
Meniscus, in Section III.H., below. We cannot estimate the volume of liquid if the meniscus is
above or below the graduation. If the meniscus is above or below the graduation, the level of
liquid should be adjusted to more closely approach the graduation.

e.g. For a 25 mL volumetric pipette, the single graduation indicates 25.00 mL. One should never
attempt to deliver any other volume with this glassware. One should never report 25.01 mL if
the meniscus is slightly above the graduation.

-Measurements with graduated glassware can never be made above the uppermost graduation.

e.g. For a 25 mL graduated cylinder, the 25.0 mL graduation is the uppermost graduation. We
can never use this graduated cylinder to measure 26 mL.

e.g. If the meniscus is above the 25.0 mL graduation, we can never estimate the true volume. We
should never report 25.1 mL for a 25 mL graduated cylinder.

-Measurements with graduated glassware can never be made below the lowermost graduation.

e.g. For many 25 mL graduated cylinders, the 5.0 mL graduation is the lowermost graduation.
We can never use such a graduated cylinder to measure 4.0 mL.

e.g. If the meniscus is below the 5.0 mL graduation, we can never estimate the true volume. We
should never report 4.9 mL for such a 25 mL graduated cylinder.

A. Beaker
These vessels come in various sizes, named for their approximated total volume. A beaker is
typically used as a temporary storage vessel for a liquid. Its wide mouth makes pouring liquids
into from the beaker relatively easy; likewise its spout makes the beaker suitable for transferring
liquids to other vessels.

100

75

50

25

A 100 mL Beaker

27
The graduations on a beaker are for estimation purposes. Any graduation on a beaker is
considered to be precise to zero digits after the decimal.

e.g. A beaker filled to the 75 mL graduation would be considered to contain 75. mL. If this
volume were delivered to another vessel, we would say 75. mL were delivered.

Applications:
-Solutions are poured from larger vessels (e.g. stock bottle) into a beaker before they are poured
into a measuring vessel such as a graduated cylinder or burette. This minimizes spilling and
waste when pouring from a large stock bottle.
-Solutions are poured from larger vessels (e.g. stock bottle) into a beaker before they are pipetted.
The pipette never comes into contact with the solution inside the stock bottle, eliminating the
possibility of contamination of the stock solution.

Cleaning Guidelines:
-Beakers should be washed with warm tap water and a small amount of detergent. A bottle brush
may be used to clean the interior surfaces.
-Rinse the beaker thoroughly with tap water to remove detergent residue.
-Rinse the beaker with a small amount of distilled water to remove remnants of tap water inside
the vessel before use or storage.

Example Workflow:
e.g. i. Task: Pour 60 mL of a stock solution into a beaker.

Suggested Method:
a. Acquire a clean and dry 100 mL beaker.

b. Pour from the stock bottle directly into the beaker.

c. Read the final volume using the graduations on the beaker. In the flowchart below, the volume
should be read as 75. mL.
Sto tion
ck
lu
So

100 100 100

75 75 75

50 50 50

25 25 25

a b c

Flowchart for e.g. i, above.

28
B. Erlenmeyer Flask
Erlenmeyer flasks are sometimes referred to as ‘conical flasks’ because of their cone-shape.

Erlenmeyer flasks, like beakers, come in various sizes named for their approximate total volume.
The narrow neck and opening of the Erlenmeyer flask allows the contents of the flask to be
swirled while minimizing the risk of spilling. Notice that the graduations on the Erlenmeyer flask
are non-linear, due to the conical shape of the vessel.

100

75

50
25

A 125 mL Erlenmeyer Flask

The graduations on an Erlenmeyer flask are for estimation purposes. Any graduation on an
Erlenmeyer flask is considered to be precise to zero digits after the decimal.

e.g. An Erlenmeyer flask filled to the 50 mL graduation would be considered to contain 50. mL.
If this volume were delivered to another vessel, we would say 50. mL of the liquid was delivered.

Cleaning Guidelines:
-Like beakers, Erlenmeyer flasks should be washed with warm tap water and a small amount of
detergent. A bottle brush may be used to clean the interior surfaces.
-Rinse the Erlenmeyer flask thoroughly with tap water to remove detergent residue.
-Rinse the Erlenmeyer flask with a small amount of distilled water to remove remnants of tap
water inside the vessel before use or storage.

Applications:
-For this course, Erlenmeyer flasks will primarily be used for titrations. Briefly, the analyte (the
sample) is transferred to the Erlenmeyer flask, typically with an indicator added. The titrant is
then added using a burette (see Section 1E). As the titrant is being added, a reaction between the
analyte and titrant occurs. Continuous swirling of the Erlenmeyer flask as the titrant is being
added ensures that the reaction mixture is well mixed as the titration proceeds and the reaction
progresses. For more details on titrations, see Section III.G below, and Experiments 3, 4, 5 and 7.

29
C. Graduated Cylinder
Sometimes referred to as a ‘measuring cylinder’, the graduated cylinder is a vessel primarily
designed to measure and deliver a specific quantity of a liquid. The narrow diameter of the
graduated cylinder allows for greater accuracy and precision of the measured volume than a
beaker.

Graduated cylinders come in various sizes named for their maximum measurable capacity. e.g. a
100 mL graduated cylinder will have 100.0 mL as its highest graduation.

Typical graduated cylinders measured volumes in mL with precision to 0.1 mL, and the digit
after the decimal must be estimated.

e.g. A 25 mL graduated cylinder filled with the meniscus centered exactly on the 24 mL
graduation would be considered to contain 24.0 mL. If this volume were delivered to another
vessel, we would report that 24.0 mL of the liquid was delivered.

100

90

80

70

60

50

40

30

20

10

A 100 mL Graduated Cylinder

Applications:
-A graduated cylinder is used to measure and deliver liquids in situations that require greater
precision than a beaker will allow, but for experiments where greater precision is unnecessary.

-Graduated cylinders are often used in situations where the volume of a liquid to be measured
exceeds what can be achieved with volumetric glassware like pipettes (see Section III.D, below).

-A single graduated cylinder is capable of measuring a range of volumes, which is not


necessarily possible with the most accurate pipettes (see Sections III.D. and III.E., below). For
instance, it may be necessary to measure 75.5 mL of a liquid, but the laboratory currently lacks
any type of pipette with this capacity. Thus a graduate cylinder would be most appropriate
glassware to measure this volume of liquid.

General Guidelines:
-Select the smallest possible graduated cylinder that will allow you to measure and deliver the
necessary volume of liquid.

30
-Graduated cylinders often have a plastic bumper around the neck. This bumper is not for
measuring the level of liquid: it is designed to protect the cylinder in case the cylinder is knocked
over. Push the bumper all the way to the top of the neck of the cylinder - it may prevent the
cylinder from breaking if it is dropped or knocked-over.

Cleaning Guidelines:
-Graduated cylinders should be washed with warm tap water and a small amount of detergent. A
bottle brush may be used to clean the interior surfaces.

-Rinse the graduated cylinder thoroughly with tap water to remove detergent residue.

-Rinse the graduated cylinder with a small amount of distilled water to remove remnants of tap
water inside the vessel before use or storage.

Example Workflow:
e.g. i. Task: Measure 94.4 mL of tap water and transfer this water to an Erlenmeyer flask.

Suggested method:
a. Acquire a clean 100 mL graduated cylinder.

b. Acquire a clean beaker and fill it with tap water. This will be used as a transfer beaker to
deliver the tap water to your graduated cylinder.
c. Pour tap water into the cylinder until the desired volume is achieved.
d. Pour the water from the graduated cylinder into the Erlenmeyer flask. It would be reported
that 94.4 mL of water was transferred to the Erlenmeyer flask.
r
ate
W

100 100
100

90 90
90

80 80
80

70 70
70

60 60
60

50 50
50

40 40
40

30 30
30

20 20
20

10 10
10

0 0
0

a b c d

Flowchart for e.g. i, above.

e.g. ii. Task: Determine the total volume of a 3.5% NaCl (aq) solution prepared previously in a
beaker.

Suggested method:
a. Acquire a clean and dry graduated cylinder.

31
b. Carefully pour all of the solution from the beaker directly into the graduated cylinder.

c. Read the volume on the graduated cylinder. The uncertain digit must be estimated. In the
flowchart below, the meniscus was centered between the 66 and 67 mL graduations. The
uncertain digit must be estimated and reported. In this case, the solution’s total volume was
reported as 66.3 mL.

100 100 100

90 90 90

80 80 80

70 70 70

60 60 60

50 50 50

40 40 40

30 30 30

20 20 20

10 10 10

0 0 0

a b c

Flowchart for e.g. ii, above.

D. Volumetric Pipette
The noun ‘pipette’ is sometimes spelled “pipet” in certain texts. Confusingly, ‘pipet’ is also used
as a verb, meaning: to measure and transfer a liquid by means of a pipette.
In this manual, the term ‘pipette’ will be used to mean the noun, the instrument, and the term
‘pipet’ as the verb, the action of using the instrument.

Examples of proper use of language:


e.g. Select a pipette from the cupboard.
e.g. Pipet 10.00 mL of distilled water into a flask.

The volumetric pipette is a high-precision, calibrated piece of glassware designed to accurately


measure a volume of liquid and then deliver this volume to another vessel. The volumetric
pipette is a long thin glass tube with a bulge along its length. A pipette pump is necessary to fill
the pipette by drawing liquid up from the bottom. The narrow diameter tube allows the volume
to be measured with great accuracy. However, a given volumetric pipette is only calibrated for a
single volume; volumetric pipettes are calibrated to deliver 1, 5, 10, 25 or 50 mL, etc.

In order to deliver a given volume of a liquid, say 50.00 mL, the appropriate volumetric pipette
must be used, and it must be filled exactly to the graduation. If the level of the liquid is visibly
above or below the graduation, the volume of the solution delivered will be inaccurate. It is
impossible to estimate the volume if the level of the solution is above or below the graduation.

32
A volumetric pipette is considered to have a precision of 2 digits after the decimal. Thus, if a 25
mL pipette is filled to the graduation, and then this volume of liquid is delivered to another
vessel, it would be reported that 25.00 mL of the liquid was delivered.

25 mL

A 25 mL volumetric pipette

Applications:
-The volumetric pipette is used whenever liquids need to be measured and delivered with the
greatest accuracy and precision possible. e.g. when preparing the analyte for a titration, typically
a pipette is used to deliver the analyte to an Erlenmeyer flask.

General Guidelines:
-Pipettes are delicate glass tubes that will break if dropped or mishandled and should be treated
with exquisite care. If mishandled, a pipette can break in your hands; for example, applying two
much pressure when inserting the pipette into the pipette pump, or attempting to bend the pipette
will both result in the pipette shattering into dangerous fragments in your hands.

-In general, you should handle pipettes near the top, above the graduation (see the diagram
above). Avoid handling the pipette tip (the bottom of the pipette) as this may introduce
contaminants into your samples / solutions.
-Avoid handling the bulb of the pipette during measurement and delivery of liquids: the heat
from your hands can affect the temperature, density and thus measured volume of the liquid.

-The volumetric pipette must be clean. See cleaning guidelines, below.

-The liquid aspirated from a beaker or Erlenmeyer flask with a pipette. NEVER pipet from the
stock or reagent bottle as this risks contaminating these solutions.

33
-The liquid is aspirated using a pipette pump attached to the top of the pipette. See the figures
below for directions as to how to place the pipette pump onto the pipette. Hold the pipette near
the top (not at the bulb or at the bottom) and gently twist the top of the pipette into the pipette
pump. Only slight pressure is required to connect the pipette pump, and a slight
counterclockwise twist will ensure a secure fit. The pipette pump should be attached securely,
but it should not be so tight that it cannot be quickly and easily removed.

-As a rule, the pipette pump should fit the pipette snugly enough that it can support the weight of
the pipette, so that simply holding the pipette pump will not allow the pipette to fall.

Inserting the pipette into a pipette pump

-When the wheel on the pipette pump is turned, it generates negative pressure inside the pipette.
If the pipette tip is inserted in a body of liquid, the liquid will be drawn up into the pipette. See
the image below for directions as to how to operate the pipette pump to aspirate a liquid into the
pipette.

Holding the pipette with a pipette pump and operating the pipette pump.

-Do not allow liquid to enter the pipette pump. This will contaminate and damage the pipette
pump.

Aspiration with the Pipette


To draw liquid into the pipette, the pipette is held vertically with a pipette pump attached to the
top. The pipette tip is submerged under the surface of the liquid to be aspirated, and then the
pipette pump is used to generate negative pressure inside the pipette. This draws liquid up the
pipette.

To aspirate liquid with the pipette:


a. Obtain the liquid to be aspirated, and pour it into a small beaker. Make sure you have slightly
more liquid than you intend to aspirate. If you need to measure 25.00 mL with a pipette, you

34
should have at about 30-40 mL of the liquid in available in a small beaker. Otherwise you will
have difficulty pipetting the desired volume without introducing air bubbles. Place a pipette
pump on the top of the pipette as described above.

b. Supporting the pipette pump with one hand, and the top of the pipette (not the bulb) with the
other hand, submerge the pipette tip into the liquid to be aspirated.

c. Turn the wheel on the pipette pump with your thumb downwards. This operates the pipette
pump and draws the liquid into the pipette.

d. Continue to turn the wheel on the pipette pump until the desired level of liquid is achieved. Be
careful not to draw air into the pipette. Keep the pipette tip submerged under the level of liquid
during aspiration.

e. Once the desired level of liquid is achieved, remove the pipette from the container of liquid.

f. Remove the pipette pump. Once you remove the pump, the liquid will drain from the pipette,
unless you block the top of the pipette with your finger.

Flowchart for Aspiration of a volumetric pipette, as described above.

Cleaning Guidelines:
-The pipette should be washed by aspirating water. See the instructions for aspiration of the
pipette above. Aspirate the pipette with tap water and then drain it completely. Repeat this
process three times before rinsing with distilled water.

-The volumetric pipette should NOT be cleaned with wire brush.

-The clean pipette should be rinsed. First rinse the pipette with 10 mL of distilled water:

To rinse the pipette:


a. Attach the pipette pump

35
b. Aspirate a small volume (~ 5 mL) of the rinsing solution.

c. Remove the pipette from the solution.

d. Remove the pipette pump from the pipette. The liquid will drain from the pipette unless you
block the top of the pipette with your finger.

e. Put the pipette pump down, and using two hands, gently turn the pipette on its side. Slowly roll
the pipette as indicated below to sheet the interior surfaces with the rinsing solution.

f. Return the pipette to vertical over a waste container.

g. Drain the rinsing solution into a waste container by unblocking the top of the pipette with your
finger.

Flowchart for rinsing a volumetric pipette, as described above.

-The distilled water rinse should be repeated a total of three times for the best results.

-A cleaned and rinsed pipette should then be rinsed using the solution that is to be measured and
delivered. Follow the instructions above for rinsing a pipette with distilled water, but replace the
water with the solution to be measured and delivered. The pipette should be rinsed with the
solution at least once before that solution is measured and delivered.

Measurements:
-Once the pipette is cleaned and rinsed, the solution to be measured and delivered can be
aspirated.

-The pipette pump may be used to adjust the volume of liquid inside the pipette until it reaches
the graduation on the pipette, although this may be difficult.

36
-As a more consistent and reliable alternative to the above, a good technique is to aspirate the
pipette to a level above the graduation. Remove the pipette pump and block the opening at the
top of the pipette with your finger (or thumb). This will prevent the liquid from draining from the
pipette. Put the pipette pump down. Gently remove your finger in very small increments to let air
into the pipette; this will allow you fine control over the liquid being drained from the bottom of
the pipette. Slowly drain the pipette in this manner into a waste container until the level of the
liquid is exactly on the graduation. Once the meniscus is centered on the graduation, block the
top of the pipette with your finger to prevent further draining. The pipette now contains the
indicated volume.

-It is advisable to raise the pipette to eye level when reading the position of the meniscus. See
Reading the Meniscus, in Section III.H., below.

Delivery:
-Solutions measured by the volumetric pipette can be delivered by reversing the direction of the
pipette pump if you are using the pump to hold the liquid inside the pipette. If you are using your
finger to block the top of the pipette to retain the liquid, simply release your finger to deliver the
liquid.

-The volumetric pipette should be held vertically during delivery.

-Deliver the entire volume contained in the volumetric pipette. Do not attempt to stop the
delivery early to deliver a smaller volume of liquid.

-Once the all of the liquid contained in the pipette is delivered, gently touch the tip of the pipette
to the receiving vessel; if there is a drop of liquid clinging to the pipette tip, this will allow it to
drain into the receiving vessel.

-Do not blow out the last drop with your mouth, finger or the pipette pump. The volumetric
pipettes in our lab are calibrated to account for a small amount of liquid to be retained by the
pipette after delivery.

Example Workflow:
e.g. i. Task: Deliver 25.00 mL of 1.20 M acetic acid solution to an Erlenmeyer flask.

Suggested Method:
a-b. Pour 40 mL of the acetic acid solution into a 50 mL beaker. DO NOT pipet from the stock
bottle.

c. Acquire a 25 mL volumetric pipette. Clean the pipette as described above. Rinse the pipette
using ~ 5 mL of the acetic acid solution as described above. Attach the pipette pump to the top of
the pipette.

d. Insert the tip of the pipette into the solution.

e. Aspirate the pipette with the solution to a level above the graduation on the pipette.

37
f. Remove the pipette tip from the solution.

g. Remove the pipette pump and quickly block the top of the pipette with your finger.

h. Slowly drain the pipette by gently releasing allowing air to enter the top of the pipette until the
meniscus is centered on the graduation. The pipette now contains the indicated volume.

i. Keeping your finger blocking the top of the pipette, and preventing the liquid from draining,
move the pipette to the receiving vessel.

j. Unblock the top of the pipette to deliver the liquid into the receiving vessel. Once all of the
liquid is drained, touch the tip of the pipette to the inner wall of the receiving vessel: in case
there is a droplet of the liquid sticking to the pipette tip, this will allow the droplet to enter the
receiving vessel.

Flowchart for example i, above.

E. Mohr Pipette
The Mohr pipette is sometimes called a “measuring pipette” or “graduated pipette”, and also
occasionally spelled “pipet”.

The Mohr pipette is a calibrated piece of glassware designed to combine the attributes of a
burette and volumetric pipette. That is, like a volumetric pipette, the Mohr pipette can be easily
used to measure liquids and deliver these liquids to other vessels. However, while a given
volumetric pipette can only deliver a single volume, Mohr pipettes have a scale of graduations
like burettes, allowing them to deliver a range of possible volumes. Thus, like a burette, the
Mohr pipette has an inverted scale of graduations, and the volume delivered must be determined
by difference.

Like a volumetric pipette, a Mohr pipette is used with a pipette pump to aspirate the liquid from
another vessel, such as a beaker or Erlenmeyer flask. The initial reading of the level of the liquid
is taken. The pipette is then used to deliver the liquid. Delivery is stopped by the operator at (or

38
as close as possible to) the required final volume. The final volume reading is taken. The actual
volume delivered is taken to be the difference between the final and initial volumes.

Importantly, the Mohr pipette readings can never be accurately taken if they fall outside the scale
of graduations. Thus the initial reading can never be less than 0.00 mL, and the final reading can
never be more than the maximum volume of the Mohr pipette.

The main advantage of a Mohr pipette over a volumetric pipette is that the Mohr pipette allows a
much wider range of volumes to be delivered, and in the hands of a skilled operator the Mohr
pipette is more accurate than a graduated cylinder. However, the Mohr pipette is somewhat more
difficult to use, and slightly less accurate than a volumetric pipette.

The Mohr pipette is considered to have precision of 2 digits after the decimal. The 2nd digit after
the decimal must be estimated when using a Mohr pipette.

e.g. If the Mohr pipette was filled, and the meniscus is centered exactly on the 2 mL graduation,
the initial volume should be reported as 2.00 mL.

e.g. If the initial Mohr pipette reading was 2.00 mL, some liquid was delivered, and the final
pipette reading is 3.57 mL, the volume of liquid delivered would be taken to be 3.57 - 2.00 mL =
1.57 mL.

e.g. If the initial Mohr pipette reading was 0.70 mL, some liquid was delivered, and the final
pipette reading is 7.40 mL, the volume of liquid delivered would be taken to be 7.40 mL - 0.70
mL = 6.70 mL, not 6.7 mL.

The Mohr pipette is similar to, and often confused with its close cousin the serological pipette,
used frequently in biology.

39
0

10

A 10 mL Mohr pipette

Applications:
-A Mohr pipette is a convenient device for delivering liquids in a range of volumes with
accuracy similar to a volumetric pipette. For example, a 10 mL Mohr pipette can deliver any
volume from 0 to 10 mL with two-decimal precision; e.g. 2.54 mL, which is impossible to
deliver with a volumetric pipette, because volumetric pipettes are manufactured only in standard
volumes, such as 1, 2, 5, 10, 25 mL, etc.

General Guidelines:
-In many respects, working with the Mohr pipette is similar to working with a volumetric pipette
(see Section III.D, above).

-Mohr pipettes are expensive and delicate, and like all glassware, dangerous if broken. Handle
the Mohr pipette with exquisite care. If mishandled, a Mohr pipette can break in your hands; for
example, applying two much pressure when inserting the pipette into the pipette pump, or
attempting to bend the pipette will both result in the pipette shattering into dangerous fragments
in your hands.

-Liquids are aspirated using a pipette pump placed on the top of the Mohr pipette. See Section
III.D for instructions on inserting the pipette into a pipette pump.

-While a volumetric pipette has only a single graduation, a Mohr pipette has a scale of
graduations. Thus a Mohr pipette can be used to measure and deliver a range of possible volumes.

40
-The scale of graduations on a Mohr pipette is inverted. The uppermost graduation is 0.00 mL.
Reading the Mohr pipette correctly and consistently is crucial for accurate results. Some
examples of reading the Mohr pipette’s scale correctly are shown below. Also see Reading the
Meniscus in Section III.H.

Aspirating liquids with the Mohr pipette:


a. Obtain the liquid to be aspirated. Make sure you have slightly more liquid than you intend to
pipet. If you intend to aspirate 10.00 mL, you should have at least 20 mL of the liquid available
in a small beaker. Otherwise you will have difficulty pipetting the desired volume without
introducing air bubbles. Place a pipette pump on the top of the Mohr pipette.

b. Supporting the pipette pump with one hand, and the top of the pipette with the other hand,
submerge the pipette tip into the liquid to be aspirated.

c. Turn the wheel on the pipette pump with your thumb downwards. This operates the pump and
draws the liquid up the pipette. Continue to turn the wheel on the pipette pump until the desired
level of liquid is achieved. Be careful not to draw air into the pipette. Keep the pipette tip
submerged under the level of the liquid during aspiration.

d. Once the desired level of liquid is achieved, remove the pipette tip from the container of liquid.

e. Remove the pipette pump. Once you remove the pump, the liquid will drain from the pipette,
unless you block the top of the pipette with your finger.

Flowchart for aspirating a Mohr pipette, as described above.

41
Cleaning guidelines:
-Methods for cleaning the Mohr pipette are consistent with cleaning a volumetric pipette.

-The Mohr pipette should NOT be cleaned with a wire brush.

-The pipette should be washed by aspirating tap water and then draining it completely. Make
sure that all of the interior surfaces are cleaned, but do not allow water to enter the pipette pump.
Repeat this process several times before rinsing with distilled water.

-The cleaned Mohr pipette must be rinsed.

To rinse the Mohr pipette:


a-b. Attach a pipette pump, and submerge the tip of the Mohr pipette into the rinsing solution.
Aspirate a small volume (~5 mL) of the rinsing solution.

c. Remove the pipette tip from the rinsing solution.

d. Remove the pipette pump. The solution will drain from the pipette unless you block the top of
the pipette with your finger.

e. Put the pipette pump down, and using two hands carefully turn the pipette on its side. Gently
rotate the pipette to sheet the interior surfaces with the rinsing solution.

f. Return the pipette to vertical.

g. Drain the rinsing solution into a waste container by unblocking your finger from the top of the
pipette.

Flowchart for rinsing a Mohr pipette, as described above.

42
-The pipette should be rinsed with ~ 5 mL of distilled water a total of three times for the best
results.

-A cleaned and rinsed pipette should then be rinsed using the solution that is to be measured and
delivered. Follow the instructions above for rinsing a pipette with distilled water, but replace the
water with the solution to be measured and delivered. The pipette should be rinsed with the
solution at least once before that solution is measured and delivered.

Measurement of liquids:
-The cleaned and rinsed Mohr pipette can now be used to measure and deliver liquids. Aspirate
the liquid as described above.

-It is not necessary to always aspirate the liquid to the uppermost (0.00 mL) graduation, although
this will be required if you need to deliver the maximum volume indicated for that Mohr pipette.
As a rule, it is advisable to aspirate slightly more liquid than you will need to deliver.

e.g. If you need the initial pipette reading to be about 1 mL, it is advisable to aspirate liquid up to
the 0.00 mL graduation.
e.g. If you need the initial pipette reading to be 0.00 mL, you will need to aspirate liquid above
this graduation.

-Remove the pipette pump and block the top of the Mohr pipette with your finger. Some liquid
will drain from the pipette during this process.

-Put down the pipette pump and read the position of the meniscus on the scale of graduations.
You may drain some liquid into a waste container if you wish to adjust the level of the liquid to a
certain level. Carefully remove your finger from the top of the pipette to drain the liquid slowly
to adjust its level to the desired position, blocking the top of the pipette when you want to stop
draining.

-Take and record the initial pipette reading. It is best to raise the pipette to eye level to read the
meniscus properly. See Reading the Meniscus in Section III.H., below. Remember that the scale
on the Mohr pipette is inverted, and must be read accordingly.

e.g. i. Pictured below is a section of a Mohr pipette, aspirated with a solution, at eye level. Note
the inverted scale of graduations. The meniscus is centered on the 3.3 mL graduation. Since the
meniscus appears to be exactly on this graduation, the uncertain digit would be estimated to be 0.
The pipette reading would be reported as 3.30 mL.

43
3

Reading the position of the meniscus in a


Mohr pipette, as described in e.g. i, above.

e.g. ii. In the figure below, we are looking at a section of the Mohr pipette, aspirated with a
solution, at eye level. Note the inverted scale of graduations. The meniscus is centered between
the 4.1 and 4.2 mL graduations. We must estimate the uncertain digit. We would estimate the
uncertain digit in this case to be 6, because it appears to be slightly below the point exactly half-
way between the graduations. Thus we would report the pipette reading to be 4.16 mL.

Reading the position of the meniscus in a


Mohr pipette, as described in e.g. ii, above.

Delivery:
-Since the Mohr pipette is meant to measure the volume of liquid delivered by difference, it is
crucial that the level of the liquid inside the pipette is held constant long enough for you to take a
reading. Ensure that the level of the liquid is constant, and take your initial reading as described
above.

-The Mohr pipette should be nearly vertical as you deliver the liquid.

-Remove your finger from the top of the pipette to deliver the liquid to the receiving vessel.
Deliver the solution until the level of the solution in the Mohr pipette reaches the intended final
volume. Once this volume is reached, block the opening at the top of the pipette with your finger
to stop the delivery. Read and record the final pipette volume, as described and outlined in the
examples above.

44
-DO NOT completely drain the Mohr pipette: for delivery of measured volumes the Mohr pipette
must not be drained below the lowermost graduation.

-Calculate the volume delivered as the difference between the final and initial readings.

-Importantly, the Mohr pipette cannot measure volumes for which readings lay beyond the
graduations of the pipette. Thus, the initial volume can never be above the uppermost graduation
on the pipette and the final volume can never be below the lowermost graduation.

Example Workflow:
e.g. iii. Task: Measure and deliver 7.45 mL of 1.195 M NaOH (aq) solution to an Erlenmeyer
flask.

Suggested Method: The volume required is not a standard volume for which a volumetric
pipette exists. Thus using a Mohr pipette is the best option for carrying out the required task.
a. Pour 25 mL of the stock solution into a beaker. Do NOT pipet directly out of the stock bottle.
Clean and rinse the Mohr pipette as described above. Place a pipette pump onto the top of the
Mohr pipette.

b. Insert the tip of the Mohr pipette into the solution to be measured, and aspirate the solution by
turning the wheel on the pipette pump downwards.

c. You can aspirate to any level that you feel is convenient which will allow you to deliver a total
of 7.45 mL. In this example, the user aspirated to a level with an initial reading of 1.25 mL.

d. Remove the Mohr pipette tip from the solution.

e. Remove the pipette pump and block the top of the pipette with your finger.

f. Important: take the initial reading at this point. It is likely some liquid will escape when you
remove the pipette pump. In this case, the initial reading is 1.40 mL. Record the initial reading.

g. Determine what the final reading should be to deliver the desired volume of liquid. In this
example, the final reading would need to be 8.85 mL.

h. Deliver the solution to the Erlenmeyer flask by removing your finger from the top of the
pipette. DO NOT completely drain the Mohr pipette. Watch the level of the liquid closely. Block
the opening at the top of the pipette once the level of the liquid reaches a reading of 8.85 mL.

i. Once the desired volume is delivered, take the final pipette reading. If the reading is different
from the desired reading, make sure to record the actual final reading. In this example, the actual
final reading was 8.90 mL. Thus the actual volume delivered was 8.90 mL - 1.25 mL = 7.50 mL.

45
Flowchart for e.g. iii, above.

F. Volumetric Flask
The volumetric flask is a high-precision, calibrated piece of glassware designed to prepare
solutions of a given final volume. Each volumetric flask is individually calibrated to contain the
volume indicated on the glassware; standard volumetric flasks are calibrated to contain 1, 5, 10,
50, 100 or 250 mL, etc.

Correspondingly, a given volumetric flask has only 1 graduation. In order to prepare a solution
of a given volume, say 50.00 mL, the solution must be prepared in the appropriate volumetric
flask and diluted exactly to the graduation. If the level of the solution is visibly above or below
the graduation, the volume and thus the concentration of the solution will be inaccurate. It is
impossible to estimate the volume if the level of the solution is above or below the graduation.

Volumetric flasks are considered to have a precision of 2 digits after the decimal. Thus, when a
100 mL volumetric flask is filled exactly to the mark with solution, it would be considered to
contain 100.00 mL of solution.

Importantly, volumetric flasks are not designed to deliver a volume of solution with high
precision; they are merely designed to measure and contain a single, specific volume. If delivery
of a specific volume is necessary, one would typically use a graduated cylinder (Section C) or a
pipette (Sections E and F).

46
100 mL
50 mL
Volumetric flasks in two sizes: 50 mL and 100 mL.

Applications:
A volumetric flask is the main method of preparing solutions for which the total volume of
solution must be accurately known.

Notably, the volume of solution typically does not equal the volume of the solvent. That is, when
a solvent is combined with a solute (or when multiple solvents are combined) the solutions may
expand or contract slightly relative to the pure solvent(s). Thus, if 5 g of a solid solute were
combined with 50.00 mL of distilled water delivered from a volumetric pipette, it IS NOT safe to
assume that the volume of solution is accurately 50.00 mL. Likewise, if a solution is prepared
with a final volume of 100.00 mL, it IS NOT safe to assume that the volume of solvent used was
exactly 100.00 mL.

If a solution with a specific final volume is required (for instance, to prepare a solution of a given
molar concentration), the final volume of the solution must be accurately measured. In this case,
an accurate knowledge of the volume of the solvent is not necessary. The volumetric flask is
specifically designed to allow the preparation of a solution of a known final volume, and thus a
known molar concentration.

The volumetric flask is useful for preparing solutions of known molarity from chemically pure
reagents, or by diluting concentrated stock solutions. Both of these workflows are detailed below.

General Guidelines:
-The volumetric flask is somewhat delicate. Its long neck will break if dropped or if too much
pressure is applied when trying to cap the flask.

-Capping the volumetric flask with a snap-type cap is an art. We suggest cradling the neck of the
flask in with both hands (with your fingers interlocked) and applying downward pressure on the
cap with both thumbs. See your instructor for suggestions and assistance.

-Generally a solution containing the solute is introduced to the volumetric flask first. Then the
solution is diluted with solvent to the final volume of the flask, indicated by a single graduation.
For the purposes of this course, the solvent will be distilled water. However, other solvents or
even other solutions may be used.

47
Cleaning Guidelines:
-The volumetric flask should not be cleaned with a bottle brush.

-Flush the inside of the volumetric flask with warm tap water and a small amount of detergent.

-Rinse the inside of the volumetric flask 5-6 times with warm tap water to remove all detergent
residue.

-Rinse the volumetric flask 3-4 times with 10-15 mL of distilled water.

Example Workflows:
e.g. i. Task: Dissolve 8.77 g of NaCl (s) in distilled water and prepare an NaCl (aq) solution
with a final volume of 100.00 mL. The resulting solution should have a concentration of NaCl
(aq) of 1.50 M.

Suggested method:
a-b. Use a balance to weigh 8.77 g of NaCl (see Analytical Balance, Section 1 above) into a
tared beaker. If the goal is to produce a specific solution with a specific concentration, you will
want to approach 8.77 g as closely as possible. In this example, a total of 8.791 g of the solid was
acquired according to the analytical balance.

c. Add distilled water to the beaker. Use a volume of distilled water less than the desired final
volume of solution (e.g. about 2/3 of the final volume).

d. Swirl the mixture until the solute completely dissolves and the mixture is homogenous.

e. Transfer the solution to the volumetric flask. You may optionally use a liquid funnel in the top
of the volumetric flask. Your goal is to pour all of the solution into the volumetric flask without
any spilling or loss of the material. Rinse the beaker with a small volume (~ 5 mL) of distilled
water and transfer the rinse into the volumetric flask. Repeat the rinsing 2 or 3 times for the best
results.

f-g. Gently swirl the volumetric flask. Then add distilled water to the volumetric flask carefully.
The neck of the flask is very narrow and will fill rapidly. Add the last few drops of water slowly
until the meniscus of the solution is centered on the graduation of the volumetric flask.

h. Cap the volumetric flask and invert the flask 5-10 times to thoroughly homogenize the mixture.
In this example, given the starting mass of 8.791 g of NaCl, the NaCl (aq) solution had a final
volume of 100.00 mL and a molar concentration of 1.504 M.

48
Flowchart for e.g. i, above.

e.g. ii. Task: Prepare 100.00 mL of a 0.375 M solution of NaCl (aq) from a stock solution of
NaCl with a concentration of 1.50 M.

Suggested Method:
This scenario requires 25.00 mL of the 1.500 M NaCl (aq) solution to be diluted to a final
volume of 100.00 mL.

a. Acquire a 25 mL volumetric pipette. Clean and rinse the pipette as described above.

b. Pour about 40 mL of the 1.50 M NaCl solution into a beaker. Do NOT pipet out of the stock
bottle.

c-d. Aspirate the volumetric pipette with the NaCl solution so that the level of the liquid is above
the level of the graduation.

e. Remove the pipette tip from the solution.

f. Remove the pipette pump. Block the top of the pipette with your finger to prevent the solution
from draining. Carefully release your finger to slowly drain the liquid until the meniscus is
centered on the graduation.
g. Move the pipette to the volumetric flask.

h. Deliver the liquid to the volumetric flask by unblocking the top of the pipette with your finger.
You may optionally use a liquid funnel in the top of the volumetric flask. Your goal is to deliver
all of the solution into the volumetric flask without any spilling or loss of the material.
Completely drain the volumetric pipette. Touch the tip of the pipette to the edge of the
volumetric flask (or liquid funnel, if you are using one) to allow the last droplet stuck to the
outside of the pipette to drain.

i. Dilute the solution in the volumetric flask with distilled water. Swirl the flask before the level
of the liquid reaches the neck.

j. Continue to fill the volumetric flask with distilled until the meniscus is centered on the
graduation. The neck of the volumetric flask will fill very quickly, so be careful.

49
k. Cap the volumetric flask and invert the flask 5-10 times to thoroughly homogenize the mixture.
The final volume of the solution is thus 100.00 mL, and the concentration of the solution is 1/4
the initial concentration.

Flowchart for e.g. ii, above.

G. Burette
Burette is sometimes spelled ‘buret’.

The burette is a high-precision, calibrated piece of glassware. A burette is essentially a long,


narrow diameter glass tube.The tube’s narrow diameter facilitates greater accuracy of
measurement of the volume of liquid the glassware contains. The burette is open at the top,
which allows liquids to be added, and has a stopcock (valve) at the bottom that controls the
delivery of the liquid from the burette to another vessel. When the valve is in the open position,
the burette will continuously deliver solution. In order to stop delivery, the stopcock must be
manually closed.

Notice that unlike beakers, Erlenmeyer flasks and graduated cylinders, the burette has an
inverted scale of graduations; the uppermost graduation is 0.00 mL, like the Mohr pipette. Thus,
the burette is designed to deliver liquids measured by difference much like a Mohr Pipette (see
Section III.E).

Thus to determine the volume of a liquid delivered, the level of the liquid inside the burette must
be read at two points; 1. Once the burette is filled, the initial volume reading is taken; 2. after the
liquid has been delivered, the final volume is taken. The volume of liquid delivered is taken to be
the difference between the final and initial volumes.

Importantly, the burette readings can never be accurately taken if they fall outside the scale of
graduations. Thus the initial reading can never be less than 0.00 mL, and the final reading can
never be more than the maximum volume of the burette, typically 50.00 mL.

A burette is considered to have precision of 2 digits after the decimal. The 2nd digit after the
decimal must be estimated when using a burette.

50
e.g. If the burette was filled, and the meniscus is centered exactly on the 12 mL graduation, the
initial volume should be reported as 12.00 mL.

e.g. If the initial burette reading was 12.00 mL, some liquid was delivered, and the final burette
reading is 33.92 mL, the volume of liquid delivered would be taken to be 33.92 - 12.00 mL =
21.92 mL.

e.g. If the initial burette reading was 0.55 mL, some liquid was delivered, and the final burette
reading is 12.45 mL, the volume of liquid delivered would be 12.45 - 0.55 mL = 11.90 mL, not
11.9 mL.

A 0 B

10

buret
20
clamp

20

30

40

50

stopcock

buret stand

A) A 50 mL burette. B) A burette clamped into a burette stand.

Applications:
A burette is most commonly used for titrations. A titration is a method of analysis that will allow
you to determine the endpoint of a reaction and therefore the quantity of reactant in the reaction
mixture required to complete the reaction. The goal is to deliver just enough of the reactant (the
titrant) to reach the reaction’s completion, and no more. This allows for quantitative analysis of
the sample (the analyte) that reacted with the titrant. See Experiments 3, 4, 5 and 7 for more
information on titrations.

To carry out a titration, the burette is clamped into a stand that supports it, as pictured above.
This allows the operator’s hands to be free to carry out the titration. The titrant (a solution of one

51
of the reagents in the reaction being studied) is poured into the burette. The analyte (the other
reagent, the sample being analyzed) is prepared in an Erlenmeyer flask. The burette is positioned
so that the titrant can be delivered to the Erlenmeyer flask with the operator controlling the
stopcock’s position. Thus the titration is carried out as the operator delivers the titrant from the
burette by controlling the delivery of this solution.

General Guidelines:
-The burette is a long delicate glass tube. Take appropriate care with the burette, as mishandling
this instrument will cause it to chip or break. The ends of the burette are especially delicate and
sensitive to being dropped.

-The burette should be clamped into a burette stand before any liquids are added.

-The burette should be clamped into the burette stand in order to operate the stopcock and drain
liquids

-Fill the burette from the top. Place a liquid funnel into the top of the burette, and pour the
desired liquid into the funnel

-Avoid pouring from the stock bottle. It is best to pour some stock solution into a clean beaker,
and then use the beaker to pour solution into the burette.

-Reading the burette correctly and consistently is crucial for accurate results. Some examples of
reading the scale of the burette correctly are shown below. Also see Reading the Meniscus in
Section III.H.

Cleaning Guidelines:
-The burette should not be cleaned with a bottle brush.

-The burette can be cleaned by flushing with tap water. Fill the burette with tap water, drain
some water through the stopcock, and pour the remaining water out the top of the burette. Repeat
this process at least 3 times to completely flush the burette.

-The wash should always be followed by rinsing the burette with distilled water.

To rinse the burette:


a-c. Approximately 10 mL of distilled water will suffice for one rinse; pour some distilled water
into the burette.

d. Open the stopcock to flush some distilled water through the valve and the tip of the burette
into a waste container.

e-f. Remove the burette from its clamp and carefully turn the burette on its side to sheet the inner
surfaces with the distilled water.

52
g-h. Pour the rinse out the top of the burette, or drain it through the stopcock into a waste
container.

Repeat the rinsing process a total of three times for the best results.

Flowchart for rinsing a burette, as described above.

-Important: The clean burette should be rinsed with the titrant before it is filled with the titrant.
Follow the instructions above for rinsing the burette, but replace distilled water with the titrant
solution. Approximately 10 mL of the titrant should be poured into the burette, some of this
liquid should be flushed through the stopcock. The burette should be removed from the burette
clamp, and carefully turned on its side to sheet the interior surfaces with the titrant. The rinse can
then be discarded into a waste container. The burette should be rinsed with titrant at least once
before it is filled with the titrant.

Delivery and Titration:


-Titration is the delivery of titrant from the burette to a receiving vessel. Most often the titrant is
one of the reagents that occurs in a chemical reaction with the analyte. See Experiments 3, 4, 5
and 7 for details.

-The goal is to quantitatively determine the volume of titrant delivered to the analyte to complete
the reaction. Care must be taken to read the burette correctly and consistently to achieve accurate
results with suitable precision.

-Every titration is somewhat specific to the reaction being carried out. What follows is a general
workflow for the experiment:

a-b. With the washed and rinsed burette in the burette clamp, place a liquid funnel in the top of
the burette.

c. Fill the burette with titrant using a liquid funnel as described above. Avoid pouring the titrant
from the stock bottle. Instead, use a small beaker to pour the titrant into the burette. DO NOT fill

53
the burette above the uppermost graduation. It is not necessary to fill the burette to a level where
the reading is exactly 0.00. This is unnecessary and a waste of time.

d. Place a waste container beneath the burette. Open the stopcock for a few seconds to drain
some of the titrant. This will flush the stopcock with the titrant, and clear any air bubbles. Close
the stopcock.

e. Remove the liquid funnel from the top of the burette before taking any readings, and set it
aside. It is best to titrate without the liquid funnel. The liquid funnel may impede air from
flowing into the top of the burette, making delivery of the titrant slow. The liquid funnel may
also gradually drip excess liquid into the burette, resulting in errors in burette readings.

f. Take the initial burette reading. Carefully read the position of the meniscus. Remember that the
burette has an inverted scale of graduations, starting with 0 on the top. Some examples are shown
below. Burette readings are always taken in mL to 2 decimal places; the uncertain digit must be
estimated. Record the initial burette reading. See Reading the Meniscus in Section III.H for
guidance.

g. Deliver the titrant to the receiving vessel (typically an Erlenmeyer flask containing the analyte)
by opening the stopcock. During delivery, the stopcock is operated with one hand and the
Erlenmeyer flask is continuously swirled with the other. Deliver the titrant directly into the
analyte solution: avoid splashing the titrant on the walls or mouth of the Erlenmeyer flask.

It is possible to adjust the rate of titrant flow by fine adjustments of the stopcock. It is possible to
deliver the titrant very rapidly, or slowly or even dropwise, depending on the stopcock setting.
Close to the endpoint of the titration, you will want to reduce the rate of delivery of the titrant so
that the titration can be stopped as close to the endpoint as possible.

h. During a titration, an indicator will indicate when to stop delivery. See Experiments 3, 4, 5
and 7 for examples. To stop delivery, return the stopcock to the closed position.

i. Take the final burette reading by reading the position of the meniscus. Remember, the burette
reading is taken in mL, to 2 decimal places, and the uncertain digit must be estimated.

54
Flowchart for delivering the titrant during a titration using a burette, as described above.

Reading the Burette:


The position of the meniscus must be carefully and consistently read for accurate results when
delivering liquids with a burette. Importantly, you should adjust the height of the burette so that
the meniscus is at eye level before taking a reading. See Reading the Meniscus in Section III.H
below for general guidelines.

Some specific examples of reading the inverted scale of graduations on a burette are detailed on
the following page:

55
e.g. i. Pictured below is a section of a burette, filled with a solution, at eye level. Note the
inverted scale of graduations. The meniscus is centered exactly on the 14.0 mL graduation. Since
the meniscus appears to be exactly on this graduation, the uncertain digit would be estimated to
be 0. The burette reading would be reported as 14.00 mL.

10

20

Reading the position of the meniscus in a burette, as described in e.g. i, above.

e.g. i. Pictured below is section of a burette, filled with a solution, at eye level. Note the inverted
scale of graduations. The meniscus is centered between the 30.2 and the 30.3 mL graduations.
The uncertain digit would be estimated to be 7, because the meniscus is clearly below the point
exactly half-way between the graduations. Thus the burette reading would be reported to be
30.27 mL.

20

30

40

Reading the position of the meniscus in a burette, as described in e.g. i, above.

56
H. Reading the meniscus

For accurate measurements of liquid volumes using graduated glassware, the volume of the
solution should be read at the bottom of the meniscus. In beakers and Erlenmeyer flasks, it may
not be possible to see a meniscus at the surface of the liquid, because the diameters of these
vessels are so large. However, when using graduated cylinders, volumetric pipettes, Mohr
pipettes, burettes and volumetric flasks, a clear meniscus should be visible at the surface of the
liquid. See the examples below.

When taking measurements with these glassware, the volume of the liquid should be read at the
bottom of the meniscus. Raise the glassware, or adjust your position so that the meniscus is at
eye level.

Placing white piece of paper immediately behind the glassware will help you to clearly
distinguish the bottom of the meniscus, so its position can be accurately and consistently read. A
white piece of paper with a black piece of tape positioned so that the black piece of tape is just
below the meniscus is even more effective. See panels a. and b. in the figure below.

Your finger may be substituted for the black mark on the white card but it is not as effective.

A. B.

The meniscus of a volume of liquid in a burette. A) The position of the meniscus can
be read with some difficulty with no coloured card in the near-background. B) The
optimal way to see and determine the position of the meniscus is by using a white piece
of paper or index card with a black mark or piece of tape. By positioning the black
mark just below the meniscus, its position is visually highlighted. In this example, the
position of the meniscus is 12.69 mL.

Parallax Error

Parallax error is an example of systematic error that may occur when reading the position of the
meniscus in graduated glassware. A parallax error may occur when you fail to read the position
of the meniscus at eye level. If you view the meniscus from above or below, it becomes difficult
to accurately assess its position, and a parallax error results.

57
In panel c. below, the meniscus in a burette is being read from a perspective above the liquid
level and the bottom of the meniscus appears to be at 15.30 mL. In panel d. below, meniscus in
the same burette is being read from a perspective below the liquid level, and the bottom of the
meniscus appears to be at 15.60 mL.

C. D.

Parallax error caused by viewing the meniscus from a perspective C) above or D) below the
surface of the liquid. The meniscus should be read at eye level to avoid this error.

I. Miscellaneous Glassware
Your drawer of laboratory equipment contains several pieces of ungraduated glassware not
described in great detail in the following sections. The most important are shown here:

liquid funnel test tube watch glass

Of these, the liquid funnel will be used most extensively in this course. Mainly, the liquid funnel
should be used to help you pour liquids from beakers into burettes. If you wish, you may also use
the liquid funnel to help you pour liquids into volumetric flasks. (See Sections III.D, III.E and
III.F above).

Test tubes are non-volumetric glassware designed to contain small amounts of liquids.

A watch glass is a small glass dish, typically NOT used as a vessel for containing liquids.
Typically a watch glass will be used to contain a small amount of a solid.

A full list of all the glassware available in your drawer appears at the beginning of this manual.

58
IV. Data Collection and Analysis

A. Significant Figures
Most numbers used in science are based upon measurements. Since all measurements have a
degree of uncertainty, it is important to only report those numbers that are meaningful, or
significant. Thus, manner in which a measurement is recorded conveys important information
regarding its level of uncertainty.

Significant figures are the number of digits used to express a measurement and its uncertainty.
The greater the number of significant figures after the decimal place, the more precise the
measurement.

To illustrate, examine the following two measurements:

1.0 mL 1.00 mL
At first glance, one might consider these measurements to be identical, but the manner in which
they are written, (the first has two significant figures and the second has three significant figures),
provides an indication of the uncertainty of each. The second value, with three significant figures,
is communicating more certaintly (less uncertainty) than the first value, which has only two
significant figures. This is because the last digit reported in any measured value is uncertain. We
call this the uncertain digit. In any measurement, the uncertain digit is estimated and thus
presumably is the most subject to error. As a rule, we typically assume that the uncertain digit
has an uncertainty of ± 1 in any reported value unless otherwise specified. With this in mind, we
will reevaluate the two measured values from above:
1.0 mL 1.00 mL
Uncertain Digit: Tenth of a mL Hundredth of a mL
Uncertainty: ± 0.1 mL ± 0.01 mL
Range of Uncertainty: 0.9 – 1.1 mL 0.99 – 1.01 mL

Thus, our uncertainty in the first value (two significant figures) is much higher than our
uncertainty in the second (three significant figures). In the first measurement the range of
uncertainty is relatively large: the actual volume could be anywhere between 0.9 and 1.1 mL,
i.e. we should assume that up to 0.1 mL of error exists in either direction (10% of 1 mL). On the
other hand, in the second measurement, the range of uncertainty is comparatively small: the
actual volume would be assumed to be somewhere between 0.99 and 1.01 mL (1% of 1 mL). In
other words, the first measurement (two significant figures) carries roughly 10-fold more
uncertainty than the second measurement (3 significant figures).

Thus, the number of significant figures used when recording a measurement is not at the
discretion of the experimenter: it is determined by the uncertainty of the equipment used and
must be accurately reported as such.

59
To explain further, and illustrate some additional important points about significant figures, we
will conisder another scenario. Consider the level of the meniscus in the burette shown below:

Here, we know the meniscus lies between graduations indicating 20.1 and 20.2 mL. Hence, we
are confident that the first three digits will be 20.1 mL. These digits are certain. However,
volumes measured with burettes of this kind are always reported with two digits after the
decimal. In other words, for a burette of this kind, the uncertain digit is the hundredth of a mL.
We must estimate the last digit in the value, in the hundredths position.

Based on the level of the meniscus in the diagram above, we would likely estimate the uncertain
digit to be 5 hundredths of a mL. Alternatively, 4 or 6 hundredths would also be acceptable. The
value of the uncertain digit varies slightly depending upon the subjective interpretation of the
person making the measurement, but the uncertain digit must always be reported: it cannot be
ignored or eliminated, even if it is 0.

Importantly, the certain digits are not open to interpretation and should always be agreed upon.

Thus the volume reading in this scenario would be reported as 20.15 mL (or 20.16 or 20.14 mL,
depending on the person making the measurement).

Recall, the greater the number of significant figures after the decimal place, the more precise the
measurement. Reporting a value with the correct number of significant figures is an important
means of communicating the precision of the measurement. This communication is critical.
When reading scientific documents, trained scientists will assume that all reported digits in
measured values therein are signficant. So when gathering data and reporting these values in lab
reports, textbooks, scientific articles etc. it is crucially important that the authors report each
value therein with the correct number of significant figures so as not to exaggerate nor understate
the true precision of these measurements.

60
The volume reading described above has a total of 4 significant figures: 3 certain digits (not
estimated), and 1 uncertain digit (estimated). It is crucial that this value be reported with all 4
significant figures intact: it should not be rounded to 20.1 or 20.2 mL – this would be
understating the precision of the measurement. And it should not be reported as 20.150 or
20.1500 mL – this would be exaggerating the precision of the measurement. Understatement and
exaggeration of the precision are both undersirable: they are akin to mistruths and scientists must
avoid this in all scientific documents.

For a final example, let us consider the difference between volume measurements made with two
very different types of laboratory glassware:

10.00 mL (made with a volumetric pipette) and 10 mL (made with a beaker).

These two measurements convey different information. The actual volume in the 10 mL value
should be interpreted as being somewhere in the range 9 mL and 11 mL (remember, the
uncertain digit is estimated). In contrast, the actual volume in the 10.00 mL value should be
interpreted as being somewhere in the range of 9.99 mL and 10.01 mL. In other words, the
10.00 mL measurement carries higher precision than the 10 mL measurement and this is the
direct result of the different glassware used to measure these volumes: the pipette (which affords
2 digits after the decimal) measures liquid volumes with much greater precision than does the
beaker (which affords no digits after the decimal). The limitations of each measuring device are
communicated in the reported values measured by these devices by correctly reporting the
number of significiant figures permitted by each.

Rules for Counting Significant Figures

Recall, anyone reading a scientific document has to assume that any values reported therein are
reported with the correct number of significant figures. What follows is a list of rules scientists
use in the interpretation of the significance of any reported value:

1. Nonzero integers always count as significant figures.

Example Number of significant figures


567.98 5

2. Leading zeros are zeros that precede all the nonzero digits. These never count as significant
figures; these zeros serve only as space holders. They are there to put the decimal point in its
correct location. They do not involve measurement decisions. When we write the numbers in
scientific notation, the non-significant zeros disappear.

Example Number of significant figures Written in scientific notation


0.00234 3 2.34 x 10-3
0.4 1 4 x 10-1
0.000035709 5 3.5709 x 10-5

61
3. Captive zeros are zeros between nonzero digits. These always count as significant figures.

Example Number of significant figures Written in scientific notation


10.06 4 1.006 x 101
300.4008 7 3.004008 x 102
0.0030505 5 3.0505 x 10-4

4. Trailing zeros are zeros at the right end of the number. They are significant only if the number
contains a decimal point.

Example Number of significant figures Written in scientific notation


25 000 2 25 x 103
300 1 3 x 102
300. 3 3.00 x 102
300.00 5 3.0000 x 102
567.89000 8 5.6789000 x 102

5. Exact numbers are not obtained using measuring devices. They can be determined by counting,
or they can be defined numbers. Exact numbers can be assumed to have an infinite number of
significant figures: they never limit the number of significant figures in a calculation.

Examples of exact numbers


3 measurements
12 experiments
6.022 x 1023 atoms in a mole
12 inches = 1 foot
3 moles of H2 react with 1 mole of N2 to form 2 moles of NH3

Rules for Calculations with Significant Figures

When performing calculations, several mathematical rules exist which take into account the
certainty of measurements. Your answer cannot be MORE precise than the least precise
measurement.

 Multiplication or Division
The result has the same number of significant figures as the least precise measurement.
Final answer reported
Example Rationale with correct number
of sig figs
5.78 has 3 sig figs
5.78 x 2.2
2.2 has 2 sig figs 13
= 12.716
Answer should have 2 sig figs
1.04 has 3 sig figs
1.04/3.665
3.665 has 4 sig figs 0.284
= 0.283765…
Answer should have 3 sig figs

62
 Addition or Subtraction
The result has the same number of places after the decimal as the least precise measurement.
Final answer reported
Example Rationale with correct number
of sig figs
1.3551 has 4 digits after the decimal
1.3551 + 136.99
136.99 has 2 digits after the decimal 138.35
= 138.3451
Answer should have 2 digits after the decimal
105 has 0 digits after the decimal
105 – 103.78
103.78 has 2 digits after the decimal 1
= 1.22
Answer should have 0 digits after the decimal

 Scientific Notation
When using scientific notation, convert the numbers to the same power of ten before any
calculations.

1.234 x 10-3 + 5.623 x 10-2 = 0.1234 x 10-2 + 5.623 x 10-2 = 5.746 x 10-2

 Logarithms
When taking a logarithm, the number of decimal places in the log is equal to the number of
significant figures in the original number.

log 563 = 2.751 105.234 = 1.71 x 105

 Multi-step Calculations
When performing a multi-step calculation, keep all the numbers in your calculation and
determine significant figures at the end. Avoid rounding intermediate values. Track the
significant figures through each mathematical operation, rounding your final value to the correct
number of significant figures.

Final answer
reported with
Example Rationale
correct number
of sig figs
110.530 has 3 digits after the decimal
110.7 has 1 digit after the decimal
Difference should have 1 digit after the
decimal (0.17 only has 1 sig fig that counts) 0.2%
Denominator has 4 sig figs
100 is an exact number
Final answer should have 1 sig fig

 Averages
As stated earlier, your answer cannot be MORE precise than the least precise measurement. This
particularly also holds true when calculating an average.

63
For example, when averaging measurements with 3 digits after the decimal point, the mean
should have a maximum of 3 digits after the decimal point as well. Consider the following data
set: 9.989 mL, 9.724 mL, 9.865 mL. Each value has 4 significant figures.

To calculate the average, we must first take the sum. These particular values yield a sum with 5
significant figures:

9.989 mL + 9.724 mL + 9.865 mL = 29.578 mL (5 sig figs)

The next operation involves dividing by the number of measurements. Note that 3 is an exact
number:

29.586 mL / 3 = 9.8593 (5 sig figs)

However, an exception to the rules for significant figures is made in this case. It does not make
sense for the average to be more precise than the measurements used to calculate it. The answer
should be reported as 9.859 mL. Since all the measurements have 3 digits after the decimal, the
average value should also be reported with 3 digits after the decimal.

B. Introduction to Basic Statistics for Chemistry Students


Consider a titration experiment consisting of 3 trials. The results obtained for the volume of
titrant required to reach the endpoint are: 11.65 mL, 11.58 mL, and 11.71 mL. Which of these
values is the “right” answer? If we make a limited number of measurements (called replicates),
some will be closer to the true value than others. This is because there are variations in the data
collected due to sources of error.

Sources of Error

In science, an “error” is anything that contributes to a measured value being different compared
to the true value. Here are some typical sources of error:

These always have the same magnitude and sign, resulting in bias (i.e., the
Systematic
measured values are consistently higher or lower compared to the true
errors
value)
A ruler missing the first 1 mm of its length: it will consistently give lengths
Example
that are 1 mm too short
These will have different magnitudes and signs, and result in a spread of
Random errors
the measured values from the true value
Any electronic measuring device: the random electrical noise within its
Example
electronic components will cause the reading to fluctuate over time
These are due to human oversight and other mistakes while reading and
Gross errors
recording data
Example Using the wrong bottle of solution to fill a buret for a titration

64
Mean

Due to the possibility of error, a single measurement of a quantity is not considered sufficient to
convey any meaningful information. When a measurement is repeated several times, we often
see the measured values grouped around some central value. This grouping can be described
with a single representative number called the mean (or average), which measures the central
value.

In order to determine the true value, we would have to obtain all possible measurement values;
i.e., make an infinite number of replicate measurements (n → ∞), and use these to calculate the
population mean. Since this is not feasible, the general approach is to perform a limited number
of replicate measurements (using the same instrument and method, under the same conditions).
This allows us to calculate the sample mean, which is an estimate of the true value. The larger
the number of replicates (the larger the value of n), the more this estimate approaches the true
value.

The sample mean is calculated as follows:

Accuracy and Precision

Accuracy: Accuracy is defined as the closeness of a result to the true value. This can be applied
to a single measurement, but is more commonly applied to the mean value of several repeated
measurements, or replicates. Accuracy can be expressed as a percent error, if the true value is
known.

Percent Error: the difference between a measured or experimental value E and a true or
accepted value A, divided by the true or accepted value A, as follows:

Precision: Precision is defined as the extent to which results agree with one another. It is a
measure of consistency. Precision can be calculated by determining the range of a set of
replicates.

Range: the difference between the largest value and the smallest value in a set of replicates; the
narrower the range, the more precise the results.

65
Alternatively, the precision of two measurements can be determined by calculating the percent
difference.

Percent Difference: the difference between the measured experimental values E1 and E2,
expressed as a fraction of the average of the two values.

Percent Change: the difference between the initial and final experimental values, divided by the
initial value.

66
0-1
Experiment 0: Introduction and Fundamentals of Analysis

This experiment should be performed and submitted individually.

Objective:
To identify the different instruments and glassware in the laboratory and gain a basic familiarity
with their use.

Prelab Assignment:
-Read the Fundamentals of Analysis section at the beginning of this lab manual.
-Read and understand the protocol for Experiment 0, below.
-Complete Prelab 0 on the last page of this experiment before you come to lab.

Introduction:

This experiment is an opportunity to develop hands-on experience with the glassware and
instruments that will be necessary for future experiments. You will learn to handle liquids and
solids, clean glassware and instruments properly and carry out the basic steps of analysis that
will be evaluated in future experiments. Throughout this exercise, you should refer to the
Fundamentals of Analysis section at the beginning of this lab manual for guidance and
instruction.

Methods:

Part A. Analytical balance

i. Select three 125 mL Erlenmeyer flasks and take them to the analytical balance. Make sure the
balance returns a reading of zero. If the reading is not zero, press the 0/T button. Make sure the
outside of the flasks are completely dry. Dry the outside of each flask with a paper towel or your
lab coat if they are wet. Label the flasks, 1, 2 and 3. Place the first flask on the balance platform
and take the mass. Record the mass in your report. Repeat for each flask.

ii. Return the Erlenmeyer flasks to your bench, and obtain a watch glass from your drawer. Place
the watch glass on the balance platform. Tare the watch glass by pressing the 0/T button. The
balance should return a mass of 0.000 g. Add 1.25 g NaCl (s) to the balance. You should use the
small spatula provided to scoop the solid carefully from the reagent bottle into the watch glass.
DO NOT transfer material from the watch glass back into the reagent bottle. Record the mass
you actually obtained in your report. Close the reagent bottle. Remove the watch glass from the
balance.

Now clean the balance. Press the 0/T button. Use the brush provided to gently clean around the
balance platform, and brush any spilled NaCl out of the balance onto the bench. Do not put
pressure on the balance platform itself.

Now clean the bench. Use a piece of paper towel to brush any NaCl on the bench the waste
container provided.

0-2
Part B. Beakers, flasks and graduated cylinders

i. Measure 50 mL of tap water using a 150 mL beaker.

ii. Identify the 10 mL, 25 mL and 100 mL graduated cylinders. Pour all of the 50 ml of tap water
measured with the beaker into a 100 mL graduated cylinder. Read the volume on the graduated
cylinder. You will have to estimate the uncertain digit. Record the volume indicated by the
graduated cylinder in your report.

iii. Use the 25 mL graduated cylinder to measure 25.0 mL of tap water. Bring the meniscus as
close to the 25.0 mL graduation as possible.

iv. Pour the 25.0 mL of tap water from the graduated cylinder into Erlenmeyer flask #1, weighed
in Part A. Do not discard the water in flask #1; this will be needed in Parts C and D.

Part C. Volumetric pipette

i. Pour about 100 mL of tap water into a 150 mL beaker.

ii. Select a 10 mL volumetric pipette. Note its single graduation above the bulb. Obtain a pipette
pump and gently insert the top of the pipette into the seal of the pipette pump with a slight twist.
For details see the Fundamentals of Analysis section at the beginning of this lab manual. The
pipette pump should not be excessively tight, but it should be tight enough to support the pipette
when held vertically. Hold the pipette pump in one hand and support the top of the pipette (above
the graduation) with the other.

iii. Insert the pipette tip into the tap water in your beaker, and aspirate about 5 mL of water by
rotating the wheel on the pipette pump downwards with your thumb.

iv. Clean and sheet the pipette with tap water as described in the Fundamentals of Analysis
section: remove the pipette pump; block the top of the pipette with your finger to prevent the
water from draining and carefully turn the pipette on its side; gently rotating the pipette to coat
the inner surfaces with the tap water; return the pipette to vertical and drain the tap water into a
waste beaker.

Rinsing the pipette with distilled water is not necessary in this case, because solutions will not
be measured or delivered at this time. At this time, the liquid being measured and delivered is
tap water, so in fact the pipette is already appropriately rinsed with the liquid to be delivered.

v. Aspirate tap water with the pipette to a level above the graduation. Remove the pipette pump
and block the top of the pipette with your finger. Do not apply excessive pressure; this could
break the pipette. However you will need to seal the opening sufficiently to prevent the liquid
from draining.

vi. Put down the pipette pump, and slowly unblock the top of the pipette with your finger. Fine
adjustments of the position of your finger will allow you to slowly let air into the top of the

0-3
pipette and gradually drain the liquid. Drain the liquid in the pipette until the meniscus reaches
the graduation. Block the end of the pipette to stop the liquid from draining once you are
satisfied that the meniscus is centered on the graduation.

vii. Deliver the 10.00 mL of water from the pipette to Flask #1. To deliver the liquid, unblock the
top of the pipette. Completely drain the pipette. Once all of the liquid has stopped, gently touch
the pipette to the wall of the Erlenmeyer in case a droplet is hanging from the pipette. Do not
attempt to blow out or shake out any more liquid from the pipette.

viii. Do not discard the water in Flask #1. We will use this water in Part D.

Part D. Mohr pipette

i. Pour 100 mL of tap water into 150 mL beaker.

ii. Select a 10.00 mL Mohr pipette. Note its range of graduations. Obtain a pipette pump and
gently insert the top of the pipette into the seal of the pipette pump with a slight twist. For details
see the Fundamentals of Analysis section at the beginning of this lab manual. The pipette pump
should not be excessively tight, but it should be tight enough to support the pipette when held
vertically. Hold the pipette pump in one hand and support the top of the pipette (above the
graduation) with the other.

iii. Insert the pipette tip into the tap water in your beaker, and aspirate about 5 mL of tap water
by rotating the wheel on the pipette pump with your thumb downwards.

iv. Clean and sheet the pipette with tap water as described in the Fundamentals of Analysis
section: remove the pipette pump and carefully turn the pipette on its side, gently rotating the
pipette to coat the inner surfaces with the tap water. Return the pipette to vertical and drain the
tap water into a waste beaker.

Rinsing the pipette with distilled water is not necessary in this case, because solutions will not
be measured or delivered at this time. The liquid being measured and delivered is tap water, so
in fact the pipette is already rinsed with the liquid to be delivered.

v. Aspirate tap water with the pipette to a level above the 5 mL graduation. Remove the pipette
pump and block the top of the pipette with your finger. Do not apply excessive pressure; this
could break the pipette. However you will need to seal the opening sufficiently to prevent the
liquid from draining.

vi. Put down the pipette pump. Keeping the top of the pipette blocked with your finger, take the
initial reading. You will need to raise the pipette to eye level to read the position of the meniscus
properly. See Reading the Meniscus in the Fundamentals of Analysis section at the beginning
of this manual for guidance.

vii. Your goal will be to deliver 5.00 mL of water to a Erlenmeyer flask #1. Aim the pipette tip
into the flask and slowly unblock the top of the pipette with your finger. Gentle adjustments will

0-4
allow you to slowly let air into the top of the pipette and gradually drain the liquid. Drain the
liquid in the pipette until the meniscus reaches the necessary final graduation. Block the end of
the pipette to stop the liquid from draining once you are satisfied that the meniscus is at the
necessary final graduation. Record the initial and final Mohr pipette readings in your report sheet.

viii. Erlenmeyer flask # 1 should now contain 25.0 mL of water added from a graduated cylinder
+ 10.00 mL of water added from a volumetric pipette + 5.00 mL of water added from the Mohr
pipette.

Take Erlenmeyer Flask #1 to the analytical balance. Dry the outside of the flask with paper towel
or your labcoat if necessary. Make sure the balance returns a reading of 0.000 g, and press the
0/T button if necessary. Place the flask carefully on the balance to determine the mass of the
flask and water. Record the mass in your report, and calculate the mass of water it contains.

Part E. Volumetric flask

i. Select a 50 mL volumetric flask. Note its single graduation. You will also need a 10 mL
volumetric pipette. Rinse each with tap water. Normally these glassware would be additionally
rinsed with distilled water, but in this experiment tap water is the solvent for each experiment, so
a rinse with tap water is sufficient.

ii. A solution of food colouring in tap water is provided. This is the concentrated stock solution.
Pour a small amount of this solution into a 30 mL beaker.

iii. Aspirate about 5 mL of the coloured solution and use it to sheet the pipette, as described in
Part C, and as explained in detail in the Fundamentals of Analysis section at the beginning of
the manual.

iv. Measure 10.00 mL of the coloured solution with the volumetric pipette as described in Part C
and deliver this to the volumetric flask.

v. Dilute the solution in the volumetric flask with tap water. Pour tap water from a beaker into
the flask to fill the flask about halfway. Swirl the liquid volumetric flask. Continue to add tap
water until the level of liquid is just below the neck of the flask. Gently swirl the liquid in the
volumetric flask again. Carefully fill the volumetric flask with tap water until the meniscus is
centered on the graduation. Be careful: the neck of the flask fills very rapidly, and it is
imperative that you do not exceed the volume indicated by the graduation. Add the last few
drops of water slowly until the meniscus of the solution is centered on the graduation of the
volumetric flask.

vi. Cap the volumetric flask. Invert the flask 5-10 times to homogenize the mixture. This is the
diluted working solution. Note the difference in colour compared to the stock solution.

vii. Discard the solutions down the sink.

0-5
Part F. Burette

i. Select a burette and set it up in the burette stand. Clean the burette with tap water as described
in the Fundamentals of Analysis section at the beginning of this manual. Normally the burette
would additionally be rinsed with distilled water and then the titrant. However, the titrant in this
experiment will be tap water, so no further rinsing is necessary in this case.

ii. Return the burette to the burette stand. Fill the burette with tap water; for details see the
Fundamentals of Analysis section. Place a liquid funnel in the top of the burette, and pour tap
water from a small beaker through the funnel into the burette. Fill the burette to a level near the
uppermost graduation. It is not necessary or recommended to fill the burette to 0.00 mL exactly.

iii. Place a waste container underneath the stopcock, and open the stopcock to the fully open
position to drain a few mL of tap water from the burette. You may notice some air bubbles being
pushed out with the liquid. Once the bubbles are cleared, close the stopcock. Close the stopcock
after a 1-2 seconds of draining. It should never be assumed that no air bubbles are trapped in the
stopcock. Air bubbles will introduce substantial error in your titration results. Always check for
and clear air bubbles as described here whenever you are filling a burette.

iv. Remove the funnel from the top of the burette and take the initial burette reading. See
Reading the Meniscus in the Fundamentals of Analysis section at the beginning of this manual
for guidance. Record the initial reading in your report sheet.

v. Place an empty Erlenmeyer flask below the burette. Open the stopcock to the vertical (fully
open) position for 10 seconds. The water will be delivered as rapidly as the burette is capable of,
in a stream. Close the stopcock. Take the final burette reading and record this in your report.

vi. The final reading from section v. above will be the initial reading for this measurement. For
this part of the experiment, slowly open the stopcock to achieve a drop wise flow of the tap water
into your flask. You need to achieve a rate of flow of about 2-3 drops per second. Once this rate
is achieved, allow the burette to deliver tap water for 10 seconds. Close the stopcock. Take the
final burette reading and record this in your report.

vii. The final reading from section vi. above will be the initial reading for this measurement.
This time, we want a very slow flow rate of about 1 drop per second or less. Deliver a total of 10
drops to the Erlenmeyer flask and close the burette. Take the final burette reading and record it in
your report.

Note: In future titrations you will be expected to be accurate to within 1 drop of liquid
delivered from the burette. Every drop beyond the 10 drops you were required to deliver
represents significant experimental error.

0-6
Part G. Cleaning glassware

Glassware must be cleaned after you use it, before you return it to its storage place. We also
suggest inspecting glassware carefully for cleanliness before you use it for experiments, and
cleaning the glassware again if necessary. Unclean glassware, or glassware contaminated with
detergent residue or other solution can all negatively impact your results.

DO NOT dry cleaned glassware with paper towel. Excess water should be poured out or drained
and then the wet glassware can be returned to its storage place.

i. Beakers, Erlenmeyer flasks and your watch glass should be cleaned in the large sinks at the
four corners of the lab. Use warm water and a small amount of detergent. Use a brush to clean
the interior surfaces. Do not force a brush that does not fit into the glassware; at best this will
scratch the glassware, at worst the glassware could break. Rinse the glassware with warm tap
water to remove the detergent. Return the glassware to your bench. Rinse the glassware with
distilled water two or three times.

ii. The burette should be cleaned at your bench by flushing it with tap water. Fill the burette with
tap water, and drain for 10 seconds through the stopcock into a waste beaker. Close the stopcock.
Then invert the burette to pour the remaining wash water down the sink out the top of the burette.
Repeat this three times.

Rinse the burette with 10 mL of distilled water. Pour 10 mL of distilled water into the top of the
burette. Remove the burette from the burette stand and carefully turn the burette on its side.
Slowly rotate the burette to sheet the inner surfaces with distilled water. Return the burette to
vertical, and drain the distilled water rinse. Repeat this three times.

iii. Volumetric pipettes and Mohr pipettes should be cleaned at your bench by flushing with tap
water. Aspirate 10 to 20 mL of tap water. Remove the pipette pump and turn the pipette on its
side, to sheet the interior surfaces. Drain the wash water into a waste beaker. Repeat this process
three times. Rinse the pipette with distilled water. Aspirate 5-10 mL of distilled water, and repeat
the rinsing process to sheet the interior surfaces. Drain the rinse water into a waste beaker.

iv. The volumetric flask should be cleaned by flushing it with tap water several times. DO NOT
attempt to use a brush to clean the volumetric flask. Rinse the volumetric flask with a small
quantity of distilled water. Tip the flask on its side to sheet the interior surfaces, before
discarding the rinse.

All liquid waste in this experiment may be discarded down the sink at your bench. The
NaCl(s) should be discarded into the waste container provided.

Wash all equipment with soap, brush and hot water at the sinks at the periphery of the labs.

0-7
Experiment 0 - Report Sheet Name ________________________

Part A. Analytical Balance Flask 1 Flask 2 Flask 3

Mass of Flask

Mass of NaCl

Part B. Beakers, Flasks and Graduated Cylinders

Volume of water measured with a 150 mL beaker

Actual volume of water measured above, indicated


by a 100 mL graduated cylinder
Volume of water measured with the 25 mL
graduated cylinder and delivered to Flask # 1

Part C. Volumetric Pipette


Volume of water delivered to Flask # 1 with the
volumetric pipette

Part D. Mohr Pipette

Initial Mohr Pipette Reading

Final Mohr Pipette Reading

Volume of water delivered to Flask # 1 by the Mohr


Pipette
Mass of Flask # 1 with water added in Parts B, C
and D

Mass of water in Flask # 1

0-8
Experimental Condition
Part F. Delivery of water with a Burette v. Fully open vi. 2-3 drops vii. 1 drop per
stopcock for 10 per second for second, 10
seconds 10 seconds drops total

Initial Burette Reading

Final Burette Reading

Volume of Water Delivered

Conclusions:

1. If the mass of a given Erlenmeyer flask is known, is it safe to assume that other flasks of the
same type and size have the same mass? Briefly explain.

2. What was the total volume of water delivered to Erlenmeyer Flask #1 in Parts B, C and D
according to the glassware used to deliver the water?

3. Using data from Part F, calculate the volume of 1 drop of solution delivered by the burette.

0-9
Experiment 0 - Prelab Assignment Name ______________________________

Answer true or false and correct any false statements.

1. You should never blow or shake out liquid remaining in the tip of a volumetric pipette.

2. After filling a burette with a reagent solution, open the stopcock to remove any air bubbles in
the tip.

3. Using a balance which measures masses to 0.001 g, a mass recorded as 2.01g is correct.

4. It is permissible to pour chemicals back into reagent bottles.

5. A funnel should be used to pour chemicals into a burette.

6. If something goes wrong with an experiment, it is permissible to use another student’s data
for one’s lab report.

7. Wet glassware should never be dried on the inside with paper towels.

8. In a titration experiment, 25.00 mL of 0.100 M HCl in a flask will be titrated with 0.100 M
NaOH from a burette. Before the NaOH solution is added to the burette, the burette should be
rinsed with a small amount of 0.100 M HCl.

9. When using a Mohr pipette, the liquid in the tip of the pipet should always be allowed to
drain completely when transferring the measured volume.

0-10
0-11
Experiment 1: Measurements of Mass and Volume

This experiment should be performed and submitted individually.

Objectives:
1. To empirically determine the density of tap water.
2. To quantitatively compare the accuracy and precision of measuring liquids using pipettes,
graduated cylinders and beakers.

Prelab Assignment:
-Read and understand the protocol for Experiment 1.
-Read and understand Section II.A and Sections III.A through III.H in Fundamentals of
Analysis at the beginning of the lab manual.
-Read and understand Section IV at the beginning of the lab manual.
-Complete the prelab assignment on the last page of this experiment.

Introduction:

Routine volumetric analysis requires accurate and precise handling and delivery of liquids.
Accuracy is the degree to which a measurement reflects the truth about a physical quantity.
Precision is the consistency or reproducibility of a measurement; the degree to which successive
replicate measurements of the same physical quantity agree with each other.

Achieving both sufficient accuracy and sufficient precision is of paramount importance in


analytical chemistry, and all wet-bench research that depends on these methods. Without
reasonably accurate results, the meaning of any analysis is called into question; what is the
purpose of carrying out an analysis if its results are so inaccurate as to have no relevance to the
true physical quantity being measured? Systematic error (sometimes called determinate error) is
a type of scientific error that contributes to poor accuracy. Systematic error is typically a
reproducible error that affects all replicate measurements more or less equally. For example,
incorrect calibration of an analytical balance would lead us to believe all of the masses
determined using that balance are slightly different than the actual mass. In this case, the error is
typically reproducible: repeated readings are consistently inaccurate, and the magnitude of the
error is always approximately the same. Systematic error can also be human error; if a piece of
measuring equipment is consistently used incorrectly by the operator, the measurement error that
results would be classified as systematic error.

Without sufficient precision, our ability to make conclusions about experimental data is severely
limited. The higher the precision for a set of replicate measurements, the more certain we are in
the validity of those measurements. High precision does not necessarily mean that the
measurements are accurate, but a series of highly precise results is at least evidence arguing in
favour of an accurate average. Uncertainty is caused by disagreement between replicate
measurements; i.e. repeating the same measurement gives different results for each trial. Some
uncertainty is inevitable, because random error is fundamental to all measurements; minute
differences in the conditions have small but quantifiable effects on each replicate, which lead the
replicates to be slightly different no matter how careful we are when taking the measurements.

1-1
For example, minute temperature variations that occur due to air currents in the room will
gradually change the temperature of a solution, causing it to expand or contract slightly, causing
random error in the measurement of the volume of the solution. Hence, random error is truly
random, and thus even perfectly executed replicate measurements are all slightly different from
each other. While some random error (and thus some uncertainty) is inevitable, a large
uncertainty often indicates other problems with the measurements. Human error is the greatest
source of irreproducible results and imprecision: lack of attention to detail and failure to carry
out experimental protocols correctly and consistently will result in unacceptably large
uncertainty.

Notably, it is possible to have relatively precise (highly reproducible) results that are nonetheless
inaccurate. This would reflect relatively reasonable random error, but some consistent source of
systematic error. Likewise, replicates that are very imprecise can nonetheless yield an average
that is reasonably accurate, but results with very high uncertainty are rarely considered
scientifically acceptable.

One of your goals as a scientist-in-training is to master the techniques required for making
measurements consistently and accurately. This experiment is designed to examine the relative
accuracy and precision of various glassware commonly used in volumetric analyses. It will also
help you develop valuable experience using these glassware that will be necessary as we move
forward into more sophisticated experiments.

Many experiments rely on accurate measurement and delivery of specific volumes of liquids: for
example in preparation of a solution of specific concentration, as we will see in an upcoming
experiment. In the current experiment, the accuracy and precision of pipettes, graduated
cylinders and beakers will be determined in Parts A, B, C and D, respectively.

In each case, these glassware will be used to measure and deliver specific volumes of water to
Erlenmeyer flasks. Tap water will be used throughout this experiment. The graduations on the
glassware will give the ‘indicated volume’ of water delivered in each case.

The mass of the water delivered to each Erlenmeyer flask will be measured by difference. Thus,
the initial (empty) mass of the Erlenmeyer must be known. Then, once water has been added to
the Erlenmeyer, the final mass is measured, and recorded. The mass of water delivered to the
flask is taken to be:

mass H2O = Mass of flask with H2O – Mass of empty flask

Importantly, the outside of each Erlenmeyer flask should be dried with paper towel before it is
placed inside the balance.

Once the mass of the water delivered is known, it can be used to calculate the volume of water
based on the density of water at that temperature according to the following:

mass material
d material 
volume material

1-2
i.e. the volume v of an object can be determined from m, the mass of that object and d is the
density of that object.

The density of the water should not necessarily be taken to be 1.000 g/mL. Importantly, density
is a temperature-dependent property, and it will be important to eliminate this source of error.
You will be provided with density charts that will give you density of water at different
temperatures; the temperature of the water will be measured and the appropriate density values
will be applied to all calculations.

We will refer to the volume determined from the mass and density in this way as the ‘actual
volume’ because, for the most part, students at this level find this method considerably more
accurate and precise than reading the indicated volume. In the long-term, your goal will be to
learn to measure the indicated volume as accurately as the actual volume. For the time being,
determining the ‘actual volume’ serves as a control; in this experiment it will provide a point of
comparison for all your measurements of the indicated volume.

In Part E of this experiment, the accuracy of your burette readings will be assessed. Unlike
pipettes, graduated cylinders and beakers, burettes are not typically used to measure and deliver
specific volumes of liquids. On the other hand, a burette is typically used in a titration to
determine the volume of a liquid required to reach the end-point of a titration. The accuracy of
such an experiment is dependent many factors, importantly including the accuracy of the burette
readings. That is, the volume actually delivered should be the same (within tolerances) as the
volume that is reported based on readings of the burette’s graduations.

As an exercise, we will assess the accuracy of your burette readings. This will give you an
opportunity to practice good techniques with the burette before doing more demanding
experiments with this instrument in upcoming labs. Water will be delivered from the burette into
an Erlenmeyer flask. As before, the mass of water delivered will be measured by difference. The
mass of water delivered will be used to determine the actual volume, based on the density of
water at this temperature. And the indicated volume will be compared to the actual volume.

For purposes of assessing the accuracy of a given piece of glassware, the % error between the
indicated volume to the actual volume will be used. In these cases, the % error will be calculated
as follows:
Indicated Volume - Actual Volume
% Error   100%
Indicated Volume

For simplicity, we will limit our calculations of the % error to assessments of the volumetric
pipette and beaker.

For the purposes of assessing the precision of a given piece of glassware, the range of actual
volumes delivered will be used. The range will be determined as follows:

Range = Largest Actual Volume – Smallest Actual Volume

1-3
Methods:

Part A. Determination of the accuracy and precision of a volumetric pipette

i. Obtain three empty Erlenmeyer flasks, and label them 1, 2 and 3. The flasks do not need to be
dry on the inside. Take the initial masses of each of the three flasks and record these values on
your report sheet.

ii. Obtain a 10 mL volumetric pipette. Clean the pipette with some tap water. Since tap water will
be delivered with the pipette, no rinsing with distilled water is necessary.

iii. Pipet 10.00 mL of tap water into Flask 1. Repeat this for Flasks 2 and 3.

iv. Make sure the outside of each flask is dried with a paper towel. Then take the final mass of
each of the three flasks, and record these values. You do not need to discard the water in these
flasks; you may save the water, and use the final mass from Part A as the initial mass for Part B.

v. Measure the temperature of the water delivered to one of the three flasks with the digital
thermometer provided. Record this temperature in your lab report in the space provided. You
may safely assume that this temperature is the same for all 5 parts of the experiment. You will
look up the density of water at this temperature on the density chart provided.

Part B. Determination of the accuracy and precision of a Mohr pipette

i. Use the same three flasks from Part A. Use the final masses for each flask from Part A as the
initial masses for Part B.

ii. Obtain a 10 mL Mohr pipette. Clean the pipette with some tap water. Since tap water will be
delivered with the pipette, no rinsing with distilled water is necessary.

iii. Use the Mohr pipette to deliver 5.00 mL of tap water to Flask 1: fill the pipette to a level
above the uppermost graduation; remove the pipette pump and use your finger to block the top of
the pipette; drain the pipette into a waste container until the meniscus is on the uppermost
graduation (0.00 mL); deliver the water to Flask 1 until the necessary final reading (5.00 mL) is
reached. Take and record the final pipette reading. The goal here is to consistently deliver as
close to 5.00 mL as possible. Make sure to record the initial and final volumes to determine the
volume of water actually delivered.

iv. Repeat step iii for Flasks 2 and 3.

v. Return to the same balance with your flasks. Dry the outside of each flask with a paper towel,
and then take the final mass of each flask, recording these data in your report.

Discard the water from all three flasks down the sink.

1-4
Part C. Determination of the accuracy and precision of a beaker

i. Obtain three empty Erlenmeyer flasks, and label them 1, 2 and 3. The flasks do not need to be
dry on the inside. Take the initial masses of each of the three flasks and record these values on
your report sheet.

ii. At your bench, measure 25 mL of tap water with your 150 mL beaker, and deliver this water
to Flask 1. Repeat this for Flasks 2 and 3.

iii. Make sure the outside of each flask is dried with a paper towel. Then take the final mass of
each of the three flasks, and record these values. You do not need to discard the water in these
flasks. You may save the water, and use the final mass from Part C as the initial mass for Part
D.

Part D. Determination of the accuracy and precision of a graduated cylinder

i. Use the same three flasks from Part C. Use the final masses for each flask from Part C as the
initial masses for Part D.

ii. At your bench, measure 25.0 mL of tap water with a 25 mL graduated cylinder and deliver
this water to Flask 1. Your goal is to approach 25.0 mL as closely as possible.

iii. Repeat step II for Flasks 2 and 3.

iv. Make sure the outside of each flask is dried with a paper towel. Then take the final mass of
each of the three flasks, and record these values. When you are finished, you may discard the
water in all three flasks down the sink.

Part E. Determination of the accuracy and precision of a burette

i. Discard any water remaining in your Erlenmeyer flasks. The inside of the flasks do not need to
be dry. Dry the outside of the flasks with paper towel, and take the mass of each flask.

ii. Fill your burette with tap water, using a funnel. Make sure to drain a few mL of water through
the stopcock into a waste container to clear any trapped air bubbles. Remove the funnel from the
burette before you take any readings. Make sure you have about 45 mL of water in the burette.

iii. Take the initial reading on your burette and record this on your report sheet. Then deliver
about 12.00 mL of water to Flask 1. You do not need to deliver 12.00 mL exactly, but make sure
the delivered volume is greater than 10.00 mL. Take the final burette reading and record it.

iv. Repeat step iii for Flasks 2 and 3.

v. Make sure the outside of the flasks are dried with a paper towel. Then take the final mass of
each of the flask and record it.

1-5
v. Discard the water in the flasks and the burette down the sink.

All liquid waste for this experiment can be discarded down the sink at your bench.

Wash all equipment with soap, brush and hot water at the sinks at the periphery of the labs.

1-6
Experiment 1 - Report Sheet Name ________________________

Temperature of water

Density of water at this temperature

Part A. Volumetric Pipette Trial 1 Trial 2 Trial 3

Initial mass of flask

Indicated volume of water delivered by


volumetric pipette

Final mass of flask

Mass of water delivered by the volumetric


pipette
Actual volume of water delivered
(based on mass and density)

Average actual volume

Volume range

% Error between indicated volume and average


actual volume

Part B. Mohr Pipette Trial 1 Trial 2 Trial 3

Initial mass of flask

Initial Mohr pipette reading

Final Mohr pipette reading

Indicated volume of water delivered by Mohr


pipette

Final mass of flask

Mass of water delivered by Mohr pipette

Actual volume of water delivered


(based on mass and density)

Average actual volume

Volume range

1-7
Part C. Beaker Trial 1 Trial 2 Trial 3

Initial mass of flask

Indicated volume of water delivered by beaker

Final mass of flask

Mass of water delivered by beaker

Actual volume of water delivered (based on


mass and density)

Average actual volume

Volume range

% Error between indicated volume and average


actual volume

Part D. Graduated Cylinder Trial 1 Trial 2 Trial 3

Initial mass of flask

Indicated volume of water delivered by


graduated cylinder

Final mass of flask

Mass of water delivered by graduated cylinder

Actual volume of water delivered (based on


mass and density)

Average actual volume

Volume range

1-8
Part E. Burette Trial 1 Trial 2 Trial 3

Initial mass of flask

Initial burette volume

Final burette volume

Indicated volume of water delivered by burette

Final mass of flask

Mass of water delivered by the burette

Actual volume of water delivered


(based on mass and density)
% Error between indicated volume and actual
volume*

Average % Error*

*Note: Your % errors for the burette readings are not being assessed for accuracy (we will be
assessing their ability to perform the correct calculation) at this time. This exercise will serve as
a formative assessment demonstrating your current skill level with this instrument. As a guide,
5% error is tolerable, but error < 2 % can be achieved by following best practices outlined in
the Fundamentals of Analysis section of the manual and will be required for the best results in
analytical experiments.

Conclusions

1. Based on your understanding of the purpose of each piece of glassware, which should be more
accurate, a volumetric pipette or a beaker?

2. Based on your data, between the volumetric pipette and the beaker, which was most accurate?
Which was most precise?

1-9
1-10
Experiment 1 - Prelab Assignment Name ______________________________

1. A burette was used to determine the density of tap water. In one trial, the initial burette
reading was 1.37 mL and then some water was delivered to a beaker. The final burette reading
was 5.41 mL. The mass of the tap water delivered was determined to be 4.0347 g. Based on these
data, what was the density of the tap water?

2. A 250 mL graduated cylinder was used to measure and deliver 250.0 mL of tap water. The
mass of tap water delivered was determined to be 246.6143 g.

a. Use the density of tap water calculated in question 1 above to determine the volume of tap
water actually delivered by the graduated cylinder.

b. What was the % error between the volume indicated by the graduated cylinder and the volume
actually delivered based on the mass and density?

3. Provide a brief definition for the term empirical.

4. Briefly explain why it is preferable to design an experiment with multiple trials instead of a
single trial.

1-11
1-12
Experiment 2: Solutions

This experiment should be performed and submitted individually.

Objectives:
1. To empirically determine the density of a solution.
2. To accurately prepare solutions of a desired concentration.
3. To accurately dilute concentrated stock solutions to a desired final working concentration.

Prelab Assignment:
-Read and understand the protocol for Experiment 2.
-Read and understand Sections II.B and III.F in Fundamentals of Analysis at the beginning of
the lab manual.
-Review Sections III.C, III.D, III.E, and III.G in Fundamentals of Analysis at the beginning of
the lab manual.
-Complete the prelab assignment on the last page of this experiment.

Introduction:

Solution chemistry is nearly ubiquitous in the applied research setting, and all of the
experiments/reactions we will study henceforth in this course will involve solutions in some
form. An understanding of solution preparation is a fundamental part of your formal education in
chemistry.

A solution is a homogeneous mixture of two or more substances. A solution is not a compound,


even though it is homogeneous, because its composition is variable; different ratios of the
substances in solutions may be combined homogenously to form solutions with different
concentrations.

Many solutions can be classified as consisting of a solvent and one or more solutes. The solvent
(usually a liquid) provides a medium into which the solute is dissolved. The solubility is the
maximum concentration of a given solute that will dissolve in a given solvent at a given
temperature. In a given solvent, a material that is relatively soluble has a high solubility at a
given temperature. The concentration of a solution is generally taken to be a measure of the
abundance of the solute in solution. The more concentrated a solution, the greater the abundance
of solute particles in a given amount of the solution.

The majority of the reactions studied at this level occur in aqueous solution (i.e. water is the
solvent). Water is a powerful solvent for polar solutes, including many ionic compounds (salts)
and acids; these materials are relatively soluble in water. However, non-polar materials tend to
be considerably less soluble in water. In practice, a wide variety of liquids may be used as
solvents to dissolve different solutes. For instance, many different organic solvents are available
and appropriate for preparing solutions of relatively non-polar solutes. Later in this course we
will use ethanol (C2H5OH) and dichloromethane (CH2Cl2) as solvents.

2-1
Notably, the solution density is often substantially different from the density of its components.
In general, it is not safe to assume the density of solution is exactly equal to the density of the
solvent. Any such assumption inherently involves some systematic error. Solution density is
defined as the mass of solution per unit volume of solution, typically expressed in g/mL as
follows:

mass of solution
d solution 
volume of solution

Note that the density of solution is not a direct measure of solution concentration.

A variety of solution preparation methods exist, but in general the goal in an analytical setting is
to prepare a solution of accurately known solute concentration. It is important that a solution’s
concentration is expressed clearly and unambiguously. There are many different ways of
expressing the concentration of a solute in solution, but the most commonplace conventions used
in the literature are percent by mass (mass %), mole fraction, molarity, molality and parts per
million (or per billion). What follows is a brief description of each convention:

The percent by mass is a convenient means of rapidly communicating the concentration of a


solute in solution. It is calculated as follows:

mass of solute
% by mass of a solute in solution   100%
mass of solution

The mole fraction has various applications that are applied later in this course. Although mole
fractions are not applied in depth in this experiment, for reference the mole fraction is defined as
follows:

moles of solute
Mole Fraction of a solute in solution 
moles of all material in solution

Sometimes the mole fraction is converted to a mole %, simply by multiplying the above by
100%.

Molarity (the molar concentration) is by far the most common convention for expressing
solution concentration. Molarity is defined as the number of moles of solute per liter of solution.
An accurate understanding of the units in this case is crucial: the volume of solution must be
known accurately, but not necessarily the volume of the solvent. That is:

moles of solute
Molar Concentration of a solute in solution  [solute ]  M solute 
L of solution

Molality (the molal concentration) is an important alternative measure of solution concentration


that will be explored later in the course. For now, we can define the molality as the moles of

2-2
solute per kg of solvent. This is a small but crucial difference between the molarity and molality
of a solution. That is:

moles of solute
Molal Concentration of a solute in solution 
kg of solvent

Lastly parts per million (ppm) is a very common unit of solution concentration encountered in
several fields of chemistry. We are not likely to encounter ppm determinations in this course.
However, for the time being we can define the ppm concentration as follows:

g of solute g of solute
ppm of a solute in solution  
Mg of solution 106 g of solution

That is, the ppm concentration is the mass of solute (in grams) per million grams of solution.

To some extent, the way the concentration of a solute in solution is expressed dictates the
method of preparing that solution, and vice versa. That is, the methods by which a solution is
prepared will often dictate which expressions of solution concentration are possible for a given
solution.

For example, to prepare a solution of a known molar concentration, the volume of solution must
be accurately known, and thus specific glassware should be used to accurately determine the
final volume of the solution. A volumetric flask is the most accurate glassware available for
preparing a solution of known molar concentration. Notably, accurate knowledge of the volume
of solvent is not required to determine the molar concentration of the solution. However, without
an accurate knowledge of the quantity of solvent in a solution, determining the molal
concentration of that solution is not possible.

It is an inherent limitation of the volumetric flask that it does not permit convenient
determination of the volume of the solvent used to prepare a solution. Importantly, the volume of
the solution does not necessarily equal the volume of the solvent. In practice, we find that when
dissolving a solute in a liquid, or when combining different liquids, the solution tends to expand
or contract somewhat, depending on the interactions between the components of the mixture.
Thus the volume of the solution does not necessarily equal the sum of the volumes of its
components, and the volume of a solution cannot be assumed to be equal to the volume of its
solvent.

In this experiment, we will prepare a series of aqueous solutions and explore molarity and % by
mass in some depth. You will be given an index card with three solutions that need to be
prepared. On the report sheet you will need to show the calculations required for the preparation
of each solution. You will prepare each solution as instructed, and then validate the preparation
by two different means: either the density of the solution will be assessed, or the absorbance of
the solution will be measured. Three different scenarios are examined in this experiment:

A. Preparation of a solution from the pure solid solute to achieve a desired % by mass.

2-3
B. Preparation of a solution from the pure solid solute to achieve a desired molar concentration.

C. Dilution of a concentrated stock solution of known molar concentration to a desired working


concentration.

Scenarios A and B are commonly encountered in research labs. In these scenarios, the solute is
of high purity. An analytical balance is used to weigh the necessary amount of the solid material
to produce a solution of a desired final volume and solute concentration. The solute is dissolved
in the solvent. In scenario A, the mass of the solvent must be known. In scenario B, the mass of
the solvent is not critical, but the volume of the solution must be accurately known, and thus the
solution will be transferred to a volumetric flask before being diluted to the final volume.

Scenario C illustrates a typical situation in which a concentrated stock solution must be diluted
to a final ‘working’ concentration - i.e. the concentration at which the solution is used for actual
experiments. To achieve the desired final concentration and final volume of the working solution,
the volume of the concentrated stock solution that is required can be determined according to the
formula:

M V M V
1 1 2 2

Where “1” indicates a stock solution and “2” indicates a working solution.
Thus M1 is the initial (stock) solution concentration and V1 is the volume of that stock
solution used to prepare a working solution. M2 is the final (dilute) working solution
concentration and V2 is the final volume of the working solution.

Thus, for scenario C, it will be necessary to accurately transfer the required volume of the
concentrated stock solution into a volumetric flask, and then dilute the solution to the final
volume with solvent in order to achieve the desired final concentration.

In all three scenarios the goal is to prepare the desired solution as accurately as possible.

Methods:

Part A. Preparation of a NaCl solution from solid material

i. Solid NaCl is provided. The index card provided indicates a specific target concentration in
% by mass for a NaCl solution that must be prepared. Record the index card # and the % by mass
on your lab report in the space provided. Calculate the mass of NaCl required to prepare 100.0 g
of this solution. Calculate the mass of solvent (water) that will be required to prepare 100.0 g of
this solution. Record these calculated masses in your lab report.

ii. Assume that the density of the tap water is 1.00 g/mL. Calculate the volume of tap water
needed to deliver the mass of water calculated above. Record this volume in your report.

2-4
iii. Obtain a clean 150 mL beaker. It does not necessarily need to be completely dry on the inside,
but the outside should be dry.

iv. Place the 150 mL beaker on the analytical balance platform, take its mass and record it on
your report sheet in the space provided. Press the 0/T button to tare the beaker. The balance
should return a reading of 0.000 g. Weigh the required amount of NaCl (s) into the tared beaker.
The mass of NaCl you obtain should be as close to the required mass as reasonably possible.
Return to your bench.

v. Use a 100 mL graduated cylinder to measure the calculated volume of tap water required, as
determined in section ii. above. The volume of water should be as close to the required volume
as reasonably possible.

vi. Pour all of the tap water from the graduated cylinder into the beaker containing the NaCl (s).
Swirl the solution gently until all of the NaCl is dissolved. Take the mass of the 150 mL beaker
with the solution and record it on your report sheet in the space provided.

vii. Transfer the NaCl solution back into the 100 mL graduated cylinder to measure its final
volume. Record the final volume.

viii. Calculate the density of the NaCl solution. Compare this to the density of the solvent (tap
water).

ix. Calculate the molar concentration of the NaCl solution.

The NaCl (aq) solution can be discarded down the sink at your bench. Any solid waste NaCl
should be discarded in the solid waste container in the fume hood.

Part B. Preparation of a CuSO4 solution from solid material.

i. Solid CuSO4·5H2O is provided. This material is a hydrated ionic solid; the solid contains 5
moles of water per mole of the ionic compound CuSO4. When preparing an aqueous solution of a
hydrated ionic compound, like CuSO4·5H2O, water molecules from the solid crystal lattice enter
the liquid phase and become part of the solution.

ii. The index card provided indicates a concentration of CuSO4 (aq) solution that must be prepared
from CuSO4·5H2O (s) and distilled water. Record this concentration in your report.

iii. Weigh the required amount of CuSO4·5H2O (s) into a tared 150 mL beaker using the
analytical balance. The mass should be as close to the required mass as reasonably possible.
Record the mass of CuSO4·5H2O (s) that you actually obtained.

iv. At your bench, dissolve the CuSO4·5H2O (s) in about 60 mL of distilled water with gentle
swirling or stirring. Once the solid is completely dissolved, transfer the solution into a clean 100

2-5
mL volumetric flask. Rinse the beaker with approximately 5 mL of distilled water, and transfer
this volume to the volumetric flask. Repeat the rinsing process 2 times, taking care not to exceed
the total volume of the volumetric flask.

v. Dilute the solution in the volumetric flask to the final volume. Cap the volumetric flask, and
mix by gently inverting the flask 5-10 times.

vi. Acquire two plastic cuvettes from the tray in the center of the lab bench. Handle the cuvettes
according to the guidelines in the Fundamentals of Analysis section at the beginning of this
manual. Rinse two cuvettes with distilled water several times. Fill one of the cuvettes with
distilled water. This cuvette will serve as the reference (blank).

Rinse the other cuvette with your CuSO4 solution: first fill the cuvette with your CuSO4 solution,
and discard this rinse. Refill the cuvette with your CuSO4 solution.

vii. The spectrophotometer will be used to determine the absorbance of your CuSO4 solution at
605 nm. First make sure the spectrophotometer is set at the required wavelength of 605 nm.

Handling the reference cuvette by the side surfaces (the front surface is marked with an arrow),
wipe the front and back surfaces of the cuvette with a Kim wipe. Place the cuvette in the
measurement port of the spectrophotometer so that the arrow on the cuvette is aligned with the
arrow on the spectrophotometer indicating the light path. Close the lid of the measurement port.
Press the ‘Ref’ button. The spectrophotometer should now return an absorbance reading of 0.000.

viii. Remove the reference cuvette from the measurement port of the spectrophotometer. Wipe
the cuvette containing your CuSO4 solution as described above and place it in the measurement
port. Close the lid of the measurement port. The spectrophotometer should now return the
absorbance reading of your CuSO4 solution. Record this value in your report sheet.

The CuSO4 (aq) solution should be discarded in the waste container in the fume hood. Any
solid CuSO4·5H2O should be discarded in the solid waste container in the fume hood.

Part C. Preparation of a dilute working solution from a concentrated stock solution

i. A concentrated stock solution of yellow dye of known molar concentration is provided. The
concentrated stock solution will be used to prepare a dilute ‘working’ solution of a desired final
concentration using volumetric glassware. The index card provided indicates the desired final
concentration of the dilute ‘working’ solution you should prepare. Record these details on your
report sheet in the space provided.

ii. Calculate the volume of the concentrated stock solution that will be necessary to prepare 50.00
mL of the working solution. Rinse a 30 mL transfer beaker in a small volume (~ 1 mL) of the
concentrated stock solution, and discard this rinse. Pour the concentrated stock solution into the
beaker; add up to 10 mL more than the calculated volume you need.

2-6
iii. Select an appropriate pipette to deliver the calculated volume. Clean the pipette. Rinse the
pipette in a small volume of the concentrated stock solution. Use the pipette to transfer the
required volume to the appropriate volumetric flask. The volume you deliver should be as close
to the required volume as reasonably possible.

iv. Dilute the solution in the volumetric flask to the graduation with distilled water. Cap the
volumetric flask, and then mix the solution by inverting it 5-10 times.

v. Acquire two plastic cuvettes from the tray in the center of the lab bench. Handle the cuvettes
according to the guidelines in the Fundamentals of Analysis section at the beginning of this
manual. Rinse two cuvettes with distilled water several times. Fill one of the cuvettes with
distilled water. This cuvette will serve as the reference (blank).

Rinse the other cuvette with your yellow dye solution: first fill the cuvette with your yellow dye
solution, and discard this rinse. Refill the cuvette with your yellow dye solution.

vi. The spectrophotometer will be used to determine the absorbance of your yellow dye solution
at 400 nm. First make sure the spectrophotometer is set at the required wavelength of 400 nm.
Handling the reference cuvette by the side surfaces (the front surface is marked with an arrow),
wipe the front and back surfaces of the cuvette with a Kimwipe. Place the cuvette in the
measurement port of the spectrophotometer so that the arrow on the cuvette is aligned with the
arrow on the spectrophotometer indicating the light path. Close the lid of the measurement port.
Press the ‘Ref’ button. The spectrophotometer should now return an absorbance reading of 0.000.

vii. Remove the reference cuvette from the measurement port of the spectrophotometer. Wipe the
cuvette containing your yellow dye solution as described above, and place it in the measurement
port. Close the lid of the measurement port. The spectrophotometer should now return the
absorbance reading of your yellow dye solution. Record this value in your report sheet.

The yellow dye solution should be discarded in the waste container in the fume hood.

Wash all equipment with soap, brush and hot water at the sinks at the periphery of the labs.

2-7
2-8
Experiment 2 - Report Sheet Name _____________________

Part A. Preparation of a NaCl solution

Index Card #

% by mass of NaCl in solution A (index card)

Calculated mass of NaCl required to prepare solution A

Calculated mass of tap water required to prepare solution A

Calculated volume of tap water required to prepare solution A

Mass of empty beaker

Mass of NaCl obtained

Mass of beaker containing NaCl solution

Mass of NaCl solution

Measured volume of the NaCl solution

Density of the NaCl solution

Molar concentration of the NaCl solution

Calculations:

2-9
Experiment 2 - Report Sheet Name ______________________________

Part B: Preparation of CuSO4 solution from solid material

Index card # Spectrophotometer #

Required concentration of CuSO4 (aq) (from index card)

Calculated mass of CuSO4·5H2O (s)

Desired volume of solution B 100.00 mL

Mass of CuSO4·5H2O (s) actually obtained

Theoretical molar concentration of CuSO4 (aq) in solution based on


the mass actually obtained

Absorbance of solution at 605 nm

Empirical molar concentration of solution B based on standard


curve provided

Calculations:

2-10
Experiment 2 - Report Sheet Name ______________________________

Part C - Targeted Dilution of a Yellow Dye Stock Solution

Index card # Spectrophotometer #

Molar concentration of yellow dye stock solution provided

Desired concentration of dilute solution C (from index card)

Desired volume of dilute solution C 50.00 mL

Calculated volume of stock solution required to prepare the dilute


solution C

Absorbance of solution C

Empirical molar concentration of solution C based on standard


curve (provided)

Calculations:

2-11
2-12
Experiment 2 - Prelab Assignment Name ______________________________

1. Concentrated HCl has a density of 1.19 g/mL and contains 37.0% by mass HCl. Calculate the
molar concentration of this solution.

2. A total of 18 g of NaOH is dissolved in water to produce a solution which contains 18 %


NaOH by mass. Calculate the mass of water used to prepare this solution.

3. How many grams of KC1 are required to make 250.0 mL of a 0.250 M solution?

4. What volume of 8.2 M NaOH is required to prepare 1.00 L of 1.20 M NaOH solution?

5. Glacial acetic acid, has a density of 1.05 g/mL and contains 99.51% acetic acid (CH3COOH)
by mass. Calculate the molarity of glacial acetic acid.

6. A total of 10.0 g of compound X was dissolved in water, producing 450.0 mL of solution with
a concentration of 0.55 M. What is the molar mass of compound X?

2-13
3-1
Experiment 3: Titrations I – Standardization of NaOH and Analysis of an
HCl Solution

This experiment should be performed and submitted individually.

Objectives:
1. To determine the concentration of a solution of NaOH by titration with KHP.
2. To determine the concentration of a solution of hydrochloric acid by titration analysis.

Prelab Assignment:
-Read and understand the protocol for Experiment 3, including the calculations required.
-Read and understand sections III.G and III.H in the Fundamentals of Analysis section at the
beginning of this manual.
-Review sections III.D and III.F in the Fundamentals of Analysis section at the beginning of
this manual.
-Complete the prelab assignment on the last page of this experiment.

Introduction:

Titrations are an important class of experiments in analytical chemistry. There are many different
types of titrations with different analytical goals, and thus different experimental conditions.
However, all titrations have a fundamental definition in common: in a titration, reagents that
participate in a chemical reaction are quantitatively combined in their stoichiometric ratio so that
neither reagent is in excess at the end-point.

Determination of the stoichiometric ratio of the reagents during a titration allows for a number of
analytical conclusions to be drawn. For example, if the reaction stoichiometry is known
beforehand, a knowledge of the concentration of one of the reagents can be used to empirically
determine the concentration of the other by titration analysis. This is the underlying principal
behind the vast majority of routine quantitative chemical analyses. Alternatively, if the reaction
stoichiometry is not precisely known, a knowledge of the concentrations of the reagents will
allow the reaction stoichiometry to be empirically determined.

In general, the reagents involved in a titration are typically referred to as the analyte (the sample
being analyzed) and the titrant (the material being added to the analyte).

Importantly, during a titration the operator will require some means of empirically determining
how much of the titrant to add to the analyte to achieve the stoichiometric ratio of reagents in the
reaction mixture. This is necessary for both of the experimental scenarios described above.
Historically, reactions involving a colour change have been the most easily studied by titrations,
because the colour change will indicate the progress of the reaction. For example, imagine a
hypothetical reaction:

A (aq) + B (aq) C (aq)


Colourless Brown Colourless

3-2
In this reaction A (aq) and C (aq) are colourless liquids, but B (aq) is a dark brown liquid. When A
and B are combined, the dark coloured solution of B is converted to a colourless solution of C,
provided some A exists in solution to react with it. Small amounts of B (aq) can be continually
added to the reaction mixture, and as long as A is in excess, the brown colour will disappear to
become colourless C. However, if small amounts of B are continually added to the reaction
mixture, at some point the brown colour will not dissipate - the brown colour will persist because
there is no longer any A in solution to react with B. All of the A has been consumed. At this
exact point (the 1st persistent appearance of brown colour in the reaction mixture) the titration is
at its ‘end-point’; that is, the ratio of the two reagents combined in the reaction mixture is as
close to the stoichiometric ratio as can be achieved.

Several different methods of quantitatively combining the analyte with the titrant are available.
The most common method is to use volumetric glassware to measure the analyte and to use a
burette to deliver the titrant. These glassware provide sufficient precision to allow for
quantitative determination of the amounts of each reagent that are combined during the titration.

A common class of titrations is the acid-base titration. In an acid-base titration, the acid and base
are combined quantitatively in their stoichiometric ratio: the result is complete neutralization of
the acid by the base and vice versa. Importantly, since ideally the acid and base are combined in
exactly their stoichiometric ratio during the titration, neither reagent is in excess at the end-point,
so there is exactly enough base to completely neutralize the acid, but not more. In general, acids
and bases can be considered to react as follows:

HA (aq) + B (aq) A- (aq) + BH+ (aq)

Commonly in this course, and throughout analytical chemistry, the titrant in an acid-base
titration is a solution of NaOH (aq) or KOH (aq), that serves as a source of the base, hydroxide ions
(-OH). For example:

HA (aq) + NaOH (aq) A- (aq) + Na+ (aq) + HOH (l)

In this case, the Na+ ion is actually a spectator, so we could write the net ionic equation as
follows:

HA (aq) + -OH (aq) A- (aq) + HOH (l)

For example, if nitric acid, HNO3 (aq) were titrated with NaOH (aq), the reaction could be written:

HNO3 (aq) + NaOH (aq) NO3- (aq) + Na+ (aq) + HOH (l)

Or if you prefer:

HNO3 (aq) + -OH (aq) NO3- (aq) + HOH (l)

Or, H+ (aq) + -OH


(aq) HOH (l)

3-3
Notably most acids and bases are clear and colourless solutions. In order to visually observe the
progress of the reaction, typically a pH-colour indicator is added to the reaction mixture. The
chemical structure of a pH indicator is pH sensitive: within a particular range of pH values the
indicator will have a given colour, and beyond these pH values the colour of the indicator will be
dramatically different.

By far the most commonly used such indicator is phenolphthalein. Phenolphthalein is colourless
from pH 0 - 8.8, and is visibly pink or fuchsia-coloured at pH 8.8-13. So in an acidic solution,
such as the analyte in an acid-base titration, the phenolphthalein will be colourless. As the titrant
(the base in such a titration) is added, the pH of the reaction mixture increases. At the end-point
of the titration, where the ratio of the analyte and titrant is very near to the stoichiometric ratio of
these reagents, the pH is typically close to neutral (certainly closer to pH neutrality than the pure
reagents). Even a small amount of titrant added beyond this point dramatically increases the pH
beyond neutrality, and as this occurs the phenolthalein in the reaction mixture turns pink. Thus
the appearance of pink colour in an acid-base titration using phenolphthalein as an indicator
allows the end-point of the titration to be visually-determined.

O
O

O
OH
-
pH = 8.8 O

OH

HO O

phenolphthalein colourless phenolphthalein pink


pH 0 to 8.8 pH 8.8 to 11

Titrations require the concentration of at least one of the reagents to be quantitatively known. In
acid base titrations, the titrant NaOH is typically standardized; its concentration is quantitatively
determined before this solution can be used in an analysis of an unknown. This is because NaOH
solutions are difficult to prepare with high precision from solid material. NaOH (s) absorbs water
from the ambient air, especially in conditions of moderate to high humidity, making quantitative
solution preparation inaccurate and inconsistent. Typically it is only possible to prepare a NaOH
solution to an approximate final concentration using volumetric glassware, as we did in
Experiment 2. The solution is then standardized by titration with a an acid of high purity.
Typically the acid used for this purpose is potassium hydrogen phthalate, C8H5O4K, often
abbreviated as “KHP”. KHP is a weak acid, but it reacts with NaOH according to the same
general formulae we explored above. That is:

3-4
-
O OH O O
O O

- +
O
-
K+ + NaOH O K + Na+ + HOH

Potassium Hydrogen Phthalate


(KHP)
-
HOOCC6H4COOK + NaOH OOCC6H4COOK + Na+ + HOH

Thus in a titration where the concentration of the titrant (NaOH solution) is only approximately
known, we can quantitatively determine the concentration of NaOH by titrating with an analyte
that contains a known amount of KHP with phenolphthalein as an indicator. Thus the NaOH
solution can be standardized.

In this experiment you will gain some experience with titrations by carrying out a series of acid-
base titrations. We will standardize a NaOH (aq) solution using the KHP. We will also titrate a
solution of hydrochloric acid with unknown concentration to quantitatively determine the
concentration of this acid.

Methods:

Part A. Standardization of a NaOH solution using KHP

i. You will need your clean 250 mL Erlenmeyer flask and three clean 125 mL Erlenmeyer flasks.
Label the large flask as #1 (this will be for your orientation trial) and the three 125 mL flasks as
#2, #3 and #4. The insides of the flasks do not necessarily need to be dry. If they are wet, make
sure to rinse them with distilled water. The outside of the flasks must be completely dry.

Place flask #1 on the analytical balance, and tare the balance. Add between 0.240 to 0.255 g of
KHP into the flask. Do not restart if you accidentally add too much KHP. Record the mass of
KHP actually obtained on your report sheet. Repeat these steps for each of the remaining three
flasks.

ii. Use a clean graduated cylinder to deliver about 25 mL of distilled water to each flask. Gently
swirl the flask until all of the KHP is completely dissolved to create an aqueous KHP solution.

iii. The NaOH solution to be used as the titrant for Part A will be provided. DO NOT use the
NaOH solution from Part B.

iv. Clean and rinse a burette. Pour a maximum of 60 mL of NaOH solution from Part A from
the stock bottle into a clean 100 mL transfer beaker. Use 5 mL of NaOH solution to rinse the
burette. Repeat this rinse. Then fill the burette with NaOH solution.

3-5
v. Make sure to remove the liquid funnel from the top of the burette. Drain a few mL of the
titrant through the stopcock into a waste container to clear any air bubbles. Record the initial
burette reading.

vi. Start with Flask #1. This flask will serve as an orientation trial, to rapidly determine the
approximate end-point volume of the titration. Add 2 to 3 drops of phenolphthalein solution to
the flask and gently swirl. Position the flask beneath the burette. Place a white piece of paper
beneath the flask. This will help you visualize the colour change of the indicator.

vii. Begin delivering the titrant to the Flask #1 by opening the stopcock to achieve a steady
stream of the titrant. As the titrant is being delivered, continuously swirl the flask in a circular
motion. As the titrant enters the reaction mixture, you will observe a pink colour appearing and
then rapidly disappearing as you swirl the mixture. This indicates that the titrant is neutralizing
the acid, and in slight excess locally where the drop of titrant is entering the reaction mixture:
upon mixing the titrant is then encountering more acid, and being neutralized.

As the titration proceeds, the pink colour may persist for a few seconds before disappearing
despite continuous swirling. This indicates that you are approaching the end-point. Continue
swirling and at some point the addition of the titrant will cause the reaction mixture to turn pink
permanently. This is the end-point of the titration. Immediately close the stopcock to stop the
flow of titrant.

viii. Take the final burette reading and record this in your lab report. The difference between the
final reading and the initial reading is the volume of titrant delivered and the end-point volume of
the orientation trial. Since the orientation trial was carried out rapidly, it likely carries with it a
large systematic error. However, it serves to rapidly determine the approximate end-point
volume.

ix. Repeat steps vi through viii for Trial 2 using Flask 2 with the following modification: titrate
rapidly until you approach the approximate end-point (1 to 2 mL less titrant than the orientation
trial) and then slow the delivery of titrant to about 1 to 2 drops per second. Swirl the flask
continuously. As you approach the end-point (pink colour persisting for longer periods) slow the
delivery of titrant to 1 drop per second or less. At this stage, be very careful to swirl between
every drop of titrant added and visually assess the colour of the mixture. The goal will be to
achieve a permanent faint pink colour without overshooting the end-point. The lighter pink the
mixture, the closer to the actual end-point, and the more accurate the titration. Stop the delivery
of titrant as soon as the mixture turns permanently faint pink. You may wish to stop delivery of
the titrant after each drop added as you swirl and assess the colour. You may also wish to deliver
‘split drops’ of titrant – a fraction of a drop at a time to achieve the most accurate results. Your
instructors will demonstrate this technique.

Once you have achieved a permanent pink end-point, stop the flow of titrant and record the final
burette reading. The end-point volume is the difference between the final and the initial volume.

x. Repeat step ix for Trials 3 and 4. Assuming Trial 2 was carried out accurately and without
issues, the end-point volume for Trials 3 and 4 should be very similar. If you are confident, you

3-6
can deliver this volume (minus approximately 0.5 mL) very rapidly, and then slow the titration
as described above.

xi. For your calculations, ignore Trial 1 (the orientation trial). It is not reliable and should not be
considered a valid data point. Use data from Trials 2, 3 and 4 to determine the concentration of
the titrant.

Liquid waste containing KHP should be disposed of in the waste container in the fume hood.
Solid KHP waste should be disposed of in the solid waste container in the fume hood.

Part B. Analysis of an HCl solution

i. You will need your clean 250 mL Erlenmeyer flask and three clean 125 mL Erlenmeyer flasks.
Label the large flask as #1 (this will be for your orientation trial) and the three 125 mL flasks as
#2, 3# and #4. The insides of the flasks do not necessarily need to be dry. If they are wet, make
sure to rinse them with distilled water.

ii. You will be provided with an analyte consisting of an HCl solution of unknown concentration.
Pour about 50 mL of this HCl solution into a clean 50 mL transfer beaker.

iii. Sheet a clean 10 mL pipette with a small amount of the acid solution. Pipet 10.00 mL of the
HCl stock solution into each of the four clean Erlenmeyer flasks. Add about 15 mL of distilled
water to each flask from a graduated cylinder and gently swirl the flask.

iv. The NaOH solution to be used as the titrant for Part B will be provided. It has been
standardized by your instructor, so the concentration of this solution is accurately known, and
will be given. Record this concentration on your report sheet in the space provided. DO NOT use
the NaOH solution from Part A.

v. Clean and rinse a burette. Pour a maximum of 60 mL of NaOH solution from Part B from
the stock bottle into a clean 100 mL transfer beaker. Use 5 mL of NaOH solution to rinse the
burette. Repeat this rinse. Then fill the burette with NaOH solution.

vi. Make sure to remove the liquid funnel from the top of the burette. Drain a few mL of the
titrant through the stopcock into a waste container to clear any air bubbles. Record the initial
burette reading.

vii. Start with Flask #1. This flask will serve as an orientation trial, to rapidly determine the
approximate end-point volume of the titration. Add 2 to 3 drops of phenolphthalein solution to
the flask and gently swirl. Position the flask beneath the burette. Place a white piece of paper
beneath the flask. This will help you visualize the colour change of the indicator.

viii. Begin delivering the titrant to the Flask #1 by opening the stopcock to achieve a steady
stream of the titrant. As the titrant is being delivered, continuously swirl the flask in a circular
motion. As the titrant enters the reaction mixture, you will observe a pink colour appearing and

3-7
then rapidly disappearing as you swirl the mixture. This indicates that the titrant is neutralizing
the acid, and in slight excess locally where the drop of titrant is entering the reaction mixture:
upon mixing the titrant is then encountering more acid, and being neutralized.

As the titration proceeds, the pink colour may persist for a few seconds before disappearing with
constant swirling. This indicates that you are approaching the end-point. Continue swirling and
at some point the addition of the titrant will cause the reaction mixture to turn pink permanently.
This is the end-point of the titration. Immediately close the stopcock to stop the flow of titrant.

ix. Take the final burette reading and record this in your lab report. The difference between the
final reading and the initial reading is the volume of titrant delivered and the end-point volume of
the orientation trial. Since the orientation trial was carried out rapidly, it likely carries with it a
large systematic error. However, it serves to rapidly determine the approximate end-point
volume.

x. Repeat steps vii through ix for Trial 2 using Flask 2 with the following modifications: titrate
rapidly until you approach the approximate end-point (1 to 2 mL less titrant than the orientation
trial) and then slow the delivery of titrant to about 1 to 2 drops per second. Swirl the flask
continuously. As you approach the end-point (pink colour persisting for longer periods) slow the
delivery of titrant to 1 drop per second or less. At this stage, be very careful to swirl between
every drop of titrant added and visually assess the colour of the mixture. The goal will be to
achieve a permanent faint pink colour without overshooting the end-point. The lighter pink the
mixture, the closer to the actual end-point, and the more accurate the titration. Stop the delivery
of titrant as soon as the mixture turns permanently pink. You may wish to stop delivery of the
titrant after each drop added as you swirl and assess the colour. You may also wish to deliver
‘split drops’ of titrant – a fraction of a drop at a time to achieve the most accurate results. Your
instructors will demonstrate this technique.
Once you have achieved a permanent pink end-point, stop the flow of titrant and record the final
burette reading. The end-point volume is the difference between the final and the initial volume.

xi. Repeat step x for Trials 3 and 4. Assuming Trial 2 was carried out accurately and without
issue, the end-point volume for Trials 3 and 4 should be very similar. If you are confident, you
can deliver this volume (minus approximately 0.5 mL) very rapidly, and then slow the titration
as described above.

xii. For your calculations, ignore Trial 1 (the orientation trial). It is not reliable and should not be
considered a valid data point. Take the end-point of the titration to be the average volume of
titrant delivered in Trials 2, 3 and 4.

All liquid containing NaOH and/or HCl from this experiment that is not contaminated with
KHP can be discarded down the sinks at your bench.

Wash all equipment with soap, brush and hot water at the sinks at the periphery of the labs.

3-8
3-9
Experiment 3 - Report Sheet Name ________________________

Orientation
Part A. Standardization of NaOH Trial 2 Trial 3 Trial 4
Trial 1

Molar Mass of KHP (C8H5O4K) 204.22 g/mole

Mass of KHP obtained

Moles of KHP obtained

Initial Burette Reading

Final Burette Reading

Volume of NaOH Solution A


Delivered to reach end-point

Moles of NaOH neutralized

Molar Concentration of NaOH


solution A
Average Molar Concentration of
NaOH solution A

Sample Calculations:

3-10
Experiment 3 - Report Sheet Name ________________________

Orientation
Part B. Titration analysis of HCl Trial 2 Trial 3 Trial 4
Trial 1
Molar Concentration of NaOH
Solution B (Provided)
Volume of HCl solution delivered into
the flask by the pipette

Initial Burette Reading

Final Burette Reading

Volume of NaOH Solution B delivered


to reach end-point

Moles of NaOH neutralized

Moles of HCl neutralized

Molar concentration of the unknown


HCl solution
Average molar concentration of the
unknown HCl solution

Sample Calculations:

3-11
Experiment 3 - Prelab Assignment Name ________________________

1. A NaOH solution of unknown concentration was standardized with potassium hydrogen


phthalate, a monoprotic acid with chemical formula C8H5O4K (204.22 g/mole). A total of 0.5760
g of KHP was obtained and titrated with the NaOH solution. 18.75 mL of the NaOH solution
was required to reach the end-point of the titration. What was the molar concentration of the
NaOH solution?

2. In a titration, 50.00 mL of 0.875 M HCl solution was titrated with 0.506 M NaOH solution.
What volume of the NaOH solution would be required to reach the end-point of the titration?

3. True or False: A titration requires two people; one person swirls the flask containing the
analyte and the other person operates the valve on the burette to control the flow of titrant.
Briefly justify your answer.

4. Where should solutions containing KHP, NaOH and HCl be disposed?


a. In the sink at your bench
b. In the fume hood sink
c. In the fume hood waste container
d. In the large sinks at the corners of the lab

5. Where should solutions containing NaOH and HCl be disposed?


a. In the sink at your bench
b. In the fume hood sink
c. In the fume hood waste container
d. In the large sinks at the corners of the lab

3-12
4-1
Experiment 4: Titrations II – Analysis of Organic Acids

This experiment should be performed and submitted in pairs.

Objective:
To experimentally determine the identity of an unknown organic acid based on its melting point,
molar mass and relatively solubility.

Prelab Assignment:
-Read and understand the protocol for Experiment 4, including the calculations required.
-Review sections III.G and III.H in the Fundamentals of Analysis section at the beginning of
this manual.
-Review the introduction to Experiment 3
-Complete the prelab assignment on the last page of this experiment.

Introduction:

The most common class of organic acids is the carboxylic acid. Carboxylic acids such as acetic
acid (CH3COOH or HC2H3O2) bear a carboxyl group which is weakly acidic. The carboxyl
group has the general structural formula R-COOH, where R indicates atoms or groups of atoms
bonded to the carboxyl group.

HO R

carboxyl group

Carboxylic acids are polar substances. These molecules have relatively high melting points
(above 100 °C). Low-molecular-weight carboxylic acids are relatively soluble in polar solvents
like water, but depending on the size and hydrophobicity of the overall molecule, the solubility
in water may be poor. Relatively hydrophobic carboxylic acids tend to be more soluble in
solvents less polar than water, like ethanol.

The carboxyl (COOH) groups common to carboxylic acids are weakly acidic. The proton bonded
to oxygen in the carboxyl group is about 106-fold less acidic than strong acids like nitric acid, but
still about 1010-fold more acidic than water; thus carboxylic acids are classified as weak acids.
Importantly, many organic acids bear non-acidic protons. For example, the C-H bonds in acetic
acid are about 1045 times less acidic than water: in other words, the C-H bonds in acetic acid are
essentially non-dissociable in aqueous conditions. For the purposes of this course, only the O-H
bonds in carboxylic acids should be considered reactive in acid-base reactions. Thus, a
carboxylic acid would react with a strong base as follows:

O O

HO R
+ NaOH
-O + Na+ + HOH (l)
-

4-2
Some other important examples of carboxylic acids are: formic acid (HCHO2), oxalic acid
(H2C2O4), benzoic acid (HC7H5O2) and citric acid (H3C6H6O7). The structures of these acids are
given below:
benzoic acid citric acid
formic acid acetic acid oxalic acid
O O OH
O O
O O HO O
HO
HO OH
HO H HO O OH HO

1 carboxyl group 1 carboxyl group 2 carboxyl groups 1 carboxyl group 3 carboxyl groups
1 acidic proton 1 acidic proton 2 acidic protons 1 acidic proton 3 acidic protons
46 g/mole 60 g/mole 90 g/mole 122 g/mole 192 g/mole

As illustrated in the examples above, organic acids may have one, two, three or more carboxyl
groups. Thus the stoichiometric ratio of the carboxylic acid to strong base in a neutralization
reaction will vary from acid to acid.

In this experiment your objective will be to experimentally determine the identity of an unknown
carboxylic acid.

You will experimentally determine the melting point of the unknown, and the unknown’s relative
solubility in water vs. an aqueous ethanol solution. Finally, a titration of a solution of the
unknown with a standardized base solution will be performed. In this case, the titration will
allow the molecular mass of the unknown acid to be determined. Taken together, these data will
allow us to determine the identity of the unknown carboxylic acid from a list of possibilities
provided by your instructor.

Methods:

You will be assigned an unknown organic acid. Record the unknown # on your report sheet in
the space provided.

Part A. Melting point determination

i. Acquire a pea-sized sample of your assigned organic acid on a clean and dry watch glass. Place
the watch glass on your bench and supporting it with one hand, use your free hand to grind the
solid powdered sample with the end of a spatula. Grind the solid to reduce the size of the crystals
to a fine powder. This will improve the accuracy of your melting point analysis. The more finely-
and consistently-ground your sample, the better and more accurate your melting point
determination will be.
ii. Fill a melting point capillary tube with some of the unknown acid, by pushing the open end of
the tube into a pile of the finely crushed sample on your watch glass.

Tap the closed end of the tube on the bench top to drop the powdered sample into the bottom
(closed end) of the capillary tube.

4-3
Repeat these steps until you have acquired a sample with a packed volume of 1-2 mm in the
bottom (closed end) of the capillary tube.

iii. Insert the capillary tube into one of the three small slots in Mel-Temp apparatus, preferably
the center slot, with the closed end of the tube pointing down. See your instructor for guidance.
Place a digital thermometer in the thermometer port. Turn the heater switch to the ON position
and set the heating rate on dial 7. Heat the sample using this setting until the thermometer
indicates a temperature of 80°C at which point the heating rate dial should be adjusted to a final
setting given by the lab instructor (dial # written on the board).

iv. Monitor your sample through the view port. Record the temperature at which the crystals
begin to melt. This is the initial melting point. Very soon after the initial melting point, you will
need to record the temperature at which the solid has completely melted. This is the final melting
point. Report the melting point range by reporting BOTH the initial and final melting points
separated by a hyphen.

For example: MP = 135.2 - 138.2 °C

Be patient. Some of the organic acids have very high melting points, and it will take several
minutes for the Mel-Temp apparatus to reach this temperature. If you find that the temperature
reading on the thermometer does not increase for 5 minutes, the heating rate setting may be
increased by increments of 0.5 every 2-3 minutes.

Capillary tubes and any excess organic solids should be disposed of in the solid waste
container in the fume hood.

Part B. Relative Solubility and Titration of the organic acid

i. You will need three clean 125 mL Erlenmeyer flasks. Label each flask, 1, 2 and 3. The outside
of the flasks must be completely dry. Each flask will contain the analyte for one titration trial.

ii. Place flask #1 on the analytical balance, and tare the balance. Weigh between 0.10 to 0.12 g of
your organic acid into the flask. Make sure to record the mass of organic acid actually obtained
on your report sheet. Repeat this step for each of the three flasks. The mass obtained does not
need to be highly consistent between trials.

iii. Use a clean graduated cylinder to deliver 15 mL of distilled water to each flask. Gently swirl
the flask. If your acid is soluble in water, it should dissolve rapidly. If your acid dissolves in
water, add another 15 mL of distilled water to each flask. Skip step iv and proceed directly to
step v. If your acid does not dissolve after 2-3 minutes of swirling, your acid is likely insoluble in
water. Proceed to step iv.

iv. If your acid is relatively insoluble in water, add 15 mL of ethanol to each flask in the fume
hood and gently swirl. In the solvent mixture of ethanol and water, your acid should rapidly
dissolve.

4-4
v. A standardized NaOH solution will be provided, and the concentration will be given at the
beginning of the lab period. Record this concentration on your report sheet in the space provided.

v. Clean a burette and rinse it with distilled water. Pour about 70 mL of the NaOH solution from
the stock bottle into a clean 100 mL beaker. Rinse the burette with a small volume of the
standardized NaOH solution.

vi. Fill the burette with the standardized NaOH solution and prepare for the titration.

vii. Start with Flask #1. Add a few drops (2-3 drops) of phenolphthalein solution to the flask and
gently swirl. Begin delivery by opening the stopcock to achieve a steady dripping flow of the
titrant.

As you approach the end-point of the titration it is advisable to reduce the flow rate to 1 drop per
second or less. At the end-point of the titration, the addition of the titrant will cause the reaction
mixture to turn pink permanently. Immediately close the stopcock to stop the flow of titrant at
the first sign of permanent faint pink colour. The goal is to achieve the faintest permanent pink
colour possible, so as to approach the end-point most accurately.

viii. Take the final burette reading and record this in your lab report.

ix. Repeat steps vi through vii for Flasks 2 and 3.

Excess NaOH solution from your titration can be disposed of down the sink at your bench.
Dispose of all liquid waste containing organic acids in the liquid waste container in the fume
hood.

Wash all equipment with soap, brush and hot water at the sinks at the periphery of the labs.

4-5
Experiment 4 - Report Sheet Name _________________________

Unknown Organic Acid #_____________

Part A. Melting point determination

Observed melting point range: __________________ °C

Part B. Solubility and Molecular Mass Determination

Was the organic acid soluble in H2O? Yes / No

Trial 1 Trial 2 Trial 3

Concentration of the standardized NaOH solution

Mass of the unknown organic acid

Initial burette reading

Final burette reading

Volume of NaOH solution delivered to reach end-point

Moles of NaOH required to neutralize the organic acid

Mass of the organic acid per mole of NaOH reacted

Average mass of the organic acid per mole of NaOH reacted

Sample Calculations:

4-6
Experiment 4 - Report Sheet Name _________________________

Your organic acid has either 1, 2 or 3 carboxyl groups; thus it has either 1, 2 or 3 acidic protons,
respectively. Depending on the number of acidic protons your organic acid has, one of three
possible balanced reactions occurred during the titration.

Complete the following reactions to give a balanced formula equation for each of the three
possible scenarios:

Scenario 1: R-COOH (aq) + NaOH (aq) R-COO- (aq) +

Scenario 2: R-(COOH)2 (aq) + 2 NaOH (aq) R-(COO)2 2- (aq) +

Scenario 3: R-(COOH)3 (aq) + 3 NaOH (aq) R-(COO)3 3- (aq) +

Based on the three possible scenarios above, complete the table below by calculating the three
possible molecular masses of your unknown organic acid.

Average molar mass of organic acid if the acid is monoprotic (Scenario 1)

Average molar mass of organic acid if the acid is diprotic (Scenario 2)

Average molar mass of organic acid if the acid is triprotic (Scenario 3)

Once you have completed the table above, your lab instructor will provide you with a datasheet
of organic acids.

Conclusion

Using all the available data (melting point, relative solubility, and possible molecular masses),
determine the identity of your unknown carboxylic acid. Compare your data to the data given for
the possible carboxylic acids on the data sheet provided. Choose the carboxylic acid that best fits
your data.

*Note that molecular masses empirically determined by titration in this way are often slightly
lower than the actual molecular mass, due to systematic error associated with using
phenolphthalein as the indicator: in fact, phenolphthalein indicates an end-point slightly past the
equivalence point of this titration. Likewise, be aware that empirically determined melting points
are also often slightly lower than the literature values due to impurities in the samples, and/or
less-than-ideal grinding and preparation of the sample prior to melting point determination.

The unknown acid was most likely _____________________which has _____ carboxyl groups
in its structure.

4-7
Experiment 4 - Prelab Assignment Name _________________________

1. A 0.219 g sample of an unknown organic acid was titrated with 0.0934 M NaOH; a total of
39.52 mL of the base solution was required to reach the end-point. The acid is known to have 2
carboxyl groups per molecule. What was the molecular weight of the unknown acid? Show all
calculations.

2. The melting point of the unknown acid in question 1, above, was found to be 179-182°C and
the acid was soluble in water. Which of the following acids is most likely to be the unknown acid?

Molecular Mass Solubility Melting Point


Acid 1 118 g/mole water soluble 147 - 149 °C
Acid 2 117 g/mole water soluble 170 - 172 °C
Acid 3 120 g/mole water soluble 185 - 186 °C

4-8
5-1
Experiment 5: Titrations III – Redox Titrations

This experiment should be performed and submitted individually.

Objectives:
1. To standardize a solution of KMnO4 of unknown concentration by a titration analysis.
2. To quantitatively determine the Fe2+ content of a commercially available iron supplement
tablet by titration analysis.

Prelab Assignment:
-Read and understand Experiment 5, including the required calculations
-Review sections III.G and III.H in the Fundamentals of Analysis section at the beginning of
this manual.
-Review the introduction to Experiment 3
-Complete the prelab assignment on the last page of this experiment.

Introduction:

Although the majority of titrations carried out in introductory labs are acid-base titrations, it
bears mentioning that the term ‘titration’ is general, rather than specific to acid-base reactions.
As was discussed in the introduction to Experiment 3, a titration is simply a quantitative
combination of the reagents that participate in a reaction. A wide variety of reactions (not just
acid-base reactions), can be carried out with a titration.

In this experiment, we will explore this concept with a redox titration.

So-called ‘redox’ reactions are a class of reaction in which one of the reactants loses electrons;
the species that loses electrons over the course of the reaction is oxidized. The ‘lost’ electrons are
donated to another reactant; the species that gains these electrons over the course of the reaction
is reduced. In such a reaction, an oxidation (a loss of electrons) is always paired with a reduction
(a gain of electrons). Thus these reactions are reduction-oxidation reactions, or red-ox (redox)
reactions.

It is often convenient to write a redox reaction as the sum of two half-reactions, which
conceptually separates the reduction process from the oxidation process. Consider the following
(unbalanced) example in which permanganate ion (MnO4-) reacts with oxalate ion (C2O42-) in
acidic conditions:

MnO4- (aq) + C2O42- (aq) Mn2+ + CO2 in acidic solution

This reaction can be separated into two half-reactions as follows:

MnO4- (aq) Mn2+ (aq) C2O42- (aq) CO2 (aq)

5-2
Separating the reaction into half-reactions is likely the most comprehensive available method of
balancing this type of redox reaction. In brief, assigning oxidation states to all species in each
half reaction will allow us to account for the electrons lost/gained in each case as follows:

5 e- + MnO4- (aq) Mn2+ (aq) C2O42- (aq) 2 CO2 (aq) + 2 e-


+7 -2 +2 +3 -2 +4 -2

Thus the MnO4- (aq) is reduced: each Mn gains 5 e- as it is reduced to form Mn2+. The C2O42- is
oxidized: each C gains 1 e- as it reacts to produce CO2. The two half reactions can never really
occur in complete isolation as we have written them above. Rather, the half reactions are just a
convenient way of thinking about the process.

Notice that for the oxidation half reaction, mass and charge are conserved, so by all indications
this half reaction is accurately balanced. However, the reduction half reaction is not truly
balanced as written above. In brief, in acidic solution, some water and protons will participate in
the reaction as follows:

8 H+ (aq) + 5 e- + MnO4- (aq) Mn2+ (aq) + 4 H2O (l) C2O42- (aq) 2 CO2 (aq) + 2 e-

At this point, the reduction half reaction is now balanced as both mass and charge are conserved.

Note: in alkaline (basic) conditions, the reaction will have to be balanced differently. For this
lab, we will use acidic conditions exclusively.

The balanced overall reaction is then taken to be the net sum of the two half reactions.
Importantly the reaction will always occur as a direct exchange of electrons between the half-
reactions. Thus the law of conservation of mass applies, even to electrons, and so the electrons
exchanged between the two half reactions must necessarily be balanced. The number of
electrons lost in the oxidation half reaction will equal the number of electrons gained in the
reduction half reaction.

In this case, the reduction half reaction involves a gain of 5 e- and the oxidation half reaction a
loss of
-
2 e . The oxidation half reaction will have to occur five times for every two times that the
reduction half reaction occurs. We can account for this and re-combine the balanced half
reactions as follows:

[ 8 H+ (aq) + 5 e- + MnO4- (aq) Mn2+ (aq) + 4 H2O (l) ] x 2

[ C2O42- (aq) 2 CO2 (aq) + 2 e- ] x 5


_____________________________________________

16 H+ (aq) + 10 e- + 2 MnO4- (aq) + 5 C2O42- (aq) 10 CO2 (aq) + 2 Mn2+ (aq) + 8 H2O (l) + 10 e-

Notice that in doing so, the electrons consumed (gained) are equal to the electrons produced
(lost). Thus there is no net consumption or production of electrons. This indicates that there is a

5-3
direct and complete exchange of electrons between the two half reactions. Thus the balanced net
reaction would be written:

16 H+ (aq) + 2 MnO4- (aq) + 5 C2O42- (aq) 10 CO2 (aq) + 2 Mn2+ (aq) + 8 H2O (l)

Thus we would predict a stoichiometric ratio of MnO4- to C2O42- of 2:5 for this reaction, not 1:1
as we had when we introduced this reaction on the previous page.

In this experiment we will carry out two redox titrations. First, we will use a potassium oxalate
(K2C2O4) solution of known concentration to standardize a solution of potassium permanganate
(KMnO4) under acidic conditions. As described above, this reaction will result in the reduction
of MnO4- to Mn2+ and the oxidation of C2O42- to CO2.

Then we will use the standardized solution of KMnO4 to analyze the Fe2+ content of a
commercially available iron supplement tablet. In acidic conditions, the MnO4- will be reduced
to Mn2+ and any Fe2+ in the sample will be oxidized to Fe3+ according to the following
unbalanced reaction:

MnO4- (aq) + Fe2+ (aq) Mn2+ (aq) + Fe3+ (aq) in acidic solution

All of the species in both reactions in this experiment are colourless in solution, except for
MnO4- which is a beautiful dark purple colour in solution. This property of MnO4- will obviate
the need for a colour indicator in this experiment. That is, as purple KMnO4 solution is combined
with the other reagents in the two titrations, the result will be a colourless solution so long as the
MnO4- remains in limiting quantities. That is, before the end-point is reached, all of the MnO4-
added to the reaction mixture is converted to colourless Mn2+. Once the end-point of the
titration is reached, any additional KMnO4 solution added should result in a permanent
pink / purple solution: i.e. MnO4- will be in excess at the end-point and beyond, and therefore
the reaction mixture turns permanently light pink to purple. The first appearance of permanent
pink / purple colour in the reaction mixture will indicate the end-point of each titration, and this
will approximate the point at which a stoichiometric ratio of reagents has been combined in each
case.

The analysis of the iron tablet presents several additional challenges. While the Fe2+ in the tablet
will be soluble under the conditions of this experiment, the tablet contains many other chemicals,
some of which are insoluble in aqueous solution.

Unfortunately, the tablet also contains many impurities that will react with MnO4-, such as
cellulose and other organic compounds. Moreover, the red dye coating the tablet will result in the
solution turning pink. This will make it impossible to detect the end-point of the titration with
KMnO4, which is also pink / purple. Thus, we will have no choice but remove the red dye and
other impurities from iron tablet, before the titration. We will extract the impurities in the iron
tablet solution by combining it with activated charcoal. Since activated charcoal is very
hydrophobic, it will not dissolve in aqueous solution; it is insoluble in water and will remain a
solid when combined with aqueous solution. Since many of the impurities in the tablet, including
the red dye, are relatively non-polar / hydrophobic, they will bind to the activated charcoal. This

5-4
allows us to extract the hydrophobic impurities from the aqueous solution. Fe2+ in the tablet
should not bind to the charcoal, because aqueous ions are highly polar / hydrophilic. The
activated charcoal and the impurities bound to it can then be removed from the solution by
passing it through a piece of filter paper. The activated charcoal and bound impurities should be
retained by the filter, because the activated charcoal is a solid and will not dissolve in aqueous
solution. The soluble Fe2+ ions should pass through the filter with the solution. The Fe2+ solution
that passes through the filter is suitably pure for titration analysis.

This two-part experiment will directly test your expertise with titrations. You will not be
required to make any solutions accurately: the solutions will be prepared by your lab instructors
and provided for you. However, the results in the first part of the experiment will directly affect
the accuracy of the second. Any errors you make during the standardization of the KMnO4
solution will be carried forward and compounded in your analysis of the iron (II) content of the
unknown. Good luck!

Methods:

Safety notice: Sulfuric acid (H2SO4) is a dangerously corrosive substance. Handle with care.
During this titration we will be heating a solution containing H2SO4. Do not allow this
solution to boil. If boiling occurs at any point, reduce the heat setting on the hotplate by 10 °C,
and immediately remove the sample from the heating surface and allow it to cool for a few
minutes before continuing.
You should also be careful handling the KMnO4 solution: although it poses no significant
health and safety risks, it will stain your skin, clothing, the floor, etc. If you splash H2SO4 or
KMnO4 on your skin, wash with soap and water immediately.

Part A. Standardization of a KMnO4 solution with K2C2O4

i. A K2C2O4 (aq) standard solution in 0.2 M H2SO4 will be provided. The concentration of K2C2O4
will be provided during the lab period. You will need three, clean, labeled Erlenmeyer flasks.
Take approximately 40 mL of K2C2O4 solution from the stock bottle in a 50 mL transfer
beaker. Measure 10.00 mL of the K2C2O4 standard solution to each of the flasks.

ii. The redox reaction between KMnO4 and K2C2O4 is very slow at room temperature - far too
slow for a titration. Heat will be used to accelerate the reaction so it can be accomplished in a
reasonable amount of time.

Place the flasks on a hotplate, and turn on the heat at a setting of 150 °C. Do not leave the
solutions on the hotplate unattended. At some point you will see condensation appearing on the
inside of the flasks. At this point, reduce the heat setting to 120°C. Do not allow the solution to
boil.

iii. While you are waiting for the solutions to heat up, adjust your burette so that it will deliver
KMnO4 into a flask on the hotplate surface. The hotplate will have a clamp and burette stand
attached, with a KMnO4 solution as the titrant already in the burette.

5-5
iv. Begin the titration by delivering approximately 1 mL of the KMnO4 titrant solution to flask #
1 and swirling the flask continuously, keeping it on the hot plate as much as possible. Once the
solution in the flask turns colourless, the solution has reached an appropriate temperature for the
titration.

v. Titrate the hot, acidic K2C2O4 solution with the titrant (KMnO4 solution). The flask should be
swirled continuously, but keep it on the hotplate as much as possible. Again, do not allow the
solution to boil. Each drop of titrant added should turn from purple to colourless (or sometimes
faint yellow) almost instantaneously, or after a moment of swirling. Continue until the reaction
mixture turns light pink/purple permanently. This is the end-point of the titration.

vi. Repeat steps iv. through v. for flask # 2 and flask #3.

vii. Keep your hotplate plugged in for Part B of the experiment.

The contents of flasks from Part A should be disposed of in the waste container in the fume hood.

Part B. Analysis of the iron (II) content of an iron supplement tablet

i. Add 25.0 mL of 0.5 M phosphoric acid to a clean Erlenmeyer flask using a graduated cylinder.
Return to your bench, and place the flask on the hotplate at a heat setting of 150 °C. Heat the
solution until it reaches about 80 °C. You should see condensation forming on the inner walls of
the flask, or heat for about 5 minutes, whichever comes first. Do not allow the solution to boil.

ii. Acquire an iron supplement tablet. Fold a large sheet of weighing paper in half to create a
crease across the center of the sheet. Weigh the iron tablet using the folded weighing paper.
Record the mass of the iron tablet on your report.

iii. Add the iron tablet to the Erlenmeyer flask containing the warm 0.5 M phosphoric acid
solution on the hot plate.

iv. While you are waiting for the sample to dissolve, set up the filtration apparatus. Attach a ring
clamp to a free burette stand. Place your funnel in the ring. Place a clean Erlenmeyer flask
beneath the suspended funnel, so that the funnel will drain directly into the flask. Take a circular
piece of filter paper and fold it in half to form a semicircle. Fold the folded filter paper in half
again, to form a cone, and place the folded paper in the funnel.

v. Once the iron tablet solution is finished heating, remove the flask form the hotplate and add
2/3 of a spatula of Norit (activated charcoal powder) to the flask and swirl. The activated
charcoal will not dissolve. Continue heating the flask for about 1 minute with intermittent
swirling to mix the contents.

5-6
vi. Pour the contents of the flask through the filter paper in your funnel, collecting the filtrate
(the liquid that passes through) into a clean Erlenmeyer flask. The filtrate should be clear,
colourless and homogenous solution.

vii. No heating is required for this titration. Unplug your hotplate and set it aside to cool.
Titrate the filtrate containing Fe2+ from the iron tablet with the KMnO4 solution. Each drop of
titrant added should turn from purple to colourless after a moment of swirling. Continue until the
reaction mixture turns light pink/purple permanently. This is the end-point of the titration.
Record the final burette reading.

The contents of flasks from Part B should be disposed of in the waste container in the fume
hood. The filter paper containing Norit and other impurities should be disposed of in the
solid waste in the fume hood. Wipe up any spills on the hotplate.

Wash all equipment with soap, brush and hot water at the sinks at the periphery of the labs.

5-7
Experiment 5 - Report Sheet Name ________________________

Part A. Standardization of KMnO4 solution

Trial 1 Trial 2 Trial 3

Concentration of K2C2O4 in standard solution

Volume of K2C2O4 solution delivered 10.00 mL 10.00 mL 10.00 mL

Initial Burette Reading

Final Burette Reading


Volume of KMnO4 solution required to reach
end-point
Molar Concentration of KMnO4 in the titrant
Average Molar Concentration of KMnO4
solution

Sample Calculations:

5-8
Experiment 5 - Report Sheet Name ________________________

Part B. Analysis of iron (II) content of an iron supplement tablet

Provide the balanced equation for the net redox reaction that occurs between Fe2+ and MnO4-:

Mass of iron supplement tablet

Initial Burette Reading

Final Burette Reading

Volume of KMnO4 delivered

Average molar concentration of KMnO4 solution (from Part A)

Moles of MnO4- reacted

Moles of Fe2+ reacted

Mass of Fe2+ in the supplement tablet

% By Mass of Fe2+ in the supplement tablet

Calculations:

5-9
Experiment 5 - Prelab Assignment Name ________________________

1. Balance the following redox reaction in acidic solution:

MnO4- (aq) + Fe2+ (aq) Mn2+ (aq) + Fe3+ (aq)

2. A sample consists of a mixture of Fe(NO3)2 (s) and Fe(NO3)3 (s) in unknown proportion. A total
of 3.65 g of the mixture was dissolved in phosphoric acid solution and titrated with 0.135 M
KMnO4. A total of 11.79 mL of the KMnO4 solution was required to reach the end-point of the
titration. Note: as per the reaction equation in question 1, above, KMnO4 reacts only with
Fe(NO3)2 in the mixture: the KMnO4 will not react with Fe(NO3)3.

What was the % by mass of Fe(NO3)2 in the solid mixture?

3. Why does condensation appear on the inside of the flask as is described in the methods during
heating?

5-10
4. What instrument(s) would you use to measure and deliver the following volumes of liquid?
Circle your answers.

Graduated Volumetric Mohr


a. 10. mL of liquid: Beaker
Cylinder Pipette Pipette
Graduated Volumetric Mohr
b. 10.0 mL of liquid: Beaker
Cylinder Pipette Pipette
Graduated Volumetric Mohr
c. 10.00 mL of liquid: Beaker
Cylinder Pipette Pipette
Graduated Volumetric Mohr
d. 9.75 mL of liquid: Beaker
Cylinder Pipette Pipette

5. A total of 9.73 mL of KMnO4 solution was required to reach the end-point in a titration with
15.00 mL of a 0.0225 M K2C2O4 solution. What was the molar concentration of the KMnO4
solution?

5-11
Experiment 6: Periodic Trends

This experiment should be performed and submitted in pairs.

Objectives:
1. To determine the relative reactivity of a series of main-group metal elements.
2. To determine the relative reactivity of the Group 17 elements (the halogens).

Prelab Assignment:
-Read and understand Experiment 6
-Complete and balance the reaction equations for Parts A and B, found in the two tables in the
report sheet.

Introduction:

Part A: The main-group metals are generally considered to consist of metallic elements
excluding the transition metals, metalloids and the lanthanides and actinides: specifically, the
alkali metals (group 1: Li, Na, K, Rb, Cs and Fr), the alkali earth metals (group 2: Be, Mg, Ca, Sr,
Ba and Ra), metals in the boron group (group 13: Al, Ga, In, and Tl) as well as Sn, Pb and Bi.
Here we will examine a selection main group metals from groups 1, 2 and 13: Na, K, Mg, Ca and
Al. These elements share many properties: they are all lustrous, soft, reactive metallic solids in
the pure state. In particular, the group 1 and group 2 metals are strong reducing agents. Due to
their reactivity, in nature these metals are most abundantly found as cations in mineral salts. That
is, in the pure metallic state, these metals readily lose electrons, oxidizing to form metal cations
given a suitably strong oxidizing agent, such as O2, halogens, mineral acids or water. For
example:

2 Na (s) + 2 H2O (l) 2 NaOH (aq) + H2 (g)

In this redox reaction, metallic sodium is oxidized: it loses electrons to form sodium ions.
Protons from water are reduced: they gain electrons to form hydrogen gas. Thus Na (s) is the
reducing agent, and H2O is the oxidizing agent. Unlike the redox reactions we examined in
Experiment 5, the above redox reaction does not necessarily require the half-reaction method to
be balanced. The above redox reaction can be balanced by inspection. If we examine the change
in oxidation states for each atom in the above reaction, we see that it is, indeed, a redox reaction,
and that the electron exchange is already balanced:

2 Na (s) + 2 H2O (l) 2 NaOH (aq) + H2 (g)


0 +1 -2 +1 -2 +1 0

That is, each sodium loses 1 e- as it is oxidized, and each proton gains 1 e- as it is reduced. A
total of 2 e- are lost and a total of 2 e- are gained in the reaction as written above: mass and
charge are conserved, and thus by all indications the redox process as written above is balanced.

6-1
Notably, different metallic solids undergo this type of reaction at different rates. For simplicity in
this experiment, we will take the relative rate of reaction to be a measure of the ‘reactivity’ of the
metallic solid.

Some metallic solids may not react with water at an appreciable rate. In part, this may be due to a
coating of corrosion (metal oxide) found on their surface that is relatively insoluble in aqueous
solution. For example, insoluble aluminum oxide on the surface of Al (s) renders the reaction with
water very slow. In these cases, a strong acid solution such as HCl (aq) may be substituted for
pure water to consume the aluminum oxide, and achieve a reaction with an appreciable rate. For
example:

2 Al (s) + 6 HCl (aq) 2 AlCl3 (aq) + 3 H2 (g)

HCl is a stronger oxidizing agent than water, and a better proton donor. These characteristics
tend to further enhance the rate and extent of the reaction of metallic solids with HCl solution
relative to pure water.

In this experiment, we will compare the relative reactivity of a selection of metallic solids
including Na, K, Mg, Ca and Al with water (and/or strong acid solution) to illustrate the trend in
the ability of these elements to ionize.

Part B: The halogens are reactive elements known for their affinity for electrons and high
electronegativity. Due to their reactivity, in nature the halogens exist primarily as halide ions (e.g.
F-, Cl-, Br - and I-) in mineral salts with metal cations. In their pure forms, fluorine, chlorine,
bromine and iodine form diatomic molecules, F2, Cl2, Br2 and I2, respectively. Under standard
conditions, F2 and Cl2 are gases, Br2 is a liquid and I2 is a solid. F2 reacts violently with water,
and thus cannot be used to prepare solutions in water. The other diatomic halogens are less
reactive than F2 and can be used to prepare aqueous solutions. However Cl2, Br2 and I2 are only
moderately soluble in water. Due to their non-polarity, the diatomic halogens are much more
soluble in relatively non-polar organic solvents such as hexane (C6H14), carbon tetrachloride
(CCl4) or dichloromethane (CH2Cl2). Thus, in a mixture of water and a non-polar solvent,
diatomic halogens partition with the non-polar solvent.

Diatomic halogens can participate in so-called replacement reactions with metal-halide salts. For
example:

Cl2 (aq) + 2 NaI (aq) 2 NaCl (aq) + I2 (aq)

The above reaction is another redox process. In this case, chlorine atoms gain electrons to
produce chloride ions, and iodide ions lose electrons, producing iodine atoms that bond to form
the diatomic molecule I2. Thus, Cl2 is the oxidizing agent, and NaI is the reducing agent in the
above reaction.

6-2
Like the redox reactions discussed in Part A, above, this reaction belongs to a class of redox
reactions that can be balanced by inspection. If we examine the change in oxidation states for
each atom in the above reaction, we see that it is, indeed, a redox reaction, and that the electron
exchange is already balanced for the reaction as written:

Cl2 (aq) + 2 NaI (aq) 2 NaCl (aq) + I2 (aq)


0 +1 -1 +1 -1 0

That is, each chlorine atom gains 1 e- as it is reduced, and each iodide ion loses 1 e- as it is
oxidized. A total of 2 e- are lost and a total of 2 e- are gained in the reaction as written above:
mass and charge are conserved, and thus by all indications the redox process as written above is
balanced.

Also note that Na+ is not an active participant in the reaction. Na+ is merely a spectator in this
reaction, and thus the net reaction would be written:

Cl2 (aq) + 2 I- (aq) 2 Cl- (aq) + I2 (aq)

The extent to which the forward reaction progresses can be taken as a measure of the relative
strength of Cl2 as an oxidizing agent. If the forward reaction occurs to an appreciable extent, it
would be concluded that Cl2 must be a strong enough oxidizing agent to oxidize I-.

In this experiment we will compare the relative ability of the diatomic halogens Cl 2, Br2 and I2 to
oxidize halide ions Cl-, Br - and I - to illustrate the trend in the relative oxidizing ability of these
elements.

Methods:

Part A. Reactivity of metals in groups I, II, and III

i. Demo performed by the instructor.


To compare the relative reactivity of sodium and potassium metals with water a small piece of
sodium will be dropped into a dish of water. The demonstration will then be repeated with a
small piece of potassium metal and fresh water. At the end of both demonstrations, the water will
be tested with phenolphthalein indicator.

ii. At your bench.


Fill a 400 mL beaker with tap water. Fill a large test tube with tap water and plug the opening
with your thumb. Carefully invert the test tube in the 400 mL beaker and then remove your
thumb so that the tube remains filled with water with a minimum of air bubbles.

iii. Drop a piece of calcium metal into the water in the beaker. Position the inverted test tube so
that the piece of metal is trapped inside the tube. Observe the reaction. You may notice a white
precipitate forming and settling to the bottom of the beaker as the reaction progresses: this is
normal, and the result of the production of Ca(OH)2 which has a limited solubility in water.

6-3
Compare the relative rate of the reaction of calcium to that of sodium and potassium in water.

iv. Pour about 3 mL of 3 M HCl (aq) into each of three test tubes. Drop a small piece of
magnesium into the first tube and observe the reaction.

v. Repeat step iv with a small piece of calcium and a small piece of aluminum using tubes 2 and
3, respectively.

Compare the relative rates of the reactions of magnesium, calcium and aluminum in HCl solution.

Dispose of all aqueous waste from Part A down the sink at your bench.

Part B. Replacement reactions of the free halogens and halide salts

Safety Note: The reagents used in Part B are toxic. They should be handled in the fume hood,
because Cl2 and dichloromethane are volatile. Gloves MUST be worn in the fume hood.

In particular, the halogens are corrosive and extremely toxic. If you spill any, immediately
neutralize with sodium bisulfite (NaHSO3) provided in the fume hood.

Dichloromethane (CH2Cl2, also known as methylene chloride or DCM) is a toxic, hazardous


material. If you spill any on your hands (even onto your gloves) or skin, remove your gloves
and wash with soap and water.

Once all of the reagents are added to each test tube, securely seal the test tube with a cork.
Only then should the test tube be removed from the fume hood, and while the test tube is
outside the fume hood THE CORK MUST REMAIN SECURELY IN PLACE.

i. Label six test tubes 1 through 6 and place the tubes in a test tube rack. Prepare a series of
reaction mixtures according to the table below. Begin by adding 2 mL of the required 0.1 M
aqueous salt solution to each tube at your bench, according to the table below. Use a test tube
rack to transport your tubes to the fume hood.

ii. In the fume hood, 1 mL of dichloromethane will be added to each tube and then 1 mL of the
halogen / water solutions according to the table on the following page.

iii. Place a cork into the top of each test tube to seal their contents and return to your bench with
the test tubes in their rack. DO NOT uncork the test tubes. Observe the colour of the bottom
layer in the test tube (dichloromethane / organic layer).

6-4
Test Halogen/Water Total
Salt Solution Organic Solvent
tube Solution Volume
0.1 M KBr (aq) Cl2 / Water Dichloromethane
1 4 mL
2 mL 1 mL 1 mL

0.1 M KI (aq) Cl 2/ Water Dichloromethane


2 4 mL
2 mL 1 mL 1 mL

0.1 M NaCl (aq) Br2 / Water Dichloromethane


3 4 mL
2 mL 1 mL 1 mL

0.1 M KI (aq) Br2 / Water Dichloromethane


4 4 mL
2 mL 1 mL 1 mL

0.1 M KBr (aq) I2 / Water Dichloromethane


5 4 mL
2 mL 1 mL 1 mL

0.1 M NaCl (aq) I2 / Water Dichloromethane


6 4 mL
2 mL 1 mL 1 mL

iv. Gently and thoroughly mix the solutions in the test tubes as demonstrated by your instructor.
Avoid contaminating the cork with the contents of the test tube. DO NOT invert the test tubes to
mix: this will soak the cork, and likely allow the reagents to leak onto your hands.

The halide salt solution and the halogen/water solution combine and mix homogenously to form
an aqueous solution in the mixture. The dichloromethane is much less polar than water: due to its
hydrophobicity, dichloromethane is insoluble in water at this temperature; the two liquids are
immiscible and will not homogenously mix. Moreover, the density of dichloromethane is
1.317g/mL - much higher than that of water. Altogether, this results in the dichloromethane
separating from the aqueous solution. Due to its greater density, the dichloromethane will sink
beneath the aqueous solution, forming a distinct layer of this organic solvent (the ‘organic layer’)
beneath the upper, ‘aqueous layer’.

Notably, the free halogen (either Cl2, Br2, or I2) present in the reaction mixture will preferentially
partition with the dichloromethane, because these materials are more soluble in this relatively
non-polar environment. Essentially, the free halogen will be extracted from the reaction mixture
by the dichloromethane. Thus after mixing the contents of the test tube, and allowing the reaction
(if any) to occur, the free halogen should be drawn into the dichloromethane layer.

6-5
Once dissolved in dichloromethane, the free halogens exhibit the following distinctive colours:

Cl2 dissolved in dichloromethane is colourless.

Br2 dissolved in dichloromethane is yellow or brown depending on its concentration.

I2 dissolved in dichloromethane is pink or red depending on its concentration.

Thus by observing the colour the dichloromethane layer of each test tube, the identity of the free
halogen present can be determined. This will allow you to ascertain whether the reactions
predicted for each reaction mixture occurred to an appreciable extent.

Dispose of all waste from Part B in the fume hood waste container.
Rinse each tube with about 1 mL of water and pour this rinsing into the waste container.
Repeat until the rinsings are colourless.

Wash all equipment with soap, brush and hot water at the sinks at the periphery of the labs.

6-6
Experiment 6 - Report Sheet Name _______________________

Part A - Reactivity of metals in groups I, II, and III

Reactants Reaction Equation (Molecular) Observations

Na / water Na (s) + HOH (l)

K / water K (s) + HOH (l)

Ca / water Ca (s) + HOH (l)

Mg / HCl Mg (s) + HCl (aq)

Ca / HCl Ca (s) + HCl (aq)

Al / HCl Al (s) + HCl (aq)

Conclusions

i. Compare the relative reactivity of the three metals that reacted with water.

ii. Compare the relative reactivity of the three metals that reacted with hydrochloric acid.

iii. Summarize the relative reactivity of the metals tested horizontally (across the 3rd period) and
vertically (down groups 1 and 2).

6-7
Experiment 6 - Report Sheet Name _____________________

Part B. Replacement reactions of the free halogens and halide salts

Halogen in
Colour of
the
Test Reactants Reaction Equation (Molecular) the organic
organic
layer
layer

1 KBr / Cl2 KBr (aq) + Cl2 (aq)

2 KI / Cl2 KI (aq) + Cl2 (aq)

3 NaCl / Br2 NaCl (aq) + Br2 (aq)

4 KI / Br2 KI (aq) + Br2 (aq)

5 KBr / I2 KBr (aq) + I2 (aq)

6 NaCl / I2 NaCl (aq) + I2 (aq)

Conclusion

Which halogen was the strongest oxidizing agent? Which halogen was the weakest oxidizing
agent? State the periodic trend.

6-8
Experiment 7: Molecular Geometry

Introduction

In this exercise you will examine models of hybrid orbitals and investigate the geometry of
molecules and ions using the VSEPR model.

According to the valence shell electron - pair repulsion theory, the shape of the molecule is
determined by the repulsions between the electron pairs in the valence shell of the central atom.

In order to determine the arrangement of electron pairs both the bonding electron pairs and non-
bonding or lone pairs on the central atom must be taken into account.

To describe the molecular geometry (shape of the molecule), only the bonding electron pairs
have to be considered.

The following table illustrates how the number of electron pairs on a central atom determines the
geometry.

NUMBER OF ELECTRON BOND LONE ARRANGEMENT OF MOLECULAR


PAIRS OF CENTRAL ATOM PAIRS PAIRS ELECTRON PAIRS GEOMETRY

2 2 0 LINEAR LINEAR

3 3 0 TRIGONAL PLANAR

2 1 TRIGONAL PLANAR ANGULAR (BENT)

4 0 TETRAHEDRAL

4 3 1 TETRAHEDRAL TRIGONAL PYRAMIDAL

2 2 ANGULAR (BENT)

5 0 TRIGONAL BIPYRAMIDAL

4 1 TRIGONAL SEE-SAW
BIPYRAMIDAL
5 3 2 T-SHAPED

2 3 LINEAR

6 0 OCTAHEDRAL

6 5 1 OCTAHEDRAL SQUARE PYRAMIDAL

4 2 SQUARE PLANAR

7-1
Hybridization is a modification of the localized electron model to account for the observation
that atoms often seem to use special atomic orbitals in forming molecules. The number of
electron pairs on the central atom also determines the hybridization of the orbitals.

Both of these models (hybridization and VSEPR) are extraordinarily valuable because they
provide pictures in our minds of what molecules might look like and how they might behave.
However, as valuable as these pictures are, we must remember that they greatly oversimplify a
very complex situation. Therefore, it is important to remember that:

- the actual geometry of a molecule can only be determined experimentally.

- hybridization is just a way of explaining the experimental facts.

- the models you are using are subject to many limitations. Note these as you go along.

Experimental

In the first part you will examine the models of hybrid orbitals placed at the end of the bench and
answer all the relevant questions using the answer sheet.

In the second part you will construct molecular models for each of the molecules given, using the
instructions given below. You will also use the tables provided on the answer sheet to respond to
the following statements.

a) Write the Lewis structure.

b) Determine the number of electron pairs around the central atom (double and triple bonds
count as single bond).

c) Describe the arrangement of electron pairs around the central atom (orbital geometry).

d) Describe the arrangement of atoms around the central atom (molecular geometry). Draw a
3-D sketch of the molecule, showing bond angles.

e) Predict the polarity of each species.

7-2
NAME: ________________________ PARTNER: ______________________________

sp orbitals

Examine the model of the sp orbitals.

1) What atomic orbitals hybridize to give sp hybrid orbitals?

Answer: ______________________________________________________________

2) How many sp orbitals are formed by hybridization?

Answer: _______________________________________________________________

3) What is the geometry of sp hybrid orbitals?

Answer: _______________________________________________________________

Construct the molecular model for BeCl2 by using the 6-hole ball (grey) for Be atom. Attach
two links at 180º from each other and attach white balls (Cl atoms) to the end of the links.

Complete the table below:

SPECIES LEWIS BOND LONE ARRANGEMENT MOLECULAR 3-D SKETCH POLAR


STRUCTURE PAIRS PAIRS OF ELECTRON GEOMETRY SHOW YES/NO
PAIRS ANGLES

BeCl2

7-3
sp2 orbitals
Examine the model of the sp2 orbitals.

1) What atomic orbitals hybridize to give sp2 hybrid orbitals?

Answer: ____________________________________________________________

2) How many sp2 orbitals are formed by hybridization?

Answer: _____________________________________________________________

3) What is the geometry of sp2 hybrid orbitals?

Answer: _____________________________________________________________

a) three bonding pairs:

Construct the molecular model for BF3 by using the 5-hole ball (brown) for B atom.
Attach three links so that they all lie in the same plane. Attach white balls (F atoms) to the
end of the links.

b) two bonding pairs and one lone pair:

Construct the molecular model for SnF2 by using the model from part a). Remove one of
the white balls leaving only the link which represents the lone pair. At this point you should
note that the lone pair occupies a greater volume than the bonding pair and hence, the lone
pair/bonding pair repulsion is greater than the bonding pair/ bonding pair repulsion.

Complete the table below:

SPECIES LEWIS BOND LONE ARRANGEMENT MOLECULAR 3-D SKETCH POLAR


STRUCTURE PAIRS PAIRS OF ELECTRON GEOMETRY SHOW ANGLES YES/NO
PAIRS

BF3

SnF2

7-4
sp3 orbitals

Examine the model of the sp3 orbitals.

1) What atomic orbitals hybridize to give sp3 hybrid orbitals?

Answer: _________________________________________________

2) How many sp3 orbitals are formed by hybridization?

Answer: __________________________________________________

3) What is the geometry of sp3 hybrid orbitals?

Answer: __________________________________________________

a) four bonding pairs:

Construct the molecular model for CH4 by using the 4-hole ball (black) for C atom. Attach
four links and place a white ball (H atom) at the end of each link.

b) three bonding pairs and one lone pair:

Construct the molecular model for NH3 by using the model from part a). Remove any of
the white balls leaving only link which represents the lone pair.

c) two bonding pairs and two lone pairs:

Construct the molecular model for H2O by using the model from part b). Remove another of
the white balls leaving only the link which represents the second lone pair.

7-5
Complete the table below:

SPECIES LEWIS BOND LONE ARRANGEMENT OF MOLECULAR 3-D SKETCH POLAR


STRUCTURE PAIRS PAIRS ELECTRON PAIRS GEOMETRY SHOW YES/NO
ANGLES

CH4

NH3

H2O

7-6
dsp3 orbitals

Examine the model of the dsp3 orbitals.

1) What atomic orbitals hybridize to give dsp3 hybrid orbitals?

Answer: ___________________________________________

2) How may dsp3 orbitals are formed by hybridization?

Answer: ___________________________________________

3) What is the geometry of dsp3 hybrid orbitals?

Answer: ___________________________________________

a) five bonding pairs:

Construct the molecular model for PCl5 by using the 5-hole ball (brown) for P atom.
Attach five links and place a white ball (Cl atom) at the end of each link.

b) four bonding pairs and one lone pair:

Construct the molecular model for SF4 by using the model for PCl5

There are two different positions in this molecule: equatorial (occupied by three atoms that are
120º apart) and axial (occupied by two atoms that are 180º apart). The lone pair can be placed
either in axial or in equatorial position. Which of these two models is better?

The guiding principle to be used is to minimize the repulsions between electron pairs:

Because of their greater size, lone pairs repel others more strongly than do bond pairs, so
we can say that LP - LP repulsions are greater than LP - BP repulsions, which in turn are
greater than BP - BP repulsions.

- When trying to quantify and compare repulsions in different models you should not be
concerned with electron pairs that are more than 90º apart.

- A configuration that contains one or more LP - LP repulsions can never be correct.

7-7
In case of SF4 lone pair in equatorial position (remove one of the white balls in equatorial
position) will yield 2 LP - BP and 4 BP repulsions.

A lone pair in axial position (remove one of the white balls in axial position) will yield 3 LP - BP
and 3 BP - BP repulsions.

Obviously, the repulsion is lower if the lone pair is in the equatorial position.

c) three bonding pairs and two lone pairs:

Construct three possible molecular models for BrF3 using a five-hole ball (brown) for Br
atom, white balls for F atoms and links for the lone pairs. The lone pairs can be placed so
that they are both in the axial positions, both in the equatorial positions or one pair in the
axial and one pair in the equatorial position.

Which model is the correct structure for BrF3?

d) two bonding pairs and three lone pairs:

Construct three possible molecular models for XeF2 using a five-hole ball (brown) for Xe
atom, white balls for F atoms and links for the lone pairs. The bonding pairs can be placed
so that they are both in the axial positions, both in the equatorial positions or one pair in the
axial and one pair in the equatorial position. Which model is the correct structure for XeF2?

8
Complete the table below:

SPECIES LEWIS BOND LONE ARRANGEMENT MOLECULAR 3-D SKETCH POLAR


STRUCTURE PAIRS PAIRS OF ELECTRON PAIRS GEOMETRY SHOW ANGLES YES/NO

PCl5

SF4

BrF3

XeF2

7-9
d2sp3 orbitals

Examine the model of the d2sp3 orbitals.

1) What atomic orbitals hybridize to give d2sp3 hybrid orbitals?

Answer: ____________________________________________

2) How many d2sp3 orbitals are formed by hybridization?

Answer: _____________________________________________

3) What is the geometry of d2sp3 hybrid orbitals?

Answer: ______________________________________________

a) six bonding pairs:

Construct the molecular model for SF6 by using the 6-hole ball (grey) for S atom. Attach
six links and place a white ball (F atom) at the end of the each link.

b) five bonding pairs and one lone pair:

Construct the molecular model for IF5 by removing one of the white balls from SF6 model.
Note that it does not matter which white ball is removed.

c) four bonding pairs and two lone pairs:

Construct the molecular model for XeF4 by removing one of the white balls from IF5 model.
This time there will be two possible structures for XeF4. Which is the correct structure for
XeF4?

7-10
Complete the table below:

SPECIES LEWIS BOND LONE ARRANGEMENT MOLECULAR 3-D SKETCH POLAR


STRUCTURE PAIRS PAIRS OF ELECTRON GEOMETRY SHOW ANGLES yes/no
PAIRS

SF6

IF5

XeF4

7-11
Multiple bonds
a) two single bonds and one double bond:

Construct the model for CH2O using the 4-hole balls (black) for the carbon and red ball for
oxygen atom. For double bond use two flexible links. (At this point you should remember
that VSEPR model treats a multiple bond as a single source of negative charge).

SPECIES LEWIS BOND LONE ARRANGEMENT MOLECULAR 3-D SKETCH POLAR


STRUCTURE PAIRS PAIRS OF ELECTRON GEOMETRY SHOW YES/NO
PAIRS ANGLES

CH2O

Answer the following questions:


- What is the hybridization of the carbon atom? _________________________
- How many σ and how many π bonds are formed? ______________________

b) one single bond and one triple bond:

Construct the molecular model for HCN by using the 4-hole balls (black) for the carbon
and yellow for the nitrogen atom. For triple bond use three flexible links.

SPECIES LEWIS BOND LONE ARRANGEMENT MOLECULAR 3-D SKETCH POLAR


STRUCTURE PAIRS PAIRS OF ELECTRON GEOMETRY SHOW YES/NO
PAIRS ANGLES

HCN

Answer the following questions:


- What is the hybridization of the carbon atom? _______________________
- How many σ and how many π bonds are formed? ______________________

7-12
Multiple bonds
a) two single bonds and one double bond:

Construct the model for C2H4 using the 4-hole balls (black) for the carbon atoms. For
double bond use two flexible links. (At this point you should remember that VSEPR model
treats a multiple bond as a single source of negative charge).

SPECIES LEWIS BOND LONE ARRANGEMENT MOLECULAR 3-D SKETCH POLAR


STRUCTURE PAIRS PAIRS OF ELECTRON GEOMETRY SHOW YES/NO
PAIRS ANGLES

C2H4

Answer the following questions:


- What is the hybridization of the carbon atom? _________________________
- How many σ and how many π bonds are formed? ______________________

b) one single bond and one triple bond:

Construct the molecular model for C2H2 using the 4-hole balls (black) for the carbon atoms.
For triple bond use three flexible links.

SPECIES LEWIS BOND LONE ARRANGEMENT MOLECULAR 3-D SKETCH POLAR


STRUCTURE PAIRS PAIRS OF ELECTRON GEOMETRY SHOW YES/NO
PAIRS ANGLES

C2H2

Answer the following questions:


- What is the hybridization of the carbon atom? _______________________
- How many σ and how many π bonds are formed? ______________________

7-13
Experiment 8: Colligative Properties

This experiment should be performed and submitted in pairs.

Objective:
To determine the molecular mass of an unknown solute by freezing point depression analysis.

Prelab Assignment:
-Read and understand Experiment 8
-Complete the prelab assignment on the last page of this experiment.

Introduction:

Colligative properties are physical properties of a solution that depend only on the number of
solute particles dissolved and are independent of the nature of the solute particles. For example,
when a solute is added to a solvent to form a solution, the vapour pressure established by the
solvent is reduced. In fact, many of the physical properties of the solution differ from the pure
solvent: the boiling point, freezing point and osmotic pressure are all affected.

In general, when a solution is formed, the osmotic pressure increases, the boiling point increases,
and the freezing point is reduced (depressed), compared to the pure solvent. The magnitudes of
these changes depend on the number of solute particles in solution, and are independent of the
nature of the particles. Thus, these physical properties are colligative properties.

The freezing-point depression that results from solution formation can be predicted as follows:

T T  T  i K m
solvent solution f

where Tsolvent = freezing-point of the pure solvent


Tsolution = freezing point of the solution
  T = the freezing point depression (by convention a positive value)
Kf = the cryoscopic constant, also known as the molal freezing point depression
constant
for the solvent in °C kg / mol
m = molal concentration of the solution (molessolute / kgsolvent)
i = the van’t Hoff factor

The simplest way to determine the freezing point depression experimentally is to prepare cooling
curves for the pure solvent and the solution. To produce a cooling curve, the liquid is first warmed
slightly above its melting point and then cooled gradually while the temperature is recorded at
regular intervals. The temperatures are plotted as a function of time. With sufficiently vigorous
mixing of the liquid as it is cooled, supercooling below the freezing point may occur just before
the liquid crystallizes. With or without supercooling, once crystallization begins, the temperature
will plateau and remain constant as the pure liquid freezes: the temperature at the plateau is the
freezing point of the liquid.

8-1
At its freezing point, a pure liquid maintains a constant temperature as it freezes. However, a
solution may experience a brief plateau, followed by a continuous gradual drop in temperature as
it freezes. This occurs in solutions where, during crystallization, the solute tends to be excluded
from the crystals that form. Thus, as the solvent solidifies, the concentration of solute dissolved in
the remaining liquid solvent rises over time, causing a continuous, gradual reduction in the
freezing point. Note that this gradual reduction in the freezing point typically only becomes
apparent later in the freezing process, after an initial plateau in temperature with respect to time.

For the purpose of determining the freezing point depression, the first plateau of temperature with
respect to time shall be considered the freezing point of the solution. Thus, for this experiment, it
will be unnecessary to cool solutions beyond this plateau and the cooling curves obtained
experimentally will have the general appearance of the curves in Figure 1.

Pure Solvent

Solution
Temperature

FPSolvent

FPSolution

time
Figure 1. Typical cooling curves for a solvent, and a solution
made from that solvent. The freezing point depression is the difference
between the freezing point of the pure solvent and the freezing point
of the solution. Supercooling is the cooling of the solvent or solution
below its freezing point.

8-2
The van’t Hoff Factor

The van’t Hoff factor i is the ratio of the actual concentration of solute particles in solution to the
theoretical concentration of the solute added to the solution based on its mass. i is a property of
the solute, not the solvent, but it’s role in predicting the freezing point depression (and other
colligative properties) is to accurately account for the actual number of solute particles in solution.

actual concentration of solute particles in solution


i=
concentration of the solute based on its mass

Different types of solutes naturally have different values of i.

Non-electrolytes are chemical species that do not appreciably dissociate into ions in solution. For
these species, i = 1, since every mole of the solute dissolved in solution results in a mole of solute
particles in solution. For example, sucrose (C22H12O11), also known as table sugar, contains no
bonds that are appreciably dissociable under typical experimental conditions. Thus, for sucrose, i
is 1.

Electrolytes are chemical species that do dissociate into ions in solution. For strong electrolytes, i
is typically greater than 1, since every mole of the solute added results in two or three or more
moles of free ions in solution. For example, NaCl has a van’t Hoff factor of 2 because each mole
of NaCl added to solution produces 1 mole of Na+ and 1 mole of Cl- free ions in solution. Thus
the ratio of the concentration of solute particles in solution to the ratio of the concentration of
NaCl (aq) is 2.

By the same line of reasoning, we would predict that i is 3 for MgCl2.

It is important to be aware that not all solutes behave as described above. For instance, weak
electrolytes often have van’t Hoff factors less than we would predict because their dissociation
into free ions does not reach completion. For example, in a solution of HF there is less than 2
moles of ions in solution per mole of HF added to the solution. Thus, HF will have a van’t Hoff
factor less than 2. The van’t Hoff factor for an aqueous solution of HF is around 1.03 under
standard conditions.

In other cases, solutes or their dissociation products may ‘pair’ in certain solutions. This occurs in
situations where the interactions between solute particles are stronger than the interactions that
occur between the solute and solvent. ‘Paired’ solutes effectively behave as one particle in
solution, and thus cause the actual number of effective solute particles in solution to be less than
what we would predict based on the mass of the solute added to solution. In such cases, the van’t
Hoff factor can be less than 1.

In this experiment, the freezing point of the solvent t-butanol (CH3)3COH, also known as tert-
butyl alcohol or 2-methyl-2-propanol, will be determined experimentally in Part A.

8-3
In Part B, a solution of an unknown solute that is a non-electrolyte will be prepared using t-
butanol as the solvent. The freezing point of the solution will be depressed (lower than) the pure
solvent. Experimentally determining the magnitude of the freezing point depression will allow the
molal concentration of the solute in solution to be determined, since K f of the solvent t-butanol is
known
(9.1 kg °C / mole). Knowledge of the molal concentration along with the mass of the solute added
to the solution will allow the molecular mass of the unknown solute to be estimated.

Methods:

This experiment will require you to follow instructions carefully and work efficiently.

Begin by assembling the necessary equipment and setting up the freezing point depression
apparatus as follows:

i. Obtain 2 hotplates, 2 digital thermometers, 2 400-mL beakers, a clean and dry large test tube,
and a stand with 2 mini-clamps.

ii. Prepare a hot water bath. Place a 400-mL beaker containing 200 mL of very hot tap water on
the first hotplate. Set the heat setting to 100. Place one thermometer in the hot water bath.

Periodically check the temperature of the hot water bath. The hot water bath should be maintained
at a temperature of 40 to 50°C.

iii. The freezing point apparatus will be set up according to Figure 2. Place the second hotplate on
the base of a ring stand. Position the hotplate and the clamps attached to the ring stand so that the
lower clamp will support your large test tube over the center of the hotplate platform and the
higher clamp will support the temperature sensor (Stainless steel). Attach the temperature sensor
to the higher clamp.

iv. Place a 400-mL beaker, containing 150 mL of very cold tap water in the center of the hotplate
platform. This is the cold water bath. Place one thermometer in the cold water bath. The heat
setting on this hotplate should be set to 0 (heat OFF). Set the stir setting on this hotplate to about
400 rpm. At higher speeds, the stir bar may become unstable.

Periodically check the temperature of the cold water bath. The cold water bath should be
maintained at a temperature around 9-12°C. Add an ice cube to the cold water bath if the
temperature climbs above 13°C.

v. Clamp the large test tube into the lower clamp and lower the clamped large test tube into the
cold water bath as pictures in Figure 2. Do not allow water to enter the large test tube. The large
test tube is an air jacket that will partially insulate your solutions as you measure their freezing
points.

8-4
vi. Your lab instructor will provide you with a clean and dry smaller test tube for this experiment.
DO NOT under any circumstances rinse or otherwise contaminate the small test tube with water
or any other material. The small test tube will be used to contain the compound for which the
freezing point is being determined (Figure 2). Contamination of this test tube or the solutions will
produce inaccurate results or possibly no result at all.

Sensor data cable

Clamp holding
temperature sensor

Clamp holding large


test tube

Large test tube (air jacket)


Temperature sensor
Small test tube containing
sample liquid
Digitalthermometer
Spirit thermometer
in
Sample Liquid water bathbath
in water
Mini Stir Bar Water bath
500 mL beaker

Stir setting display 400 rpm


Hotplate
Heat setting display 0 (Off) STIR HEAT with
magnetic
stir function
Ring stand

Figure 2. Freezing point determination apparatus.

Part A. Determination of the freezing point of t-butanol

i. The small test tube must be clean and completely dry. Place a clean and dry magnetic stir bar
into the tube. Take the mass of the tube containing the magnetic stir bar. Use a tared Erlenmeyer
flask to support the tube while you take its mass. Record the mass of the tube and stir bar.

ii. In the fume hood, add 8 mL of warm t-butanol to the small test tube using the graduated
cylinder provided.

8-5
iii. Carefully lower the small tube containing the stir bar and t-butanol into the hot water bath.
Allow any frozen t-butanol in the tube to melt. Once the t-butanol is completely melted, allow it
to rest in the hot water bath for 2-3 minutes.

iv. The temperature sensor (Stainless steel) will be used only to measure the temperature of your
solutions. Gently clean the tip of the temperature sensor with a dry Kimwipe. Secure the
temperature sensor in the upper clamp of your stand (Figure 2). Do not immerse the tip of the
temperature sensor into the water bath.

v. Setting up Capstone for temperature reading

1. Click on Capstone on the desktop to open the program.

2. From “Tools” palette along left of the window, click on “Hardware


Setup”. A sidebar will pop up with a picture of the interface (Science
Workshop 500).

a. Click on the yellow circle that matches where the


sensor is plugged into the interface (labelled
“Analog Channels” A, B, and C on the physical
interface).

b. A list of possible sensors will appear. Type “t” to


narrow down that list and then select
“Temperature sensor (Stainless steel)”.

c. The “Hardware Setup” picture will now show a green line connecting the interface
to a thermometer icon. Click on “Hardware Setup” again to hide the sidebar.

3. From “Displays” palette along right side of the window, click and drag
“Graph” (top icon) into the center of the window. This will open a graph on the
main tab (“Page #1”).

a. On the y-axis, click on “<Select Measurement>”.


b. Select “Temperature (C)”.
c. By default, the x-axis will automatically update to “Time (s)”.

Save your file! Save experiment as a file with both your and your partner’s name. Make sure you
save the file in the Desktop folder “NYA-Colligative properties-[SEMESTER]”.

8-6
vi. Remove the small test tube containing the t-butanol from the hot water bath. Wipe the outside
of the tube dry with a paper towel, and then gently place the small tube into the large test tube (the
air jacket, Figure 2). Lower the test tubes in the cold water bath. The stir bar should spin
consistently and stably in the bottom of the small test tube, stirring your
t-butanol. If the stir bar does not spin properly, re-adjust your setup so
that the tubes are directly over the center of the hotplate.

vii. Lower the clamp supporting the temperature sensor so that the tip
of the sensor enters the t-butanol inside the small test tube. The tip of the sensor should be
suspended 3 to 5 mm above the stir bar.

viii. On the Capstone window, from the “Controls” palette along the bottom of the window, click
“Record”.

Wait a few seconds to see the plot appear on the graph. Observe the temperature changes and the
evolution of the graph keeping an eye on the t-butanol sample. Notice the supercooling effect on
the temperature curve and the plateauing following after. The t-butanol will turn from liquid to
solid - frozen (it appears as a slosh). After a few seconds of plateauing, click “Stop” on the
“Controls” palette.

Check the value of the temperature on the plateau by using the “Coordinates” tool:
To obtain an accurate reading of the x/y coordinates click on “Add a Coordinates Tool” from
the toolbar above the graph. From the dropdown list, select “Add Coordinates / Delta Tool”. The
“Coordinates” tool will appear on your graph. Drag the “Coordinates” tool to the plateau points
on the graph, and the tool will display the x/y coordinates at that data point. If the temperature
reading is constant, Click “Stop” on the “Controls” palette.

Save your data after each trial.

Part B. Determination of the molecular mass of an unknown solute

i. You will be assigned an unknown solute number by your instructor. Record your solute number
on your report sheet in the space provided.

Solute Mass of solute to add to t-butanol


number
1 0.1 g
2 0.1 g
3 0.2 g
4 0.2 g
5 0.05 g

ii. Remove the temperature sensor from the small test tube and wipe the tip with a Kimwipe.
Remove the small test tube from the air jacket and gently place it in the warm water bath for 5
minutes to melt the t-butanol.

8-7
iii. Remove the small test tube from the warm water bath and wipe the outside of the tube with a
paper towel. The outside of the tube must be completely dry. Take the mass of the tube as was
done in Part A. Use a tared Erlenmeyer flask to support the tube while you take its mass. Record
the mass of the tube containing the stir bar and t-butanol.

iv. Tare the small test tube and the Erlenmeyer flask. Carefully add the mass of your assigned
unknown solute to the t-butanol one drop at a time into the middle of the test tube until the mass
of solute specified in Table 1 has been added. Make sure the solute drops fall into the t-butanol
and do not cling to the sides of the test tube. Record the mass of the solute you added to your t-
butanol. The solute should completely dissolve in the solvent, forming a solution.

v. Place the small test tube containing the solution in the hot water bath. Once all the t-butanol is
melted, allow the test tube to rest in the hot water bath for 2-3 minutes.

vi. Place the small test tube containing your solution in the air jacket. Lower the clean and dry
temperature sensor into the solution. Lower the test tubes in the cold water bath.

Repeat the procedure as in Part A. viii.

When you click “Record” for your second run, the plot from Run #1 will no longer be visible on
the graph, and a new plot for Run #2 will appear. This is normal and does not mean that the data
from Run #1 is lost.

Save your data after each trial.

vii. After stopping the measuring of the temperature, raise the temperature sensor out of the
solution and make sure to wipe the tip with a Kimwipe. Remove the small test tube containing the
solution from the air jacket and place it in the hot water bath to melt the t-butanol solution. Once
the t-butanol solution melted, dispose of the content of the small test tube containing the solution
and the stir bar in the labelled waste beaker in the fume hood. Clean the test tube with soap, brush
and hot water and return it to your instructor.

viii. On the Capstone window:

1. To show both trials on the same plot, click on the “Data


Summary” tool at top of graph (“Page #1”). A dark outline will
appear around it, showing that the button is “pressed in.”

a. Click on the dropdown menu to the right of the Data Summary


tool. Select each of the runs. A checkmark will appear beside
them and they will show up on the same graph.
b. If the Data Summary button is not “pressed in”, clicking on
other runs will switch between which run is shown on the
graph.

8-8
2. Click on “Add a Coordinates Tool” from the toolbar above the graph. From the
dropdown list, select “Add Coordinates / Delta Tool”. The “Coordinates” tool will appear
on your graph. Drag the “Coordinates” tool to any point on the graph, and the tool will
display the x/y coordinates at that data point. Use this tool to obtain the value of the
temperature at the freezing point and record it in your report sheet.

Dispose of waste from the test tube (solution and stir bar) in the labelled waste beaker in the
fume hood.

Wash all equipment with soap, brush and hot water at the sinks at the periphery of the labs.

8-9
8-10
Experiment 8 – Report Sheet Name
______________________________

Part A. Determination of the freezing point of t-butanol.


Dry mass of the small test tube with stir bar
Freezing point of t-butanol

Part B. Determination of the molecular mass of an unknown solute.


Mass of the small test tube containing stir bar and t-butanol
Mass of t-butanol
Unknown solute #
Mass of unknown solute added to the solution
Freezing point of solution
Freezing point depression (T)
Molal concentration of the unknown solute in solution

Molar mass of the unknown solute

Calculations:

If the solute evaporates after weighing, how would your experimentally determined freezing point
of solution differ from what you would have expected had there been no evaporation? Briefly
explain.

8-11
8-12
Experiment 8 – Prelab Assignment Name
________________________

1. Benzene has a freezing point of 6.00 °C, a density of 0.8787 g/mL and a molal freezing point
depression constant (Kf) of 5.12 °C kg/mol. When 0.515 g of an unknown nonelectrolyte solute
was dissolved in 12.3 mL of benzene, the solution that formed was found to have a freezing point
of 4.10 °C. What is the molecular mass of the unknown solute?

2. The following data were obtained in an experiment to determine the freezing point of
cyclohexane (C6H12).

Time (s) T (°C) Time (s) T (°C)


0 12.5 105 5.8
15 11.1 120 5.5
30 9.5 135 5.7
45 8.2 150 5.9
60 7.2 165 5.9
75 6.5 180 5.9
90 6.0 195 5.9

a. From the data above, over what range of time was the cyclohexane supercooled?

b. What is the freezing point of cyclohexane?

8-13
8-14

You might also like