You are on page 1of 67

Laboratory Manual

for

CHEM 0720/0770
UHC
General Chemistry 2

Tamika A. Madison and Peter E. Siska


University of Pittsburgh
Spring 2016 Printing

Ar*

Table of Contents
Introduction to Honors General Chemistry Laboratory II!

iii

Safety in the Laboratory!

iv

Experimentation and Data Taking!

vi

Writing your Laboratory Report!


Experiment #1: Molecular Mass of a Volatile Liquid!

vii
1

Experiment #2: Chemical Kinetics: Reaction Rates, Rate Laws, and Reaction
Mechanisms!
6
Experiment #3: Quantitative Absorption Spectroscopy (Vernier Version)!

12

Experiment #4: The Law of Mass Action: The Ionization of Picric Acid (Vernier
Version)!
18
Experiment #5: pH Measurements I: Acid-Base Equilibria and Titrations (Vernier
Version)!
23
Experiment #6: pH Measurements II: Buffers, Solubility, and the Common Ion
Effect (Vernier Version)!
29
Experiment #7: Redox Activity Series!

34

Experiment #8: Electrochemistry!

38

Experiment #9: Reactions of Iron!

45

Experiment #10: Equilibrium, Free Energy, and Metal - Ligand Equilibrium in a


Complexation Reaction (Vernier Version)!
49
Experiment #11: Aqueous Complexes of Nickel (II): Spectroscopy and Stability
(Vernier Version)!
56

ii

Introduction to Honors General Chemistry Laboratory II


Laboratory work gives you the kind of hands-on and sensory experience that no lecture
or computer-simulated experiment can provide. You can never truly learn chemistry
unless you do it. Observation in science is paramount, accurate observation and
manipulation of experiments requires foresight, care, and skill acquired through
practice. Perhaps more important, good scientific work entails a scrupulous honesty in
the observation and recording of data. If you dont see the result you expect, record
what you do see. The meter reads what it reads, not what it should read. An important
part of each lab consists in discovering any sources of error: if your result was not what
you expected, why? It could be that the error is in your calculations!
Each laboratory experiment is designed to illustrate a topic covered concurrently in the
lecture. In Honors Chemistry, we are moving away from highly prescribed chemistry
experiments common in an entry-level course, not because we think you know
everything already, but because you are clever enough to find out. Not every procedure
or reaction will be spelled out in detail, so you may have to do a bit more preparation
before the lab beyond simply reading the experimental description, and after
experimenting when it is time to analyze your data. In some of the labs, you will be
given the option of striking out on your own, by inventing based on what you have
already seen, further experiments to test some idea or result. In the second term, there
is more emphasis on instrumentation and graphical analysis of results
This manual is locally produced,making it exquisitely sensitive to your needs; it is
revised after every use. We always value, and usually act on, feedback from you and
your laboratory instructor. Please let your instructor know of any errors, inconsistencies
or other problems; we would also like to know what you liked about the labs, so we can
include more of that.
Dr. Peter E. Siska
Spring 2008

iii

Safety in the Laboratory


Playing with chemicals is popularly considered to be a somewhat dangerous activity
practiced by nerds. In fact, with a few simple precautions, it can combine learning with
a lot of fun. Safe laboratory practice is also a matter of your own enlightened selfinterest, and concern for your fellow students.
Wear eye protection whenever you are working at the lab bench. Eyes are the most
vulnerable to permanent damage from chemicals and fumes. Goggles that fit the face
snugly are required. If you ordinarily wear eyeglasses, wear the goggles over your
glasses. Do not wear contact lenses of any type in the lab, even with goggles. You
may remove your goggles only when away from the bench to do calculations or to
consult with your laboratory instructor.
Come to lab appropriately dressed. Do not wear your best clothes. Wear loose clothing
that minimizes areas of exposed skin. Shorts, skirts, bare midriffs, shoes that do not
cover your entire foot, or other overly brief clothing are forbidden. Long hair must be
tied back.
Instead of attempting to list all possible hazards here, safety notes will be given at the
beginning of each experimental write-up, and your instructor will remind you of any
specific precautions when first briefing you. Try to anticipate the hazards as you look
over the procedure before coming to lab. Accidents in the lab are usually due to:

Fire due to volatile solvents and other chemicals.

Spills and splashing of concentrated chemical reagents.

Cuts, usually from broken glassware.

Burns, usually from picking up hot apparatus with bare hands.

Toxic Fumes

Electric shock

Some general safety rules that you must be conscious of during every session are:
1. Do not carry a stock bottle of flammable solvent or reagent to the lab bench.
Take just what you need at the reagent bench when you need it, and return to the
lab bench and use all of it immediately. Flammable reagents must be kept away
from flames or heat.
2. Do not handle or touch any solid or liquid chemical with your bare hands. Clean
up chemical spills at once; ask your instructor for help.

iv

3. If any dangerous reagent spills or splashes on you or your clothing, wash it


immediately with plenty of water. For large scale mishaps, remove contaminated
clothing and make use of the safety shower in the corridor. An eyewash station
is also in the corridor; make sure you know where it is and how to use it.
4. Do not exert great force on glassware; it will break and cut you. Do not clean up
broken glassware with bare hands. Dispose of broken glass only in appropriately
labeled containers.
5. Always handle hot crucibles, beakers, etc. with a dry pad or tongs.
6. Do not hide any accident. Tell your instructor at once and cooperate fully in
making an official report and obtaining follow-up treatment, if needed.
7. Do not smoke, eat, or drink anywhere in the lab. Clean up your work area and
wash your hands after completing the lab.
You will be asked to sign a sheet verifying that you are aware of these and other safety
rules before your first lab session.

Experimentation and Data Taking


All experimental data must be recorded in a laboratory notebook with the facility for
making a removable carbon copy. The recommended notebook is the Hayden-McNeil
50-set notebook, in stock at the Book Center. Each original page of data must be
signed and dated by you, and initialed by your instructor. Carbon copies are to be
turned in to your instructor at the conclusion of each lab session.
Each experiment has been planned to fit into a three-hour session, including
calculations. In order to make this possible, you must read the lab write-up before
coming to lab and prepare one or more pages of your notebook, with your name and the
title of the experiment at the top, and labeled spaces for recording observations and
numerical data such as masses, volumes, temperatures, etc. This preparation is
required before you can begin experimentation. Please observe the following practices:
(a) Recording of data is to be in ink only in your notebook; temporary notes on
scraps of paper are forbidden. If you must change an incorrect datum, cross out
the original entry with a single ink line, and enter the new datum beside it.
(b) Each recorded numerical value must reflect the precision of the measurement by
means of the appropriate number of significant figures.
(c) When weighing out samples precisely, do not waste time trying to adjust the
mass of the sample to exactly the suggested amount. For example, the
experiment might call for you to weigh 0.5 g of calcium metal to 4 decimal
places... Since only one significant figure is specified, any mass from 0.45 g to
0.55 g will do, and the important fact is the precise mass of that sample (e.g.,
0.4623 g, 0.5127 g, etc.) Your instructor would regard a recorded mass of
0.5000 g with suspicion.
In most cases, the success of the experiment hinges on the cleanliness of your
containers and the purity of the reagent chemicals. A contaminated reagent can
invalidate an entire sessions work. Be clean and neat. Take care not to contaminate
the reagent bottles, balances, or other equipment, and recap opened bottles
immediately after use.
If you have made an egregious error that invalidates your results, there will often be
time to repeat at least part of the experiment. Your instructor can advise you as to
which sections should be repeated.

vi

Writing your Laboratory Report


Good laboratory reports are neat, brief, and to-the-point. They should consist of:
(a) Your name, date, title of the experiment, and the name(s) of your lab partner(s).
(b) A neatly organized tabulation of your experimental procedures, data, and
observations.
(c) Any required calculations, including error assessment.
(d) Conclusions
(e) Answers to the questions at the end of the experiment.
A copy of your lab notebook data sheet(s) will already be in the possession of your
laboratory instructor. Your instructor in this course is generally an advanced graduate
student with extensive teaching experience; and is given some freedom to require
anything additional, such as a statement of purpose, etc., and to assign credit for the
components of your report, including your data sheet. The instructor may also include
assessment of your state of preparation for the lab as part of the report grade. Make
note of other requirements below.
NOTES:

vii

Experiment #1: Molecular Mass of a Volatile Liquid


Safety
The samples to be used in todays experiment are organic liquids, whose vapors are
toxic. Avoid inhaling their vapors and keep your apparatus under a fume hood.
Some of the unknowns are highly flammable. Keep all samples and reagent bottles
away from open flames. Take care when handling hot glassware.
Background
In the century or so since the ideal gas law has been considered a reliable tool of the
chemist, its major use has been in the determination of molecular masses. Long
before its universal acceptance however, Jean Baptiste Andre Dumas developed a
method for extending the use of the ideal gas law to substances that are not gasses
at room temperature. Dumas, the most well known French chemist of the early
nineteenth century, exploited a well-established fact. When a liquid is boiled away to
dryness in a small-necked flask, its vapor fills the flask and drives out the air. If the
flask is then sealed, cooled, and weighed, the mass due to the recondensed liquid is
equal to the to the mass of the vapor that occupied the flask when it is hot. Knowing
the temperature and volume of the hot flask, and the external (atmospheric)
pressure, one can use the ideal gas law to find the molecular mass of the liquid. To
maintain a constant high temperature, Dumas employed boiling water, oil, and
molten metal baths. The hotter baths were reserved for less volatile liquids. In
todays experiment, your unknowns will be highly volatile liquids, for which a boiling
water bath is adequate.
Pre-Lab
You should prepare your laboratory notebook before coming to lab using the
guidelines on page xx. Be sure to reserve room for mass and volume readings, the
atmospheric pressure, and any other data. You should also leave room for any
observations. What you wish to record is up to you, but in science it often happens
that seemingly minor details and barely noticeable features are important.
You also need to estimate the volume of liquid to use in this experiment. Assume
that the molecular mass of the liquid is 100 g/mol and the density is 1 g/mL. Make
sure there is enough liquid to fill the flask with vapor at least three times over,
rounding the volume to the nearest mL. This surplus guarantees that all air will be
swept out of the flask when the sample boils away.

Procedure
Make a small square of aluminum foil with a pinhole in the center to cap the mouth
of a 250 mL Erlenmeyer flask. Measure the diameter of the pinhole to the nearest
0.1 mm. Precisely weigh the dry flask and foil cap before setting up the experiment.
Set up a water bath by filling an 800 mL beaker ~2/3 with water. Add boiling stones
to the water and place the beaker on a hot plate inside of a fume hood. Heat the
water to boiling while monitoring the temperature with an alcohol thermometer.
Obtain the letter of your unknown liquid from your instructor. Using thoroughly dry
glassware, measure out the estimated volume of your unknown and pour it into the
flask. Cover the mouth of the flask with the foil, making sure to center the pinhole.
Crimp the foil around the rim of the flask to seal. Clamp the rim (not the neck) of the
flask with a two-jaw clamp and immerse it as fully in your water bath. Fasten the
clamp to a vertical monkey bar inside the fume hood. Take care that water does not
enter the flask or wet the aluminum foil.
Observe the liquid sample in the flask. When it has boiled away entirely, wait 30
seconds, and then transfer the flask to an 800 mL beaker halfway filed with room
temperature water. Keep the flask in the room temperature water for about one
minute, do not remove the foil cap. Remove the flask from the water, unclamp, and
thoroughly dry the outside of the flask. Take care that no water remains in the folds
of the foil cap. Weigh the flask with the cap in place.
Repeat the experiment twice more with the same unknown. After the last run, obtain
the actual volume of the Erlenmeyer flask by using a graduated cylinder to measure
the precise volume of water needed to fill it to the top.
Further Exploration
After carrying out the above experiment and calculation of molecular mass (MM)
three times, you are free to explore further. It would be nice if you could do this
experiment on something known, like good old H2O! Decide whether the same
technique, or some modification of it, is usable for water vapor (Hint: A salt-water
solution boils at a higher temperature than pure water). Try a modified experiment
with water as the volatile liquid, and see how close you come to MM = 18.0 g/mol.

Report
Your data should allow you to calculate MM from the ideal gas law. A correction to
your results for the mass of air excluded from the flask by the equilibrium vapor
pressure of the unknown should be made to improve the accuracy of the results.
This correction is analogous to that which is made to correct for the presence of
water vapor in gas collected over water, using Daltons Law of Partial Pressures. In
the initial weighing, the flask is filled with air, but after vaporization part of the air has
been displaced by the vapor in equilibrium with the liquid. Therefore, the final
weighing should be corrected by adding the mass of air that has been displaced by
the vapor. The vapor pressure at 25C for your unknown can be obtained from your
instructor. The molecular mass of air is taken to be 29.0 g/mol and the total
pressure in the flask can be assumed to be that of the atmosphere (in a future
experiment, you will learn to measure the vapor pressure of a liquid).
Calculate a value of MM from each run, uncorrected for the vapor pressure of the
unknown and average with error. You can calculate the error by first calculating the
standard deviation (sx):

sx =

PN

x)2

i=1 (xi

N represents the number


of trials which is 3 for this experiment. To get the error, you
p
should divide sx by N . Do the same after making the correction for each run.
Questions
1. A student t-test can be used to determine if two sets of data are significantly
different. To perform a students t-test on your corrected (M1) and uncorrected
(M2) sets of molar masses, you will calculate a t value using the equation
r
|x1 x2 |
M 1 M2
tcalculated =
spooled
M1 + M2
Here, x1 and x2 are the averages for the for the corrected and uncorrected sets
of molar masses, M1 and M2 are the number of data points in each set of molar
masses, and spooled is
s
s21 (M1 1) + s22 (M2 1)
spooled =
M1 + M2 2
Note that s1 and s2 are the standard deviations for the corrected and uncorrected
sets of of molar masses. Compare your calculated value for t to the t values in
Table 1 for M1 + M2 - 2 degrees of freedom. If your calculated t value is greater

than the tabulated t value, then your corrected and uncorrected molar masses
are considered be significantly different. Was the vapor pressure correction
significant?
2. A student inadvertently terminated the heating of the flask before all liquid had
boiled away. How would this affect the value of MM obtained? Show your
reasoning using an appropriate equation.
3. In this experiment, you rely on the rapid diffusion of air through the pinhole in
your foil cap. The kinetic molecular theory formula for effusion (not diffusion)
through a hole of area A. The equation:

Z = 14 [X]vA
can provide an upper limit to the diffusion rate. For a pressure behind the hole of
1.00 atm of air and a temperature of 100C, find the rate Z in molecules of air
per second, and the mass flow rate dm/dt in grams of air per second. According
to this result, how long will it take for all of the air in the flask to escape from the
pinhole, assuming a constant rate? What is wrong with this calculation and how
could it be improved?

Explanatory Notes
The early results Dumas obtained with his new method were controversial, and
caused growing doubt over the validity of Avogadros hypothesis, although Dumas
himself firmly believed it. For example, when Dumas experimented with bromine
liquid and iodine crystals, he found MM = 160 and 254 g/mol respectively, while
gravimetric analysis of silver bromide and iodide yielded half these values. As we
now know, this was evidence for the diatomic nature of halogen vapor particles, and
not a counterexample to Avogadros idea. Experiments with sulfur and phosphorus
gave inconsistent results. Later, mass spectrometric analysis of the vapors of these
elements have shown that they consist of several different molecular species, with
the composition depending on T (e.g. in sulfur vapor there are S, S2, S3, and S8
molecules in variable proportions).

Experiment #2: Chemical Kinetics: Reaction Rates, Rate Laws, and


Reaction Mechanisms
Background
The law of mass action, which you will explore later this term in laboratory, had its
origins in observation of reaction rates for forward and reverse reactions. Studies of
the reaction rate of the hydrolysis of the ester ethyl acetate
(1) CH3COOCH2CH3 + H2O CH3COOH + HOCH2CH3
by Berthelot and St. Gilles in 1860 showed that the rate of reaction depended
linearly on the concentrations of ethyl acetate and water. This was later expressed
by vant Hoff in the form of a rate law.
d[CH3 COOCH2 CH3 ]
dt

= k[CH3 COOCH2 CH3 ][H2 O]

where the time derivative expresses the rate of loss of water and k is a rate
constant. The reaction is said to be first order in both ethyl acetate and water, and
second order overall. The order refers to the power of the concentration appearing
in the rate law. Berthelot and St. Gilles also studied the reverse reaction and found
its rate to be first order in both CH3COOH and HOCH2CH3. When equilibrium in an
equimolar system was reached, the products were found to comprise 34% of the
reaction mixture, no matter in which direction the reaction was studied. With the
concept of the rate law, this meant that the reverse reaction was four times faster
than the forward for given concentrations of starting material, and that, to balance
the rates at equilibrium, the products had to be present in lower concentration.
The reaction was found to be greatly accelerated by increasing the temperature of
the reaction mixture. The rate law remained unchanged, implying that the rate
constant k depends strongly on temperature. The form of this dependence was first
proposed by our old ionic friend Svante Arrhenius as

k = Ae

Ea /RT

where A is called the pre-exponential factor, and Ea the activation energy. In


modern times we interpret A as an attempt rate which is a measure of how often the
reagents encounter each other. Ea, which appears in the exponential Boltzmann
factor, can be interpreted as an energy hill that the reagents must climb before they
can succeed in reacting. At higher T, the Boltzmann factor grows, increasing the
fraction of reaction attempts that succeed. The form of the Arrhenius equation leads
to a vant Hoff - like expression for the dependence of k on T.

d ln k
=
d( T1 )

Ea
R

A plot of ln k versus T will be linear be linear with a slope of

Ea
R.

Vant Hoff himself later studied the acid catalysis of Reaction (1), finding that the
overall reaction rate could be enhanced many - fold by adding minute amounts of a
strong acid, but that the composition of the system at equilibrium was unaffected.
This led to speculation concerning the mechanism by which the acid catalyst
performs its magic. From that day, the question of mechanism has dominated
discussions of inorganic, organic, and biochemical reaction rates. A mechanism is
formally as a sequence of elementary steps by which the a reaction is thought to
actually proceed. That is, each step is imagined as an actual occurrence involving
anywhere between one and three molecules. The steps are themselves reactions,
and when summed, they must yield the overall balanced chemical equation. For
example, the mechanism for the acid catalyzed version of Reaction (1) may consist
of two steps, the protonation of the ester followed by the a reaction of the protonated
ester with water.
Step 1:

ester + H+ esterH+

Step 2: esterH+ + H2O acid + alcohol + H+


When these two reacts are added, they give Reaction (1), the H+ catalyst and the
esterH+ reaction intermediate canceling out. Note that the catalyst H+ is not
consumed in the reaction, hence a single proton can catalyze a large batch of
reactions. Often, a catalyst will change the rate law as well as the rate constant.
The rate of a multi - step sequence is greatly determined by the slowest, or rate
limiting step. In the above example, if the protonation step limits the rate, the rate
law will become first order in H+ and zero order in H2O. Kinetic data, such as what
you gather today, can rule out incorrect proposed mechanism. However, mere
agreement between a reaction mechanism and an observed rate law never proves
the correctness of the mechanism; this is a dictum often neglected in the rush to find
explanations for chemical happenings.
In todays experiment, you will examine the kinetics of an aqueous redox redox
reaction, the oxidation of iodide ion by peroxydisulfate.
(2)

2I-(aq) + S2O82-(aq) 2SO4-2-(aq) + I2(aq)

As in most redox reactions, the position of the equilibrium for Reaction (2) lies far to
the right, so that you do not have to worry about the occurrence of the reverse
reaction. The easiest molecule to use as a monitor of the rate is I2, since it forms a
blue complex with starch indicator. Normally you would see a gradual buildup of
blue color as the I2 product accumulates. With a spectrophotometer and fast

electronics, you can monitor the time profile for the reaction ([I2] as a function of
time). In the absence of these tools, you can use another, very fast reaction that
consumes the I2 as a chemical timer or clock. In this case, the reduction of I2 by
thiosulfate (S2O32-), which as you may recall from last term, was the basis of
iodometric redox analysis, serves this purposes nicely.
(3)

I2(aq) + 2S2O32-(aq) 2-(aq) + S4O62-(aq)

Reaction (3) will consume the I2 as fast as it is made by Reaction (2), thereby
preventing the formation of the blue complex. Thus, if you add a very small, known
amount of S2O32- to the reaction mixture, the solution will flash blue after a certain
time when the supply of S2O32- is exhausted. At that time, the moles of I2 produced
by reaction (2) is known to be half that of S2O82- initially present due to the
stoichiometry of of Reaction (3). This gives you a single point on the time profile. By
keeping [S2O32-]0 constant for each trial, you always detect the time interval t
during which a fixed amount of I2 is produced ([I2]). This allows you to find the
[I2 ]
2rate,
t . By keeping [S2O3 ]0 small, you insure the measurement of the rate
near the very beginning of the reaction, the so-called initial rate.
You will first attempt to determine a rate law for Reaction (2), by assuming the
general form

d[I2 ]
= k[I ]x [S2 O82 ]y
dt
and finding the orders x and y by systematically changing first [I-]0 and then [S2O82-]0
and observing their effect on the reaction time and rate. For example, if the reaction
time is halved (or the rate doubles) when you double the [I-]0, x = 1 and the reaction
is first order in I-. These measurements then allow you to determine the rate
constant k, and may help you to distinguish between possible proposed mechanisms
for this reaction. You will also observe the effect of varying the temperature, and
testing a possible catalyst for the reaction.
Procedure
A. Finding the rate law and rate constant
Before you begin, measure and record the ambient room temperature.
The following stock solutions will be available to you: 0.20 M KI, 0.0050 M
NaS2O3 containing 0.4% starch indicator, and 0.10 M K2S2O8. Note that the [KI]
and [K2S2O8] are such that a stoichiometric mixture for Reaction (2) will result
from equal volumes. To insure uniform conditions for all runs, use a constant
total volume of 50.0 mL in a 125 mL or 250 mL Erlenmeyer flask. In addition,
maintain a [S2O32-]0 of 0.0010 M for each run. Use the dilution equation
(M1V1 = M2V2) assuming additive volumes to determine the volume of each
8

solution to add. All solutions should be delivered from burets, and the total
volume of solution should be maintained with boiled distilled water when
necessary. Since you must time the reaction, you should first make up a solution
containing the iodide, thiosulfate, and any water needed, and then add the
peroxydisulfate from a separate beaker or flask. Begin timing as soon as the
peroxydisufate is added, using a magnetic stirrer to insure thorough mixing.
You need a standard or reference reaction mixture, so that other runs, in which
initial conditions are varied, may be compared to it. For this, use a stoichiometric
mixture (equal volumes of KI and K2S2O8) with no dilution by water. Perform
three runs or more until reproducible reaction times are obtained. These data will
also be used to determine the rate constant at ambient temperature.
To determine the rate law, vary [I-]0 and [S2O82-]0 one at a time. In each case, use
a smaller volume of reagent solution and dilution with water to reduce its
concentration. Perform (at least) three runs in which the initial concentration of a
given reagent is varied, down to 1/4 of its reference concentration; that is at least
six runs in all.
B. Temperature dependence of the rate constant
For these runs, use your reference reaction mixture. Before adding the
peroxydisulfate, warm the I-/S2O32- using a hot plate, or cool it using an ice water
bath. Measure reaction times for at least three different non-ambient
temperatures. Avoid boiling or very high temperatures since the I2 - starch
complex becomes unstable and difficult to form. Measure the temperature just
after timing the reaction.
C. Testing a possible catalyst
Certain metal ions, which can act as Lewis acids, often make effective catalysts.
Try a reference reaction mixture with one drop of 0.1 M CuSO4 added just before
adding the peroxydisulfate. Repeat once more to test reproducibility.
D. Further exploration
Upon showing your data tables to your instructor, and analyzing a subset of data
to obtain estimates for x, y, and k, you may explore further. Here are a few
suggestions.
1. Does the catalyst change the rate law? Do a few more runs to examine this
question.
2. Is Cu2+ unique to its catalytic activity (if any)? Try a drop of some 0.1 M
Arrhenius acid, or some other transition metal ion.

3. How good a clock is the thiosulfate reaction? Does the reaction keep good
time? Devise a few runs to test the clock.
4. Use your calculated rate constant to predict the reaction time for a
combination of initial concentrations that you did not measure. Then carry out
the experiment, thereby testing the predictive power of your experimental rate
law.
Report
Report your reaction timing data as a table with columns indicating the run number,
temperature, volumes used, initial concentrations, reaction times, and reaction rates.
Make two plots of log (rate) vs. log [ ]0, one for each set of runs that varied [I-]0 and
[S2O82-]0. What do the slopes of these plots represent? Find values of x and y for
the rate law from these plots, rounding them to the nearest half-integer to reflect
experience with simple reactions such as Reaction (2). Describe the order of
reaction with respect to each reagent and overall. Use your reference run along with
the values of x and y to complete rate constant with proper units. You should
calculate an average k with error.
1

Use your temperature dependence data to prepare a plot of ln k versus T (K)


(an Arrhenius plot) and determine the activation energy Ea from it. Also report the
factor by which the Cu2+ catalyst affects the reaction rate.
Consider the following three possible mechanisms for the uncatalyzed reaction:
I. Step 1: I-(aq) + S2O82-(aq) SO42-(aq) + SO4IStep 2: SO4I-(aq) + I-(aq) I2(aq) + SO42-(aq)

(slow)
(fast)

II. Step 1: 2I-(aq) I22-(aq)


Step 2: I22-(aq) + S2O82-(aq) I2(aq) + 2SO42-(aq)

(slow)
(fast)

III. Step 1: 2I-(aq) I22-(aq)

(fast equilibrium)

Step 2: I22-(aq) + S2O82-(aq) I2(aq) + 2SO42-(aq)

(slow)

When the slow step is Step 1, the observed rate law reflects the bimolecularity of
that step. When the second step is slow, the rate law involves a reactive
intermediate, whose concentration must re-expressed in terms of the first step. For
mechanism II, the equilibrium condition for Step 1 may be used to do this . Write the
rate laws that are consistent with these three mechanisms, and state which of these
(if any) is supported by your results.

10

Questions
1. Reaction (2) is exothermic with H = -82.9 kcal/mol. Combine this with your
measured activation energy to sketch an energy profile for this reaction (a hill
separating the reagent and product enthalpy levels.
2. An old rule of thumb in chemical kinetics states that the rate of reaction doubles
for every 10C rise in temperature. How well does Reaction (2) conform to this
rule. To what value of Ea does the rule of thumb correspond?
3. The equilibrium constant for Reaction (2) is 2.6 1043. On the basis of this,
sketch time profiles of all species in the reaction on a common set of labeled
axes for the reference conditions. Label the equilibrium region. Include beneath
your sketch the relationship among the the rate of loss or gain of the four species
in the reaction. On your sketch, show the location of your clock measurement.
What were the rates for the other species at that time?
4. Identify the reaction intermediates in mechanisms I - III and attempt to draw
structures for them.
Explanatory Notes
Chemists have always been fascinated by reaction intermediates, unstable weird
molecules that do not exist before or after, but only during a chemical reaction. The
role of a catalyst is generally thought to be the creation of a new, more reactive type
of intermediate species. For example, the Cu2+ ion may complex with S2O82-,
creating a new species in which the negative charge has been neutralized and the
SOOS linkage is held in a rigid confirmation, making it more susceptible to attack by
I-.
The exact treatment of a mechanism actually involves the numerical integration of
coupled rate equations, in which the rate laws refer to elementary steps. The right
handed sides of the rate laws will contain both production and loss terms, and all
intermediates are included. This is generally necessary if two steps in the
mechanism occur with similar rates. As an example, in mechanism II, the rate
expression for I22- would be
d[I22 ]
= k1 [I ]2
dt

k2 [I22 ][S2 O82 ]

where the first term on the right is a production term from step one with a rate
constant k1, and the second is a loss du to step 2 with the rate constant with a rate
constant k2. If, as in mechanism II, the intermediate never builds up to an
appreciable concentration, its rate of change can be set equal to zero. This leads to
what is known as the Steady State Approximation, which you will learn more about if
or when you advance in your chemical adventures.

11

Experiment #3: Quantitative Absorption Spectroscopy (Vernier Version)


Background
In the coming weeks, you will often be dealing with colored aqueous solutions. The
color is due to the absorption of a certain range of visible wavelengths () by a
solution molecule or ion. The wavelengths that are not absorbed, usually the
complimentary color to that absorbed, are what you see when you view the solution
against a light background. In solution, absorption lines are not seen; instead a
broad, structureless band of wavelengths is absorbed by the molecule. Absorption
in the visible region of the spectrum is always due to an electronic transition (an
electron hops between orbitals). The breadth of the band aries from simultaneous
excitation of myriad vibrational degrees of freedom of the absorber and fluctuations
in solvation. The vibrational excitation occurs because of changes in the bonding
geometry that accompany the electrons promotion. Most colored solutes also
absorb in the ultraviolet region of the spectrum ( < 400 nm), due to excitation of
higher lying states.
Last term you performed some qualitative color/wavelength solution spectroscopy
using hand-held spectroscopes. The spectroscopes will again be used to get an
overview of the spectrum, particularly with regard to the relationship between
absorption and perceived color. For quantitative measurements, you will employ the
Vernier SpectroVis Plus. Our objectives are twofold: to relate concentrations to the
the extent of light absorption, and to relate the wavelength of maximum absorption to
some aspect of the molecular structure of the absorbing species. The relationship
between absorption and concentration is given by the Beer-Lambert law.

A =

log10

I
I0

= " `c

A is called the absorbance at wavelength , I0 is the intensity of a monochromic


(single ) beam of light in the absence of absorption (no absorber present), I is the
intensity with solution present, is the path length (in cm} that the light must travel
through the solution, and is the molar absorption coefficient. is different for
each molecule, reflecting the probability that an electron will hop when light of light of
the proper is present. The above equation predicts a direct proportionality
between absorbance and concentration. Once for a given molecule is known, the
concentration of unknown can be readily determined.

12

Procedure
A. Collecting Absorption Spectra
You will measure spectra of a selection of compounds in aqueous solution:
CoCl2, NaC6H2O(NO2)3 (sodium picrate), Fe(SCN)Cl2, NiSO4, and the acid-base
indicator bromothymol blue. You will use the hand-held spectroscope (for
qualitative observation) and the SpectroVis Plus units will be used to measure
absorption as a function of wavelength. The picrate ion will be the subject of next
weeks lab and the iron complex and nickel solutions those of future labs.
You will need to prepare your samples in the following manner:
NaC6H2O(NO2)3. Combine 2.0 mL of 1.0 10-3 M picric acid solution with 2.0
mL of 1.0 M NaOH and 6.0 mL distilled water.
Fe3+ and FeSCN2+. You will measure a separate spectrum of the stock
0.10 M Fe(NO3)3 solution. To synthesize a sample of FeSCN2+, combine 5.0
mL of 0.100 M Fe(NO3)3 solution, 1.5 mL of 2.00 10-3 M KSCN solution, 2.5
mL of 1.0 M HNO3, and 1.0 mL of distilled water.
Ni2+ and Ni(NH3)62+. You will measure a separate spectrum of the stock
0.10 M NiSO4 solution. Then, slowly ad 5 M NH3 to the NiSO3 solution until a
color change is observed. Mix the solution thoroughly.
Bromothymol Blue. Obtain 10 mL each of 0.10 M NaOH and 0.10 M HCl.
Add 10 drops of bromothymol blue to each sample. Also, add 10 drops of
bromothymol blue to 10 mL of a clear pH 7 solution. You will measure the
spectrum of all three solutions.
Obtain a SpectroVis and LabQuest Mini unit. Plug the SpectroVis unit into one of
the data ports on the LabQuest Mini. Then, connect the LabQuest Mini to a PC
using the USB adapter provided. Open the LoggerPro software on the PC; you
will use the default file that opens up.
You will need to calibrate the SpectroVis unit before collecting any spectra. You
will use a blank cuvette filled with water for the calibration. Place the cuvette
into the cuvette holder of the SpectroVis unit. Go to the Experiment menu and
select Calibrate. Follow the instructions in the dialog box, then click OK. This
completes the calibration of the SpectroVis unit.
To record a spectrum, use a pipette to add a solution to a clean cuvette and
place it in the cuvette holder. Click on the green Collect button. The absorption
spectrum will be displayed on the screen and will be refreshed every 2 seconds
(this usually does not change the spectrums quality over time). When you are
satisfied with your spectrum, click the red Stop button. You should label the
spectrum with the name of the solution. To do this, go to the Insert menu and
13

select the Text Annotation item. A text box with a line going to the spectrum will
appear. You may adjust the position of the text box and line if you wish. Before
moving on to the next solution, examine the data table on the lefthand side of the
screen and find the wavelength at which the absorbance reaches a maximum
value (max, Amax). Record this data in your lab notebook.
Follow the above procedure to record the spectra for the remaining solutions.
When you click on the Collect button, a dialog box will open up. Make sure that
you select Store Latest Run to keep previous spectra. Your new spectrum will
be plotted on the same screen and two new columns will appear in the data table
for the new spectra.
When you have collected all of your spectra, you may print them. Go to the File
menu and select Print Graph. In the dialog window, do not check the box
labeled Print Visible Spectrum on Wavelength Graphs. Select Landscape
mode so that the spectra fill most of the sheet. You should include your spectra
with your report.
B. The Beer - Lambert Law
To verify the equation in the Background, prepare samples of 0.10 M, 0.05 M,
and 0.02 M CoCl2 by carefully diluting small samples of stock solution with
distilled (not deionized) water. Be sure to mix these solutions thoroughly.
You will need to configure the spectrometer so that you can create a
Beer - Lamberts Law plot. First, save your absorption spectra file and open a
new file. You will then need to collect a spectrum of CoCl2. Insert the most
concentrated of your CoCl2 solutions into a clean cuvette and insert it into the
cuvette holder. Click the Collect button. When your spectrum has stabilized,
click the Stop button.
Now, click on the Configure Spectrometer button, which is two places to the left
of the Collect button. This is an Events with Entry data collection mode. In the
drop-down list at the foot of the box, select Individual Wavelength. You should
select your max for CoCl2 from Part A (or a close to it). Click the appropriate
box in the list or click on the appropriate position on the displayed spectrum.
Then click OK. In the dialog box that opens up, select NO, in order to erase the
displayed spectrum. The screen should now show a table with two columns and
a graph of Absorbance vs. Concentration.
You can now measure the absorbance of each of your solution, beginning with
the least concentrated solution. Transfer your first solution to a clean cuvette and
place it in the cuvette holder. Click the Collect button, then let the signal
stabilize. Then click the Keep button, which appears in the Toolbar. DO NOT
click on STOP!! In the Edit box that opens, type in the molar concentration of
your CoCl2 solution, then click OK. This entry and the absorbance value at your
max should be displayed in the data table and the data point plotted on the
14

graph. Continue with the other samples, pressing Keep for each data point.
When you have collected absorbances for each solution, click the red Stop
button.
At this point, the graph should show the data points, but no lines. If this is not the
case, right-click on the graph and select Graph Options. Make sure that the
Point Protectors box is checked and the Connect Points box is NOT checked.
Click Done. Refer to the Instruction sheet for details.
To draw a best fit line for your plot, click the Linear Fit button (which is four
places to the left of the Collect button) and then click OK. The best straight line
through the data points should now be displayed on the graph and a box should
appear with the slope and intercept.
Calculations and Report
In addition to including your printed spectra, give a neatly arranged table that
includes Amax and max for each sample in Part A. For each Amax, calculate the
percentage of light with wavelength max that is transmitted, the %T, and include it in
your table.

%T = 10

100%

Based on your values of Amax and the known concentrations of each sample,
calculate max for each to add to your table. Briefly discuss your findings as to
relative absorbing power for a given concentration. An acid-base indicator itself is
an acid HA, which exists as HA in acid but as A- in base. Point out which forms you
are observing in the first two bromothymol blue spectra you collected (in acid and in
base). What can you conclude about the presence of HA and A- in the pH 7 buffer?

15

Questions

1. The Beer-Lambert Law results from assuming that a thin band (dx) containing
independent solute molecules (just like an ideal gas) attenuates the light by a
small amount (dI) according to

dI =

I cdx

where is a proportionality constant, dx is a small distance of travel along the


light beam, and I and c are intensity and concentration as defined in the
Background section. Integrate this equation to give a relationship of the same
form as the equation in the Background section. Identify in terms of (the
law breaks down at large c or ).
2. From your experimental max values, find the energy level spacing E (in eV)
between the highest occupied molecular orbital (HOMO) and lowest unoccupied
molecular orbital (LUMO) for each absorber. Tabulate these E alongside the
max values in your report.
3. The picrate ion has the structure

in which the bonds are all conjugated (alternating with single bonds).
Approximating this planar ion by a 2D square box containing the electrons (i.e.,
Particle in a Box), find the box edge length (L) that reproduces the observed max
for the picrate ion. The box energy levels are given by

E nx ny

(n2x + n2y )h2


=
, nx , ny = 1, 2, 3, ...,
8mL2

where h is Plancks constant, m is the mass of an electron, and nx, ny are


quantum numbers that characterize each of the two dimensions. First, make an
energy level diagram by considering the quantum number combinations (nx, ny) =
(1, 1), (1, 2), (2, 1), (2, 2), etc, being careful to show any degeneracies explicitly.
Then, count the number of electrons (twice the number of double bonds in the
above spectrum), and allow them to occupy the levels on your diagram. Make
16

sure that your occupation diagram obeys the Aufbau and Pauli exclusion
principles. The transition you observe occurs between the HOMO and LUMO.
Assign quantum numbers to the transition as (__, __) (__, __). You can now
find L from the energy level difference calculated in Question 2.
4. The transition metal ions, Co2+, Fe3+, and Ni2+ absorb visible light due to an
energy splitting in their d orbitals induced by their surroundings. In which
direction did the splitting change in Ni2+ when NH3 was added? (This observation
will be of interest in the discussion of transition metal chemistry and the
accompanying lab later this term.)

17

Experiment #4: The Law of Mass Action: The Ionization of Picric Acid
(Vernier Version)
Safety
You will be using high concentrations of HCl and NaOH, both of which are highly
destructive to eyes, skins, and clothing. Please take appropriate precautions.
Background
In the 1860s several scientists, including Guldberg and Waage in Sweden, Berthelot
in France, and vant Hoff in the Netherlands, were investigating reactions which do
not go to completion, leaving behind some amount of unreacted starting material.
They found that, a certain ratio (K) of product to reagent concentrations was always
achieved, independent of the initial concentrations of reagents, provided the
temperature was held constant. For example, for an incomplete reaction
aA + bB cC + dD (the lower case letters are stoichiometric coefficients), it was
found that

K=

[C]c [D]d
[A]a [B]b

One could even run the reaction backwards, and still achieve the same ratio. This
became known as the Law of Mass Action. Its connection with thermodynamics
remained obscure until the late 1870s, when J.W. Gibbs proposed the free energy
change (G) as the criterion for the for determining the direction of a chemical
reaction. G = 0 indicated a reaction with no preferred direction, a reaction at
equilibrium, and Gibbs showed that this condition led directly to K. K had also been
found to a strong to be a strong function of T, as revealed mainly in vant Hoffs
work. This was also shown by Gibbs to follow temperature dependence of the
entropic contribution to the free energy.
In todays experiment you will explore the equilibrium properties of one of these
incomplete reactions, the ionization of picric acid (2, 4, 6-trinitrophenol,
HOC6H2(NO2)3 or HPic) to the hydronium ion and picric (Pic-) in aqueous solution,
(1)

HPic(aq) + H2O(l) Pic-(aq) + H3O+(aq)

In the previous experiment, you have measured the absorption spectrum of the
picrate ion, which is yellow in solution; picric acid itself does not absorb in the visible
region of the spectrum. By using the Beer - Lambert Law, and the absorption
coefficient at a fixed visible wavelength, = 440 nm from that experiment, you can
measure the equilibrium concentration of Pic-. Knowing the initial concentration of
undissociated acid [HPic]0 and the stoichiometry of Reaction (1) allows you to find
[H3O+] and [HPic] at equilibrium as well. You could then evaluate the mass action

18

constant, usually called the equilibrium constant K from the mass action
(equilibrium) condition

K=

[H3 O+ ][P ic ]
[HP ic]

In general, this ratio of concentrations is called the reaction quotient Q. To


establish this ratio as a law for this reaction, you can vary the [H3O+] in the solution
by adding large excess of HCl, which can be assumed to be 100% ionized in water
to give H3O+ and Cl-. Then, [H3O+] is determined entirely by the added HCl, and
[Pic-] and [HPic] must readjust themselves to restore equilibrium if K is truly a
constant. (You can qualitatively predict what will happen to the concentrations by
invoking LeChteliers Principle, see Chapter 14 in your textbook.) The temperature
dependence of K can also be assessed by performing the absorption measurements
at different temperatures.
Procedure
You will be supplied with stock solutions of 1.0 10-3 M picric acid, 1.0 M NaOH, and
2.0 M HCl. To insure accurate volume readings, you should use a buret to delver
each solution. You should prepare each solution in a small beaker; each with a total
volume of 20 mL. You will need to determine the volume of picric acid stock solution
needed to make each solution 2.0 10-4 M in picric acid (i.e., [HPic]0 = 2.0 10-4 M);
use the dilution formula (M1V1 = M2V2). Add this predetermined volume of picric
acid, any other stock solution (HCl or NaOH) and dilute to 20 mL with distilled water.
The concentrations of the stock solutions have been carefully chosen to allow all
required solutions to be made up without exceeding the pre-selected total volume.
In the real world, you would be responsible for this condition as well.
Dont worry about slight deviations in your actual delivered volumes from your
calculated volumes, simply record the actual volumes used, and dilute with water to
20 mL. You can later compute the actual concentrations using the dilution formula.
Swirl each solution well to insure complete mixing and attainment of equilibrium.
Extra solution may be required if you elect to do temperature dependent
measurements. Record the laboratory temperature to the nearest 0.1 C using your
alcohol thermometer.
A. Calibration
Obtain a SpectroVis and LabQuest Mini unit. Plug the SpectroVis unit into one of
the data ports on the LabQuest Mini. Then, connect the LabQuest Mini to a PC
using the USB adapter provided. Open the LoggerPro software on the PC; you
will use the default file that opens up.
You will need to calibrate the SpectroVis unit before collecting any spectra. You
will use a blank cuvette filled with distilled water for the calibration. Place the
cuvette into the cuvette holder of the SpectroVis unit. Go to the Experiment
19

menu and select Calibrate. Follow the instructions in the dialog box, then click
OK. This completes the calibration of the SpectroVis unit.
Since each SpectroVis Plus unit is slightly different, you should check your
absorption coefficient measurement from last weeks work. Prepare a solution
with 2.0 10-4 M HPic and 0.20 M NaOH, diluting your total solution volume to
20 mL. Record the actual volume of stock solution used next to your planned
volumes. The excess NaOH insures that the (Bronstead - Lowry) acid base
reaction
(2)

HPic(aq) + OH-(aq) Pic-(aq) + H2O(l)

will be driven to completion, making [Pic-] = [HPic]0. To record a spectrum, use a


pipette to add a solution to a clean cuvette and place it in the cuvette holder.
Click on the green Collect button. . When you are satisfied with your spectrum,
click the red Stop button. You should label the spectrum in a way that
distinguishes it from the others you will collect. To do this, go to the Insert menu
and select the Text Annotation item. A text box with a line going to the spectrum
will appear. You may adjust the position of the text box and line if you wish.
Examine the data table on the lefthand side of the screen and find the maximum
absorbance. Record this value along with the wavelength and a calculated
percent transmission in your lab notebook. Compare this with your results from
last week. Save this solution.
B. Establishing the Mass Action Law
Make up five (or more) solutions, each containing 2.0 10-4 M HPic and various
concentrations of HCl ranging from 0.20 M to 1.0 M; dilute each solution to
20 mL. Again, record the actual volume of each solution next to your planned
volumes. Observe each solution by eye against a white background and rate the
color intensity of the solutions.
Record a spectrum for each solution, making sure to annotate each
appropriately. For each spectrum, find the maximum absorbance at and record
it, along with a calculated percent transmission in your lab notebook. How do the
absorbances compare with your visual observations? After this series of
measurements, remeasure the maximum absorbance for the calibration solution.

20

C. Further Exploration
After completing the above measurements and calculating at least two
experimental K values as described below (to make sure that your experiments
have worked well), you may explore further. Here are three suggestions; feel
free to explore your own ideas, checking them with your lab instructor first.
1. Does your K value determined in Part B depend on temperature? Choose on
solution from Part B and use it to determine the temperature of the
absorbance. Use a dry dropper pipet to add this solution to a clean cuvette.
You will need to adjust the volume of solution in the cuvette so that the bulb of
your thermometer can be inserted in it without causing spillage. Then,
immerse the cuvette in an ice water bath, monitoring its temperature T (C).
Observe the color relative to the parent solution at ambient T. Record a
spectrum, then quickly remeasure T (C). Take the measurement
temperature to be an average of the temperature readings before and after
you recorded the spectrum. Repeat with water baths at different
temperatures, both hot and cold, to give 3 - 4 non - room ambient spectra and
absorbances. Report your results using a table to display K vs. T (C) and
1
T (K). Also calculate ln K, and T (K) , and use Excel to prepare a plot of ln K
1

vs. T (K) . Insert a best fit line through you data. According to the vant Hoff
equation

d ln K
=
d(1/T )

H
R

H
the slope of your best fit line should equal
R , where
H is the heat of Reaction (2). Report a value for H . Note that here you
can obtain a H without doing calorimetry.

2. According to LeChateliers Principle, adding H3O+ to Reaction (1) at


equilibrium will cause it to shift to the left, meaning less ionization and less
color. Explore this by putting a small portion of picric acid stock solution in a
test tube, noting its color, and adding concentrated HCl (12 M) dropwise while
observing. What can you conclude about the absorption spectrum of
undissociated picric acid?
3. All of the above measurements employed the same initial concentration
[HPic]0. How do you predict things would change if [HPic]0 were different?
Devise some experiments to test your predictions, and carry them out, report
numerical measurements and comparisons.

21

Report
In addition to reporting ambient temperature and all absorbance and percent
transmission values, use your absorbance at = 440 nm from Part A to obtain the
product " ` = [PAic ] , and use it to compute and tabulate [Pic-] for each solution in
Part B. Because HCl is completely ionized and [H3O+] >> [HPic] in all cases, you
may set [H3O+] = [HCl] in the K expression for Reaction (1). Deduce what to use for
[HPic]; note that [HPic] [HPic]0, since some has ionized. For each of these
solutions, calculate K, tabulate, and calculate an average and error. Assess how
well the Law of Mass Action holds
Questions
1. Use the K obtained from Part B to calculate the standard free energy change
G at laboratory temperature (See Chapter 17, Equation 17.15). By
considering the extreme possible values of K, or by using
pebbles (differentials), give an error estimate for G. What does your value
imply about the spontaneous direction of Reaction (1).
2. Discuss your results on the concentration dependance of the position of
equilibrium in terms of LeChteliers Principle (See Chapter 14).
Explanatory Notes
The ionization of picric acid was chosen not only because it is a well balanced
reaction, but it is an example of acid ionization, a topic that will soon occupy us in
both lecture and lab. Note that in reactions of type A B + C + ..., if the initial [A] is
known, only one concentration needs to be measured at equilibrium to completely
determine the composition of the system. This is worked out in general by
introducing an extent of reaction variable, an analysis you will be learning to use in
coming labs and in lecture as well.

22

Experiment #5: pH Measurements I:


Acid-Base Equilibria and Titrations
(Vernier Version)
Background
Aqueous solutions of compounds we call acids and bases, and their combinations,
produce a wide range of [H3O+], typically between 1 M and 10-14 M. To express
these concentrations conveniently as nearly-whole numbers, the Danish chemist
Srensen introduced the pH scale in 1909. pH stands for the negative of the power
of the Hydronium ion (H3O+) concentration and is defined as

pH =

log[H3 O+ ]

Later this was generalized to include other solute concentrations, as well as


equilibrium constants. Thus, we write,

pOH =

log10 [OH ]

pCu =

log10 [Cu2+ ]

pKa =

log10 Ka

for example. The acidic and basic ends of the pH scale are delineated by the pH of
pure water (pH = 7.00, [H3O+] = 1 10-7 M ). Acids have pH < 7 ([H3O+] > 1 10-7 M)
and bases have pH > 7 ([H3O+] < 1 10-7 M). In aqueous solution, [H3O+] and [OH-]
are related by the water autoionization equilibrium, leading to

pH + pOH = pKW = 14.00


Measurement of such small concentrations was very tedious before the invention of
the glass electrode in the 1930s. This small, portable electrochemical device made
the measurement of pH rapid and accurate over the entire [H3O+] range.
Technological improvements since then in both the electrode and the measuring
electronics allow the regular use of the device, now known simply as a pH meter, in
introductory laboratory programs, research, and testing laboratories. Today you will
be using the Vernier pH sensor along with the LoggerPro software, which works very
much like a traditional pH meter. The pH sensor, like the glass electrode is
expensive, delicate, and easily damaged by mishandling. Pleas follow the
instruction sheet on page xx when using it. You are urged to read Chapter 19 in
your textbook for a description of how a pH meter works.
Today, you will be using the pH sensor as a tool to probe the equilibrium behavior of
a variety of acids, bases, and salts, and to determine the changes in pH that occur in
the course of acid - base reactions by means of titrations. The pH varies smoothly
during a titration, forming what is known as a pH titration curve, a plot of pH vs.
moles or volume of titrant added. In next weeks experiment, you will test the
23

effectiveness of a buffer in resisting changes in pH and measure the solubility of a


slightly soluble metal hydroxide. Further background reading may be found in
Chapter 16 of your text.
Procedure
A. Set up and Calibration
Obtain a pH sensor and LabQuest Mini unit. Plug the pH sensor into one of the
data ports on the LabQuest Mini. Then, connect the LabQuest Mini to a PC
using the USB adapter provided. Open the LoggerPro software on the PC; you
will use the default file that opens up.
You should carry out an initial two point calibration, first using a pH 7.0 buffer
(yellow solution) and secondly a pH 4.0 buffer (pink solution). Instructions for the
calibration will be provided by your lab instructor. If during the experiment you
suspect that the sensor is not working properly, you can repeat this calibration
procedure.
For Part B, you will read the pH from the display in the bottom lefthand corner of
your screen. For Part C, you will need to open the Acid-Base Titration file
provided by the Vernier Company. In this file, the data collection mode, Events
with Entry has already been fully set up. There will be a table with two columns,
labeled Volume (mL) and pH on the left and a graph on the right with similarly
labeled axes. The pH will still be displayed in the lower lefthand corner.
B. pH of acid, base, and salt solutions
You will be supplied with the following stock solutions (all 0.100 M): HCl,
CH3COOH (acetic acid), NH4Cl, NaCl, CH3COONa (sodium acetate), NHCO3,
Na2CO3, NH3, NaOH. Group these into acidic, neutral, and basic sets. First,
measure the pH of the acidic and neutral solutions, and tap and deionized water
as well. Using the HCl and CH3COOH stock solutions, prepare solutions diluted
to 0.01 M and 0.001 M, and measure the pH of each. Put 1 - 2 drops of universal
indicator into each of your beakers after the measurement and record the
observed color. Next, measure the pH of the basic solutions, again adding
indicator after the measurement. Study the effect of dilution for NH3 and NaOH is
optional.
C. pH Titration Curves
Your instructor will assign you HCl and acetic acid solutions of unknown
concentration (concentrations may range from 0.05 to 0.2M). Make sure you
record the unknown letters in your lab notebook. Recalibrate the pH sensor if
necessary. Titrate the HCl first. Use a 10 mL volumetric pipet to deliver
20.00 mL of your unknown solution into a 150 mL beaker. Place the beaker on a

24

magnetic stirrer unit and insert a magnetic stir bar, adjusting so that you are
stirring at a moderate rate.
Insert the pH sensor, then click on the green Collect button. When the pH
reading on the bottom lefthand corner of the screen is steady, click on the Keep
button. In the edit box that opens, enter 0 mL, then click OK. The data point
will automatically be entered in your data table.
Next, rinse and fill a buret with the standardized 0.10 M NaOH solution to the
0.00 mL mark. Begin to add base, stopping to record the volume (to 0.01 mL)
and pH after every mL or when the pH changes by 0.5 units or more. For each
measurement, click on the Keep button and record the cumulative volume
added. At the first sign of rapid pH change, begin to add base drop-wise,
recording the volume and pH after each drop. After passing through the rapidlychanging region, where the endpoint occurs, you may resume adding larger
amounts of base until you are about 5 mL beyond it. When you are finished with
your titration, click the Stop button. Use your endpoint volume to calculate the
concentration of your HCl. To print your titration curve, go to the File menu and
select Print Graph. Select Landscape mode so that the plot fills most of the
sheet. You should also copy your data table into your lab notebook.
Repeat the above procedure for your acetic acid unknown. You should save your
HCl data file and open a fresh copy of the Acid-Base Titration file before
beginning.
D. Reactive preparation of a precipitate
You may have seen in class (and will now see again for yourself) that when
calcium metal reacts with water, displacing hydrogen, a milky-white Ca(OH)2
precipitate is formed:
(1)

Ca(s) + 2H2O(l) Ca(OH)2(s) + H2(g)

In preparation for next weeks experiment, you are asked to carry out this
reaction and store the resulting solutions and precipitates. Add approximately
200 mL distilled water to a clean 600 mL beaker. Slowly add Ca(s) to the water
and allow the reaction to proceed until you notice cloudiness, indicating the
formation of Ca(OH)2; then add a little more Ca. Stir the solution with a clean
glass stir rod to insure complete reaction. Record your observations on the
reaction and the appearance of the solution. The total mass of Ca should not
exceed 0.5 g or so. Monitor the temperature of the solution, and allow it to cool
somewhat with occasional stirring. Now stir the cooled solution to suspend the
precipitate and immediately pour into a 500 mL plastic bottle for storage. Label
the bottle 0.5 g Ca. Clean and dry the beaker, and carry out the same reaction,
but use roughly 1.0 g Ca. This should result in more precipitate being formed.
Note any visual differences. After cooling, stir and store the solution as before,
labeling the second bottle 1.0 g Ca.
25

E. Further exploration
After completing the above experiments, and show your instructor your data
tables and titration curves, you are free to explore further. Here are a few
suggestions; please check with your instructor before pursuing your own ideas.
1. Bring a sample of one or more liquids, or solids that can be dissolved, from
dorm or home to test pH. Report measured pHs along with your estimate of
the source compound that yields acidity or basicity.
2. Among the exploratory experiments suggested in Experiment #4 was the
temperature dependence of the picric acid equilibrium. Attempt a similar set
of measurements on water ionization or any of the weak acid - base
ionizations. Report a vant Hoff plot and estimated heat of reaction and
compare with literature values. (You have probably measured H for a
reaction that is the reverse of water ionization!)
3. Titration curves are supposed to turn upside down when a weak base is
titrated by a strong acid. Check this out by using your (now known) HCl
solution to titrate a 0.1 M NH3 sample, generate a titration curve, and
compare it to the week acid case, and to equilibrium-based predictions.
Report
For background acid-base equilibrium theory and relationship needed to analyze
your results, see the text (Chapters 16 and 17).
Present your results of Part B in a table showing identity of solute, concentration,
and pH reading. In a separate column, predict the pH of HCl and NaOH solutions
based on known concentrations and assumed 100% ionization. Use the pH value
obtained for CH3COOH and NH3 to calculate their ionization constants, Ka and Kb
respectively and pKa and pKb, comparing them to their literature values. Then use
your experimental K values to predict the pH of each solution and their conjugate
salts CH3COONa and NH4Cl. Enter these predictions next to your experimental
values in the table. From formulas given in the text (Chapter 16), use the pH
measured for NaHCO3 and Na2CO3 to calculate Ka1 and Ka2 for carbonic acid H2CO3
and compare them to literature values.
For part C, preset the pH titration curve data in the form of a tabulation (which can
be reproduced from your LoggerPro table) and copies of your printed titration
curves. Determine the endpoint for each as the point of steepest slope, interpolating
your graph if necessary, and use the volume of base added to calculate the
concentrations of the solutions. Also, present a short table containing the pH at the
start of each titration, at the exact midpoint, and at the endpoint. Using the HCl
molarity you have just determined, predict the pH at the start and midpoint for the
HCl titration and enter these values in a separate column next to the measured
values. Use the measured pH at the endpoint to estimate the water autoionization
26

constant KW. Use the measured pH values for acetic acid and its computed molarity
to determine separate estimates of the equilibrium constant Ka at each of the three
points. Compare these values to each other and to that obtained in Part B. Use
your determinations to calculate the pH at two other points along the acetic acid
titration curve on either side of the midpoint and compare with your measured pH
values.
Questions
1. Use your average value of Ka for acetic acid to compute a value for the standard
free energy change G for the ionization of acetic acid, with error. Is the
reaction spontaneous under standard conditions? Why does it happen under
your conditions?
2. Hesss Law applied to free energies can allow you to predict the equilibrium
constants K and standard free energy change G for a number of acid-base
reactions you studied in the heat of reaction lab from last term. Make such a
prediction for the three acid-base reactions you studied in the heat of reaction
lab,
HCl(aq) + NaOH(aq) NaCl(aq) + H2O(l)
HCl(aq) + NH3(aq) NH4Cl(aq)
NH4Cl(aq) + NaOH(aq) NaCl(aq) + NH3(aq) + H2O(l)
as well as the reaction
CH3COOH(aq) + NH3(aq) CH3COO-(aq) + NH4+(aq)
Compare with literature values. What can you say about the spontaneity of these
reactions under standard conditions? What was assumed about the spontaneity
and extent of reaction in the H lab? Are these assumptions valid in light of your
present results?
3. A pH meeter functions by measuring the free energy difference between a 1.00M
HCl solution behind the glass membrane and the sample solution inside (i.e., it
measures the free energy change of dilution). Compute this G for sample
solutions of pH 1.00, 7.00, and 13.00. What is the proportionality factor relating
pH to G?

27

Explanatory Notes
Membranes like the glass membranes that serve as sensors for a variety of specific
ions are now commercially available. These ion selective electrodes or pIon
electrodes operate on the same principle as the pH electrode, through the exclusive
transmission of ions of a certain type across a membrane. Gas permeable and
biocatalytic membranes also allow electrochemical detection of neutral molecules
ranging from gases such as O2 or CO2 to small proteins (chains of amino acids).

28

Experiment #6: pH Measurements II:


Buffers, Solubility, and the Common Ion Effect
(Vernier Version)
Background
The efforts of Bertholett in ca. 1799 - 1803 (not Berthelot, who came much later)
toward the solubility of salts, using both heat and compounds with elements in
common, gave us our first inkling of the possibility of equilibrium and reversible
reactions.. Today we understand the effect of adding an ion in common with a
reaction product on both buffer solutions and precipitates in terms of its influence on
the equilibrium reaction quotient. In todays world, buffers, solubility, and
precipitation play a role in a wide range of phenomena from water hardness and
pollution, to antacid and other gastrointestinal medication, to clean-up of nuclear
waste.
Last week you analyzed an acetic acid CH3COOH unknown using pH
measurements and a standardized NaOH solution. Now you can use your results to
prepare and test a buffer solution. You also prepared two-phase mixtures containing
a solution and a precipitate by reacting calcium metal with water. This method was
selected (as opposed to mixing a salt solution with a soluble hydroxide) to produce a
system where the only ions present are Ca2+ and OH-. As you will see, the pH of the
supernatant liquid of your equilibrated mixtures in considerably greater than 7,
indicated the presence of OH-(aq), and therefore incomplete precipitation. The two
phase equilibrium involved is
(1)

Ca(OH)2(s) Ca2+(aq) + 2OH-(aq)

If the system is at equilibrium, the solution is saturated with Ca(OH)2, i.e., the free
energy change of reaction (1) under the prevailing system conditions is zero, or the
free energy of the precipitate is balanced by that of the solvated ions. The amount
(molarity) of the product that remains dissolved is called the solubility of the salt,
and may be determined from the solubility product

Ksp = [Ca2+ ][OH ]2


when GT = RTlnKsp for reaction (1). Note that the precipitate itself does not
appear in the Ksp expression since no work is required to take it from in standard
state to the actual condition in the system. This in turn means that the equilibrium
position established in reaction (1) cannot be influenced by the amount of precipitate
present as long as there is some. However, if there is a common ion (Ca2+ or OH-)
already present in appreciable concentration or if such an ion is added to the
equilibrium system, the equilibrium position will change in accord with Le Chateliers
Principle, as quantitatively represented by the Ksp expression, although Ksp remains
constant. If the temperature is changed, however, the value of Ksp itself will alter,

29

depending on the sign of H for reaction (1), and shift the equilibrium. In todays
experiment, you will be able to explore these effects, and learn how to control one of
the more important and practical types of chemical reaction.
Procedure
A. Set up and Calibration
Obtain a pH sensor and LabQuest Mini unit. Plug the pH sensor into one of the
data ports on the LabQuest Mini. Then, connect the LabQuest Mini to a PC
using the USB adapter provided. Open the LoggerPro software on the PC; you
will use the default file that opens up. You will read the pH from the display in the
bottom lefthand corner of your screen.
You should carry out an initial two point calibration, first using a pH 7.0 buffer
(yellow solution) and secondly a pH 4.0 buffer (pink solution). Instructions for the
calibration will be provided by your lab instructor. For Parts C-F, you should
repeat the calibration using the pH 7.0 buffer and a pH 10.0 buffer. If during the
experiment you suspect that the sensor is not working properly, you can repeat
this calibration procedure.
B. Making and testing a buffer solution
Using the same acetic acid unknown you analyzed last week, combine
appropriate volumes of you unknown and the standardized NaOH solution to
make about 40 - 50 mL of 1:1 acetic acid/sodium acetate buffer in a 150 mL
beaker. Making up the solution involves reacting exactly half the acetic acid in
your sample with NaOH, thereby providing a concentration of acetate ion (the
common ion) equal to that of the unreacted acetic acid. Record the precise
volumes and measure the pH of the buffer using your pH sensor. Add 2-3 drops
of universal indicator, noting the color. Test the resistance of the buffer to strong
acid or base by first adding 1.0 mL of 0.1 M NaOH, then adding 2.0 mL of
0.1 M HCl and remeasuring the pH. Save your buffer solution if you wish to
explore it further a la suggestion 3 below.
Now take a volume of distilled water the same as that of the buffer in a second
beaker, add universal indicator, and add 0.001 M HCl dropwise to give a pH
about the same as that of the original buffer. Then repeat the addition of base
and acid as before, noting the pH and color changes.
C. Determination of [OH-] in the supernatant liquid of a precipitation
Take care not to shake or otherwise disturb your equilibrated precipitation
mixtures. Using the 0.5 g Ca mixture, pour off a small portion of the
supernatant solution and measure its pH; then add universal indicator and note
its color. The pH measurement allows you to determine [OH-] and and Ksp.

30

D. The common ion effect


To 50.0 mL of the 0.5 g Ca supernatant solution, add about
0.02 mol CaCl22H2O(s), pre-weighed; observe. Note that based on the solubility
rules, CaCl2 is highly soluble. Stir the solution thoroughly with a stirring rod, and
allow it to return to ambient temperature. Repeat the pH and indicator
observations of Part C for this solution.
E. Varying the amount of precipitate
Repeat Part C with the 1.0 g Ca mixture.
F. The effect of temperature
Using supernatant solution from either mixture, transfer 50.0 mL to a clean
150 mL Erlenmeyer flask. Clamp the neck of the flask and immerse it in a
600 mL beaker of water with a magnetic stir bar. Place the flask and beaker on a
magnetic stirrer and set the stirrer for a moderate speed, Set the heater to
moderate heat (about 3 or so) and heat the assembly, monitoring its temperature.
Every 5C, shut off the stirrer (but not the heat), and measure and record the pH
of the contents of the flask, up to 50C.
G. Further exploration
Upon completing your experiments and calculating at least one Ksp value, you
are free to explore further. As usual, you may pursue your own ideas or follow a
suggestion; check with your instructor before proceeding.
1. Explore the distinction between equilibrium and non-equilibrium conditions.
As you observed last week, the Ca(s) + H2O(l) reaction is exothermic, and
were equilibrium notions to apply to the reaction in progress, heating the
water would slow the net reaction down by encouraging the reverse reaction.
Design a small scale experiment to test this idea in the reaction of Ca with
cold and hot water; report your data and conclusions concerning any
distinction.
2. What should happen if the saturated Ca(OH)2 system is cooled? Try taking
the contents of the reaction beaker of Part D and cooling it while stirring in an
ice bath before letting it settle. Measure the temperature and the pH and
report your conclusions. Are these results consistent with your other
measurements.

31

3. Establish that the solubility of Ca(OH)2 is pH dependent (see Question 2


below). After you are assured that all experiments above are completed to
your satisfaction, thoroughly shake the 1.0 g Ca bottle, pour out about
10 mL into a small beaker, and slowly add the buffer solution from Part A.
Record your observations, and perform any relevant calculations.
Report
For Part B, use the measured pH of your buffer to determine pKa and Ka for acetic
acid from the Henderson-Hasselbalch equation (See Chapter 17), and compare it to
those obtained from last weeks measurements, Parts 5A, 5B, and 5C. Average
your various estimates to give a final experimental value of Ka with error, and
compare it to the literature. Using the H-H equation with your own buffer value of
pKa to predict the new pH as a result of adding the strong acid and base and
compare it with your measurements. Also predict the outcome of the unbuffered
addition of strong acid and base, and compare it with your experiment.
For Parts C, D, and E, group data and calculations from the separate runs together
but compare runs in a table at the end. Compare Ksp from the pH measurements for
all runs. Is Ksp from Part D consistent with Parts C and E? In analyzing your results
from Part D, dont forget to account for the Ca2+ added and that present due to
dissolved Ca(OH)2; otherwise you will miss the point of the exercise! Does the
amount of precipitate affect the position of equilibrium? Find an average Ksp with
error at ambient temperature. Also, give experimental values for the solubility of
Ca(OH)2, i.e., its molarity in a saturated solution at ambient temperature, and the
mass (in g) of Ca(OH)2 that will dissolve in 100 mL of solution. Compare Ksp as a
function of T from Part F, and prepare a vant Hoff plot. Derive values of H from
the slope, and S from G and H, and discuss these quantities in terms of the
interactions and arrangements of the reagent and products of reaction (1). Discuss
the consistency of your results of Parts D, E, and F with Le Chateliers Principle.
Questions
1. What is the reaction quotient for the Ca(s) + 2H2O(l) Ca(OH)2(s) + H2(g)
reaction? Find G and K for this reactions from tables, and predict what you
can about the the state of the system at equilibrium. Use this information along
with your experimental data to predict G and K for the reaction
Ca(s) + 2H2O(l) Ca2+(aq) + 2OH-(aq) + H2(g).
Compare the degree of spontaneity under standard conditions of the reaction
with that given it above. Give a free energy diagram that illustrates the
comparison.
2. The solubility of hydroxides is strongly pH dependent. At what buffered pH would
the solubility of Ca(OH)2, expressed as [Ca2+]eq be 1.0 10-4 M?

32

Explanatory notes
Although Ca(OH)2 precipitates readily, as you have seen, it ranks as one of the more
soluble of the metal hydroxides from Group IIA rightward. Some of the transition
metal hydroxides, such as Fe(OH)3, are so insoluble that the measurement of
OH-(aq) is hopeless, and electrochemical methods must be used. For a fixed ionic
charge (say 2+), the metal ions tend to become better Lewis acids as one moves
from left to right on the Periodic Table, strengthening and lending more covalent
character to the bonding in the salt. Hydroxide precipitates of the rightmost
transition metal ions, such as Ni2+ or Cu2+, can often be taken back from their watery
graves by strong ligands such as ammonia or cyanide to make soluble transition
metal complexes.

33

Experiment #7: Redox Activity Series


Safety
Aside from the use of sulfuric acid as a reagent, the hazards in this experiment are
minimal.
Background
The idea that chemicals have an affinity for each other is a very old alchemical
concept, but it was not until the seventeenth century that Gauber, Boyle, and Newton
(our physicist turned chemist) began to recognize relative affinities in single
displacement reactions, where an element would take the place of another in its
compound. For example, copper metal would replace silver in nitre of silver (silver
nitrate) to make nitre of copper and silver metal. Written in our modern way, this
would be
Cu(s) + 2AgNO3(aq) Cu(NO3)2(aq) + 2Ag(s).
This is a one way street; silver metal will not replace copper in its nitrate. The early
chemists believed this to be due to the greater affinity of Cu for nitrate.
The French apothecary E.F. Geoffroy is credited with assembling the first table of
relative affinities or activities in 1718; he called it a Table des rapports. A half
century later, when the distinction between elements and compounds had become
clearer, the Swedish chemist Torbern Bergman, a contemporary of Cavendish,
devoted his entire career to an obsession with ranking elements and compounds in
order of their activity toward a given test substance. Metal-acid reactions played a
major role in his work and he recognized for the importance of water -- there were
separate wet and dry activity rankings. Electrochemical findings by Davy and
Faraday in the early nineteenth century cast a new light on the problem, suggesting
that the origin of affinity is electrical in nature. The activity series was finally put on a
quantitative basis decades later by Walther Nernst, who measured the electrical
potentials generated between dissimilar metals and their soluble compounds in
voltaic cells, using Arrheniuss ion theory to interpret his results. A complete
description still had to await the discovery of the electron by Thompson, making
possible our present-day picture of a single displacement reaction as an electron
transfer or redox reaction that only occurs in a certain direction. We now recognize
Bergmans distinction between wet and dry activities is due to solvation energy
liberated when an ion dissolves in water.
Our twentieth century understanding of the structure of the atom and the Periodic
Table has finally allowed us to appreciate the origins of the activity series in terms of
atomic properties (principally ionization energy IE), with exceptions to predicted
periodic trends attributed to differences in solvation energies of the ions formed.

34

Electrons always want to roll downhill, and chemists, and now you young uns, can
discover which way is down through clever experimentation.
Procedure
A. Chemical investigation of activity
Your goal will to be to devise a series of reactions with the reagents below that
utilize the minimum number of reactions necessary for the activity ranking of the
six elements: copper, hydrogen, iron, magnesium, sodium, hydrogen, and zinc.
For reasons having to do with surviving in the lab, elemental sodium and
hydrogen will not be available, so your reaction scheme will have to do without
them. Some of the reactions may require a few minutes to show appreciable
product, so you must be patient.
Available Reagents
Elements

Compounds
(aqueous solutions)

copper

copper (II) sulfate

iron

iron (II) sulfate

magnesium

magnesium sulfate

***

sodium sulfate

***

sulfuric acid

zinc

zinc sulfate

For each reaction that is observed to occur, write a balanced net ionic equation.
Then compose an activity series, most active element first, based on your
results. From the activity series, predict the results of a few reactions you did not
have to try, and experimentally test your predictions.
B. Mystery Element
After completing Part A, obtain a sample of an unknown element (call it X) and a
solution of its compound from your instructor. By means of chemical tests,
decide where it fits in your activity series. Speculate on its identity.

35

C. Further exploration
Take a suggestion or strike out on your own
1. In some early studies, the rate at which a metal would evolve hydrogen gas
when reacted with acid was used as a gauge of its activity. In some cases,
this was found to be unreliable. Does it work for our selection of elements?
Set up a gas collection apparatus, and devise experiments to test this idea.
Be quantitative, and report any conclusions.
2. The more active metals will react directly with water (no acid needed), and
water near the boiling point is much more reactive than at room temperature.
Test those elements that are most likely to react with hot water, based on your
activity series. Report results, chemical equations where appropriate, and
conclusions.
3. Set up a voltaic pile by sandwiching a paper towel soaked in sodium sulfate
between two pieces of metal. Use a multimeter to measure any voltage
generated by touching the leads to the metal pieces. Do the pieces have to
be dissimilar metals to get a voltage? How does the sign of the voltage
depend on the way you measure it? When the voltage is measured, a small
voltage passes through the pile; what is carrying the current? Try shorting the
two pieces to each other with another piece of metal; observe. Report
numerical results, and relate the measurements to your activity series.
Report
Clearly indicate your reaction scheme, and the logic behind it, along with a log of
attempted reactions, observations, balanced equations, and conclusion. State your
activity series clearly, along with the location of X in it.
Questions
1. Next to each element as arranged in the activity series you found, tabulate the
ionization energy IE, electron affinity EA, and electronegativity . Which of the
quantities is arranged most nearly in a monatomic sequence that matches your
series? Are there any exceptions? What other energetic feature(s) of the
displacement reactions might be important factor(s)? Speculate on the atomic
properties of X.
2. Choose one of the metal-acid reactions you run and illustrate its mechanism
using Lewis structures for reactants and products, using loops and curly
arrows.

36

3. In last terms experiments, you observed a number of single displacement redox


reactions; list as many as you can. Identify the more active element in each
case, and classify each as wet or dry. Try to incorporate or corroborate these
observations with your present results.
Explanatory Notes
A precise knowledge of the activity series is important in metallurgy and metal
recovery from native materials and ores. The alkalis and alkaline earths, the most
active metals, can only be recovery by electrolysis of molten salts, as discovered by
Sir Humphrey Davy at the turn of the nineteenth century. Early transition metals can
then readily be recovered by the reaction of purified ores or ore extracts directly with
alkalis or alkaline earths, for example the rutile ore extract TiCl4 with Mg(s) directly
by redox replacement to yield titanium metal, useful for lightweight, high strength
tools. Rare earths (lanthanides) such as samarium Sm, used in high field magnets,
are recovered by reactions with their aqueous salts with a sodium-mercury
amalgam. Later transition metals such as iron and nickel are usually obtained by
reduction of their oxides with carbon or hydrogen. Because any element lying to the
left of an element of interest can reduce it to its elemental form, many relatively
unexplored possibilities exist for further development of metal-winning methods.

37

Experiment #8: Electrochemistry


Safety
In the electrolysis experiment of Part A, lethal voltages are used. A good rule to
follow is: plug in last; unplug first. You will also be making a chlorine half-cell;
avoid inhaling the fumes and make use of the fume hoods.
Background
An understanding of redox, or electron transfer, reactions really was born of
electrochemical investigation, beginning with the pile of alternating metal plates
separated by brine-soaked paper of Alessandro Volta (1800). The use of these piles
(now understood to be a series of electrochemical cells, batteries connected in
series) by Sir Humphrey Davy and his student Michael Faraday in the next four
decades, did much to elucidate the electrical nature of redox chemistry. They did so
through their discovery of electrolysis, the decomposition of a sample by the
passage of an electrical current through it. Out of this early electrolysis era came
Faradays Laws of Electrolysis, which can be succinctly stated in modern form as the
triple equality

q = It = nF ,
where q is the total charge (Coulombs C) passed through the sample, I is the
electrical current (Amperes A), t is the time (s) during which the current is passed, n
is the moles of electrons involved in the decomposition, and F is called Faradays
constant, with the modern value 96,485.3 C/mol (or J/V).
Approaching the stability of the products of a redox reaction, or equivalently the
driving force behind redox chemistry, had to await the ionic theory of Svante
Arrhenius (1887). Walther Nernst (1889) immediately used this theory, which
postulated the existence of ions in aqueous solution, to explain the results of his
measurements of the voltages generated by electrochemical cells, and formulated
the Nernst Equation for the prediction of cell voltage,
E =E

0.0592V
log10 Q ,
n

where E is the cell voltage, E is that voltage under standard conditions, n is defined
as above, and Q is what you now know as the reaction quotient, a ratio of the ion
concentrations raised to powers equal to their stoichiometric coefficients. J. Willard
Gibbs later showed that E is proportional to the change in free energy:

G
=
n

FE ;

38

and the Nernst equation follows directly from the free energy relation you should
now be familiar with. These advances gave us a better understanding of how Voltas
piles really worked and how aqueous solutions and fused melts of ionic compounds electrolytes - were able to conduct electricity in the earlier electrolysis studies.
In todays experiments you will explore both electrolysis and the characteristics of
electrochemical cells. By electrolyzing a solution of copper sulfate with appropriate
electrical measurements and weighings, you will be able to determine a value for
Faradays constant, F . By measuring the voltages by a series of cells with one
fixed half cell, you will be able to develop a small table of standard reduction
potentials (E ) for half cells, and to test the predictive ability of the Nernst equation.
Finally, you may briefly investigate the electrochemical of corrosion, the rusting of
iron, and how it might be presented.
Procedure
A. Electrolysis
Using a pair of copper electrodes immersed in 0.5 M CuSO4 solution and a
simple rectified direct current (DC) supply, you will run a timed electrolysis
experiment, using the setup shown in Figure 8.1.

Figure 8.1

39

It is your responsibility to wire the electrolysis circuit, an ammeter (the circled A in


Figure 8.1) in series with the electrolysis cell, using the alligator clips provided.
Please observe safety precautions to avoid a lethal shock. If you are unsure
about your set up, ask! The ammeter will be a multimeter, which you must set to
read DC milliamperes. The electrodes must be firmly attached to the sides of the
beaker so that they do not shift or touch each other during the run.
Prepare the copper strip electrodes by scouring with a pad and soap. Rinse
thoroughly, dry, and weigh each to the nearest 0.0001 g, being careful to handle
it by the edge, away from the part to be immersed (grease from your fingers will
cause electro-deposited metal to adhere poorly). Put some distinguishing mark
on one of the electrodes so that you can identify it later, and indicate it in your
notebook next to the corresponding mass reading. Prepare the cell by first
hooking the electrodes over the sides of the beaker, connecting the external
wiring (power supply unplugged!), and then adding about 40 mL of 0.5 M CuSO4
solution. Make a rough sketch of your set-up in your notebook, noting the
polarity of your leads as indicated on the power supply box, particularly with
respect to the marked electrode.
After checking your set-up for correctness and integrity, plug in the power supply,
and carefully note and record the time. The lightbulb in the DC power supply will
remain lit while the experiment is running. Record the current at the start and at
regular 10 minute intervals throughout the experiment. After 30 - 60 minutes,
stop the experiment by pulling the plug, again recording the time. Carefully
remove each electrode, dip in distilled water, then in acetone. Allow the
electrodes to dry in air, and reweigh. During the course of this experiment, you
should be engaged in setting up electrochemical cells for Part B.

40

B. Cell Potentials
In this experiment, you will make voltage measurements on a related series of
cells set up like that shown in Figure 8.2.

Figure 8.2

Here, the circled V represents the multimeter, now set to read DC volts. The
inverted U-tube is the salt bridge, which you must assemble from a U-tube and
the 1 M KNO3 provided. Plug the ends of the tube firmly with rolled paper towel;
you may have to add an additional small amount of KNO3 solution to refill the
tube after the paper plugs are wetted. Clean all metal electrodes as before,
weighing is not needed here. One of the half-cells to be examined is Cl2, Cl-, for
which an inert carbon (graphite) electrode will be used with a saturated Cl2 in 1 M
Cl- as the electrolyte. Another half-cell requiring an inert electrode is Fe3+, Fe2+.
In this example, it is not possible due to limited solubility to prepare a a solution
that is 1 M in each. In constructing the cells, try to use close to 30 mL of solution
in each half-cell; greatly different volumes will cause siphoning through the
salt bridge.
1. Prepare and measure the standard potentials for the following cells:
ZnZn2+Pb2+Pb
ZnZn2+Cu2+Cu
ZnZn2+ Fe3+, Fe2+ C
ZnZn2+ Cl2, Cl- C
41

The same ZnZn2+ half-cell may be used for all cells. Use separate graphite
rods for the Fe and Cl cells. To avoid asphyxiation, assemble the Cl cell
under the fume hood. The non-Zn end of the salt bridge should be set into
1 M KNO3 between cells.
2. Prepare a cell from two of the four couples, excluding the Zn couple, and
measure the voltage. Before inserting the salt bridge, rinse both ends briefly
in 1 M KNO3.
3. Rebuild either the Zn-Pb or Zn-Cu cell, but reduces one of the metal ion
concentrations by a factor of 100 using 1 M KNO3 as the diluent and measure
the voltage. Rinse the salt before use as in 2.
C. Corrosion protection
Underground iron pipes can be electrochemically protected from rusting by use
of a sacrificial anode, that is by electrically connecting the pipe to a piece of
metal that is more easily oxidized (or difficult to reduce). The first step in the
corrosion of iron is the oxidation Fe Fe2+ + 2e-. The standard reduction
potential for the reverse of this oxidation is E = -0.41 V. Therefore, any metal
whose E lies below FeFe2+ is a possible sacrificial anode; a common choice is
MgMg2+. The typical situation is diagramed in Figure 8.3.

Figure 8.3

The circuit is completed by moist soil, which forms the salt bridge. The iron is
the cathode and the absence of Fe2+ in the soil, water itself will be reduced on
the surface of the pipe, 2H2O + 2e- H2 + OH-, since E for this reduction also
lies above MgMg2+.
To simulate this situation, place a clean iron nail in sand wetted with tap water
containing a few drops of phenolphthalein. By means of alligator clips, connect
the nail through an ammeter (multimeter set to DC milliamps) to a weighed strip
of magnesium, insert the magnesium strip into the sand about 2 cm from the nail,
and record the current. Note any color changes, and where they occur. Then
switch the meter and leads to DC volts and record the voltage.
42

D. Further exploration
In the early days of electrochemistry, setting up cells was normally a trial-anderror procedure. Experiment with different arrangements of electrolytes and
electrodes, in an attempt to answer some questions you may have about the cell
set-ups in Part B. Do you really need a copper electrode when the coper half-cell
is the cathode? Do you really need ZnSO4 electrolyte when the zinc half-cell is
the anode. Do the reduction potentials of the Fe3+, Fe2+ or Cl2, Cl- half-cells
depend on what electrode is used? The possibilities are nearly endless!
Report
For Part A, your notebook data should include all masses as recorded, starting and
ending time, elapsed time, all current readings, the average current, and a sketch of
your set-up. On the sketch, or on a separate page of your report, indicate the
direction of motion of ions in the electrolyte and electrons in the external circuit
based on the polarity of the power supply leads. Identify the anode and cathode,
write the half reactions occurring at each, and on this basis decide whether the
marked electrode should have gained or lost mass. Does this agree with your
observations? Compute the mass loss and gain from your weighings. Either
average these mass differences or choose the one you believe to be more reliable
and calculate Faradays constant F (with proper units) from the mass change, the
average current, and the time using Faradays Laws. Compare with the accepted
value, and discuss possible sources of error.
For Part B1, present your results in the form of a table of columns with the line
notation for each cell at the left, then the measured voltages, and the cathode halfreactions. At the right, tabulate the standard reduction potentials vs. the hydrogen
electrode, by combining your voltages with the accepted value for Zn2+Zn of
-0.763V. Also give literature values next to your own. Comment on sources of error.
For one of your cells, give a fully labeled sketch, showing the direction of motion of
ions in the half-cells and salt bridge, and the electrons in the external circuit. For
Part B2, give the line notation for the cell you selected, written in proper order.
Compute form your own measured voltages of B1 the expected voltage of your cell,
and compare it to the direct measurement. For B3, give the line notation of this nonstandard cell. Compare your measured voltage to that predicted from the Nernst
Equation, using your own measurement of E . Describe the possible sources of any
discrepancy.
For Part C, give the overall cell reaction for the sacrificial cell. Report your current
and voltage measurements and any observed color changes. Based on the mass of
Mg and the current, calculate how long the sacrificial anode would last using
Faradays Laws. Compare your measured voltage to the standard voltage
computed from tables, and discuss any disagreement

43

Questions
1. For the electrolysis of Part A, verify by the use of standard reduction potentials
that the redox processes of lowest free energy deficit -- the ones that will most
likely occur -- are the oxidation and reduction of copper. Ignore any possible
process involving SO22-.
2. For the ZnZn2+ Fe3+, Fe2+ C cell of Part B, verify by writing the Nernst
Equation for the cell that the use of electrolyte 0.1 M in both Fe3+ and Fe2+ will
yield the standard cell potential.
3. Compute E for the oxidation of iron by water reduction from tables. Is this
reaction spontaneous under standard conditions? Write the Nernst Equation for
this reaction and point out some conditions for which the direction of spontaneity
might be reversed.
Explanatory Notes
It is amazing to think about what really is going on in the CuSO4 solution during the
electrolysis of Part A. The ions Cu2+ and SO42-, must carry baggage around with
them, wherever they go, in the form of a shell of solvating water molecules. This
impedes their progress to the electrodes and reduces conductance i.e., the current
that can be passed at a given applied voltage. In addition the ionic atmosphere
due to other ions in solution act as a further impedance. A Cu2+ ion, for example,
tends to be surrounded by SO42- ions (at some distance, of course, with lots of H2O
molecules in between), and must overcome an atmospheric drag from the
Coulomb attractions to the SO42- ions as it makes its way to the cathode. Another
way to look at this effect is that the electric field between the Cu electrodes felt by
Cu2+ ion is modified due to its other ions. It is the net field that provides the driving
force for ionic motion. When Cu2+ finally arrives at the cathode, it finds many other
Cu2+ ions arriving at about the same time, like passengers rushing to board a train,
all dragging baggage (solvent molecules and negative ions) along. This gives rise to
a double layer of ions at the electrode surface, a Cu2+ layer closest to the surface
and a SO42- layer farther off. This double layer is a dynamic one, though as copper
ions are continually being reduced (boarding the train) and replaced by new arrivals.
Water also plays a hidden, but critical role in defining the various half-cell potentials
measured in Part B. Some appreciation for this can be gained by considering the
following sequence
Cu(s) Cu(g) Cu2+(g) Cu2+(aq)
The free energy of vaporization corresponds to a voltage od ~5 V, the ionization
~10 - 20 V, and the hydration ~10 V, making the voltage of the cell
H2H+ Cu2+Cu (the Cu2+/Cu half cell potential) a rather delicate balance among
these factors.
44

Experiment #9: Reactions of Iron


Background
Iron, the fourth most abundant element in the earths crust, and the second most
abundant metal after aluminum, is one of the earliest useful metals to be discovered
by man. The symbol Fe for iron stems from the Latin ferrum, meaning that which
bears or carries, and indeed, true to its symbol, irons main use remains as a load
bearer in buildings, bridges, etc. However, as you may have found in Experiment 7,
Fe is a relatively active metal, readily corroded by moisture and oxygen in the
chemical reaction we call rusting, the oxidation of iron. Rust destroys millions of
dollars worth of structures and equipment each year making an understanding of the
chemistry of iron oxide, and compounds derived from it, of some importance and
interest. In addition, higher living organisms with vascular systems have adopted
iron in its oxidized form as a oxygen/carbon dioxide carrier in the circulating fluid we
call blood. The red color of blood arises directly from the iron ions in it.
In the very first experiment of the first honors lab, you saw a way of recovering iron
from its oxide by reduction with aluminum in the thermite reaction. In todays
experiment you will explore the chemistry of iron in combined form, where it displays
a mix of redox and acid-base properties central to the understanding its chemical
behavior. In its compounds, Fe displays two oxidation states, +2 and +3, and
transforming from one to the other forms a significant part of its chemistry. These
changes are accompanied by color changes in solutions and compounds that form
as a result. Iron ions are also fairly strong Lewis acids, forming precipitates and
complexes with a variety of aqueous anions. Drawing on your growing knowledge of
chemical behavior, you will be asked to identify reaction types and write balanced
chemical equations based on informed observation.
Procedure
1. Obtain a 2 - 3 mmol sample of Fe2O3(s); the precise mass is not important here.
Place the sample in a 50 mL beaker, and add 10 mL concentrated (12 M) HCl,
covering the beaker with a small watch glass. Warm it gently until the oxide
dissolves completely. Record your observations and give balanced molecular
and net ionic equations for the reaction.
2. Place 2 mL portions of your solution into small beakers labeled 1, 2, 3, and 4.
Then, pour the rest of your solution onto a large watch glass to dry.
3. Slowly dilute samples 1 and 2 to 50 mL with deionized water, observing closely.
Set sample 1 aside as a reference solution.
4. To sample 3, add 3 mL of syrupy (14.8 M) phosphoric acid H3PO4, stir, and dilute
to 50 mL. Observe. Assuming that the product is the soluble complex
Fe(H2PO4)3, write balanced molecular and net ionic equations for the reaction.

45

5. Add sample 4 to a 50 mL beaker containing iron nails to a depth of ~4 mm.


Leave it in contact until most of the color is gone, roughly 5 minutes. Then rinse
it into a larger beaker, diluting to about 50 mL. Record your observations, and
write balanced molecular and net ionic equations. Note that the beaker of nails
can be used by several students. If you find them rusty, soak the nails briefly in
concentrated HCl, then rinse with deionized water.
6. Compare samples 2, 3, and 4; note any differences and correlate them with your
conclusions from the above observations and equations.
7. To sample 2, add 6 M NH3 slowly and carefully while stirring, until a permanent
change persists, then add 1 mL more. Repeat for samples 3 and 4. Observe.
Using solubility rules and your knowledge of the contents of each solution, decide
on the identity of the reaction products, and write balanced equations for each
reaction.
8. To sample 4, slowly add 5 mL of bleach (NaOCl) while stirring. Observe. By
comparing with other samples, deduce the products and write balanced
equations.
9. Sample 4 still contains a slight excess of bleach; to remove it, add concentrated
HCl until a change occurs. Carefully smell the solution, first fill your lungs with
air, then waft the beaker fumes toward your nose. Record your observations,
and write balanced chemical equations.
10. Boil sample 4 to remove all traces of gas. After cooling, slowly add 6 M NH3, as
before, compare the result with sample 2, and deduce balanced chemical
equations.
11. Store the watch glass with the original acidified sample in your drawer until next
week, to look for crystals of FeCl36H2O.
Further Exploration
1. Iron (III) most strikingly combines the properties of oxidizing power and Lewis
acidity. If reagents are available, try test tube reactions of portions of sample 1
with KSCN, K4Fe(CN)6, SnCl2, KI, NaBr, Na2CO3, etc. Record your observations,
and where a reaction occurs, give balanced molecular and net ionic equations, if
you can, and classify them.
2. Iron (II) is, on the other hand, a fair reducing agent. Use a small amount of
bleach to make I2(aq) and Br2(aq) form KI and NaBr solutions, respectively, and
try reacting these with iron (II) solution. Report your observations and give
balanced chemical equations if possible.

46

3. Solution conductivity can provide an additional diagnostic for the contents of a


reaction solution. Try repeating a few experiments where you expect big
changes in conductivity, and test your predictions.
4. Use a spectroscope to observe the changes in absorption that occur during the
bleach reaction. Record wavelengths and relative absorption strengths for the
starting solution and at intermediate stages. Does the peak absorption slowly
shift as the color changes, or does it change instantly. If you have done previous
measurements of this type (e.g., with acid-base indicators), compare those
results with these.
Report
Organize your report as a table with columns for reaction number (1 - 11 as above),
sample number, contents of sample, added reagent, observations, and balanced
chemical equations. You can use multiple lines for each table entry to allow room for
full descriptions.
Questions
1. Classify reactions 1 - 11 as redox or acid base; further classify into types of acidbase or redox if possible.
2. During World War II, sailors would keep rust off the decks of their ships with
toothbrushes and a pink goop known as navel jelly, a gelatinous suspension of
phosphoric acid. Based on your chemical experience gained in this lab, suggest
the chemistry behind the rust removal , and describe what the sailors may have
observed as they scrubbed.
3. You may have wondered why the iron solutions are colored and why the color
changes when various reagents are added. The phenomenon of color occurs
frequently in solutions containing transition metal ions with partially filled d subshells. Give electron configurations and orbital occupation diagrams for Fe2+ and
Fe3+.
Explanatory Notes
In a gas phase iron ion -- a rare (flying!) fish, -- all five d orbitals are degenerate, but
in solution, the d orbitals are disturbed by water molecules or other species
swimming nearby, and split up into two or more distinct energy levels. When white
light (e.g., ambient room light) shines on the solution, d electrons in the lower level
will absorb those wavelengths corresponding to the energy level differences, and
jump to the higher level(s). This absorption often occurs in the visible range of
wavelengths (400 - 700 nm) and the color you observe arises from wavelengths not
absorbed. Thus, an orange solution results when blue light is absorbed.

47

As implied above, the Fe ions are not floating about freely in the solution, but are
surrounded by neighboring water molecules or other solute particles. The nearest
neighbors (called ligands) arrange themselves in a regular array around the ion to
form a rather strongly bound complex. The identity of these neighbors determines
the magnitude of the energy splittings among the d orbitals, and hence the color of
the solution. A color change also results when, e.g. Fe2+ is substituted for Fe3+,
while keeping the ligands the same. The color change from orange to yellow,
occurring when you first dilute your Fe3+ sample with water, arises from replacing Clligands by H2O ligands.
FeCl63- + 6H2O Fe(H2O)3+ + 6ClTry to think about other color changes you observed in those terms. What has
happened if the solution turns nearly colorless? Why are the corresponding salt
crystals also colorless? Why do they often contain waters of hydration? We will
discuss such complexes later this term.

48

Experiment #10: Equilibrium, Free Energy, and Metal - Ligand Equilibrium


in a Complexation Reaction
(Vernier Version)
Safety
You will be using 1M HNO3, nitric acid to stabilize the iron complex that is the subject
of study in this experiment. HNO3 is a powerful oxidizing agent that can cause
severe burns. Please observe safety rules when handling it. KSCN potassium
thiocyanate is poisonous if ingested; as always, wash your hands thoroughly after
lab.
Background
Most compounds that produce colored aqueous solutions are either large organic
molecules with extensive, delocalized bonding such as picric acid or bromothymol
blue, or compounds containing a transition metal ion. The latter were first
characterized in 1906 by Alfred Werner and were for many years known as Werner
complexes; now they are simply called transition metal complexes. When a
transition metal ion, such as Fe3+ as in todays study, dissolves in water, it forms
bonds with six neighboring water molecules in a Lewis acid-base adduct written as
Fe(H2O)63+. The metal ion is the Lewis acid and the water molecules, referred to as
ligands, are Lewis bases, donating their electron pairs to the positive metal center.
The understanding of the nature of the bonding, and its relationship to the color of
the complex, had to await the molecular orbital analysis of Harry Gray and others in
the late 1960s. The combination of the simple structure of these complexes, their
visible light absorption, and their lability (reactivity) due to the relative weakness of
the metal-ligand bonds, provides us with a unique opportunity to use thermodynamic
data to probe the structure of the metal ligand interaction.
In todays experiment you will play with a particularly simple complex,
FeSCN(H2O)52+, in which a single thiocyanate ligand SCN-, has displaced one of the
water ligands. This complex has a proper name, thiocyanatopentaaqua iron (III), but
you can refer to it simply as iron thiocyanate. The reaction in which it is made is
called a ligand substitution reaction,
Fe(H2O)63+(aq) + SCN-(aq) FeSCN(H2O)52+(aq) + H2O(l)
The extent of this reaction for various initial concentrations of reagents can be
monitored by the blood red color of the thiocyanate complex, which arises from the
absorption band at the blue end of the spectrum as you may have measured in a
previous experiment. Because water is the solvent, the displaced water ligand does
not appear in the reaction quotient Q, which can therefore be written simply as

49

Q=

[F eSCN 2+ ]
[F e3+ ][SCN ]

where we, as usual, the square brackets denote molar concentrations


As in the picric acid experiment, you can use the Beer-Lambert law to obtain the
[FeSCN2+] from the measured absorbance at max for a solution to which known
amounts of reagents have been added. The stoichiometry of the above reaction can
then be used to obtain [Fe3+] and [SCN-] from the measured [FeSCN2+] as follows.
We define initial reagent concentrations c1 and c2 and an extent of reaction variable
x (see Chapter 14), and make an ICE table
Fe3+(aq)

SCN-(aq)

FeSCN2+

Initial

c1

c2

Change

-x

-x

+x

Equilibrium

c1 - x

c2 -x

For constant solution volumes n c, so concentrations may be substituted for x, so


concentrations can be used directly. When the equilibrium concentrations are
substituted into the reaction quotient, and the condition Qeq = K noted, we get

K=

(c1

x
x)(c2

x)

Since x = [FeSCN2+], the measured complex concentration may be substituted for x


in the above equation and K obtained for each set of initial concentrations. As in the
picric acid experiment, the Law of Mass Action suggests that K should be a constant,
independent of c1 and c2. K yields G at ambient temperature, and a temperature
dependent study will allow you to derive H and S, as well. This information will
be combined with the spectrum to yield estimates of the relative and absolute bond
energies of the Fe3+OH2 and Fe3+NCS- bonds.
Procedure
You will be provided with the following stock solutions: 0.10 M Fe(NO3)3,
4.0 10-3 M Fe(NO3)3, 4.0 10-3 M KSCN, 2.0 10-3 M KSCN, and 1.0 M HNO3.
Solutions will be made up by delivering precisely known volumes of these stock
solutions from burets, achieving a final total volume of 20.0 mL in every case by
topping off with distilled water. Se the example in the picric acid lab (#4) to help with
calculating the volumes to use. Fill a clean, dry cuvette for absorbance
measurements. Each solution should be 0.25 M in HNO3; this stabilizes the complex
against decomposition to Fe(OH)3(s) that occurs in neutral or basic solution. Stir or

50

swirl each solution before filling the cuvette, to insure thorough mixing and
attainment of equilibrium.
A. Calibration
Obtain a SpectroVis and LabQuest Mini unit. Plug the SpectroVis unit into one of
the data ports on the LabQuest Mini. Then, connect the LabQuest Mini to a PC
using the USB adapter provided. Open the LoggerPro software on the PC; you
will use the default file that opens up.
You will need to calibrate the SpectroVis unit before collecting any spectra. You
will use a blank cuvette filled with water for the calibration. Place the cuvette
into the cuvette holder of the SpectroVis unit. Go to the Experiment menu and
select Calibrate. Follow the instructions in the dialog box, then click OK. This
completes the calibration of the SpectroVis unit.
To check your result from Experiment #3, prepare a solution containing
3.0 10-4 M SCN-, 0.050 M Fe3+, 0.25 M HNO3, and distilled water, making sure
that the total volume is 20.0 mL as stated above. Record the precise volumes of
stock solutions used. The large excess of Fe3+ in the solution drives the
equilibrium to the right (LeChteliers Principle), insuring the essentially complete
conversion of SCN- to FeSCN2+, that is [FeSCN2+] [SCN-]0.
To record a spectrum, use a pipette to add a solution to a clean cuvette and
place it in the cuvette holder. Click on the green Collect button. . When you are
satisfied with your spectrum, click the red Stop button. You should label the
spectrum in a way that distinguishes it from the others you will collect. To do this,
go to the Insert menu and select the Text Annotation item. A text box with a line
going to the spectrum will appear. You may adjust the position of the text box
and line if you wish.
Examine the data table on the lefthand side of the screen and find the peak
absorbance. Record this absorbance along with its wavelength in your lab
notebook and use the absorbance to determine "

`=

A
[F eSCN 2+ ] . This will

allow you to the complex ion concentration from the absorbance of an unknown
solution. Compare your absorbances to those you obtained in Experiment #6.
B. Stoichiometry verification and equilibrium constant determination
By using comparable (but small) concentrations of Fe3+ and SCN-, you can
achieve equilibria in which all three species have non-negligible concentrations,
an essential condition for obtaining K. A method often used for this purpose
holds the total moles of reagents constant (c1 + c2 = constant), while varying the
ratio of one reagent to the other. This method has the advantage of allowing you
to verify the stoichiometry of the complex (many complexes involve displacement
of two or more water ligands). We suggest taking c1 + c2 = 2.5 10-3 M, to keep
51

the complex ion concentration comparable to that from your calibration


measurements, while varying the ratio c1/c2 from 1/4 to 4. Make seven or more
solutions in this range and some outside it if you choose. As before, each
solution should also contain the requisite [HNO3].
Record a spectrum for each solution, making sure to annotate each
appropriately. For each spectrum, find the absorbance at = 460 nm and record
it. From these data, the calibration, and the above equation for K, you can
determine a value of K for each solution.
C. Temperature Dependence
Choose one of the solutions from Part B (a solution with c1 c2 is suggested, as
it should have given maximum absorbance), to measure the effect of
temperature. First, measure the temperature of the solution in the cuvette to the
nearest 0.1C by inserting a thermometer into the solution, and verify the
absorbance reading for this sample. Then cool or heat the cuvette in water baths
and measure absorbances at 3-4 non-ambient temperature (see Further
Exploration 1 in Experiment #3 for more details). Note qualitatively whether
complex ion formation is enhanced or not when the solution is heated.
D. Further Exploration
After carrying out the above experiments and calculating at least two K values to
insure consistency, you are free to explore. Please check with your instructor
before proceeding; here are a few suggestions.
1. Why the nitric acid? Try making up the calibration solution from Part A, but
without any HNO3. Record your observations. Now add HNO3 slowly from a
buret, swirling the solution, and noting any changes. Record the volume of
HNO3 added when the solution returns to normal. What is the pH of the
solution at this point? Use your result to estimate the solubility product of
Fe(OH)3.
2. Try a solution with a higher value of c1 + c2 than used in Part B. Is your K
consistent with your other results? What errors might be incurred at these
higher concentrations?
Report
Your report should contain all measurements as originally recorded. In a separate
table, enter your calculations of concentrations (using the actual volumes of stock
solutions), and " ` (Part A) or K (Part B) for each solution. Compute the average
value of K from Part B and calculate the error (standard deviation of the mean).
Present data graphs for the stoichiometric verification (100 - %T vs. c1 from Part B)
and temperature dependence measurements (ln K vs. 1/T, a vant Hoff plot),
choosing the axis units to make sure that the data takes up most of the space of the
52

plot area. From the first graph, verify the stoichiometry; the composition of the
complex ion will be the ratio c1/c2 at the maximum 100 %T. Use the vant Hoff plot to
derive a value for H in kcal/mol and kJ/mol. From your average K, calculate G
at ambient temperature with error estimate, and use it along with H to calculate
S for the reaction, again with error estimate . Compare with literature values.
Point out any assumptions made in the calculations and enumerate major sources of
error.
Questions
1. Solve the K expression in the Background section for x. Calculate x from your
average K for the solutions you need for Part C and compare it to your measured
complex concentration. Explain any discrepancy.
2. G is a criterion for spontaneity of a reaction under standard conditions. For a
reaction in solution this means 1.00 M concentrations of all species, both
reagents and products. If we begin out complexation reaction under these
conditions, would it spontaneously proceed to the right? Use your experimental
K value to find the final concentrations of all species under these conditions, and
thereby verify your qualitative conclusion.
3. Is there a temperature at which the spontaneous direction of the reaction under
standard conditions would be reversed? If so, give an estimate of this
temperature based on your experimental results.
4. Interpret your findings in Part C qualitatively, in terms of LeChteliers principle.
5. H may be taken to represent the difference in bond energies for the iron-water
and iron-thiocyanate bonds. Write an equation that represents this idea, and
draw a three level enthalpy diagram that reflects it. max, on the other hand, in
the MO theory of metal-ligand bonding reflects the metal-ligand bond energy
directly in the case of a single distinct ligand:
D(F e3+

N CS )

hc

max

Use three assumptions to give values (kcal/mol) for both the iron-thiocyanate and
iron-water bond energies.

53

Explanatory Notes
The strong absorption in the blue (450 nm) region of the spectrum by
FeSCN(H2O)52+ owes to the effect of the thiocyanate ligand on the 3d Fe orbitals in a
crystal field description (see Chapter 24 in your textbook). Fe(H2O)63+ is an
octahedral complex, and substitution of H2O by SCN- breaks the octahedral
symmetry. SCN- exerts a stronger influence on the d orbitals than H2O, and hence a
new energy level is created. An energy level diagram is given below.

We have assumed that SCN- is brought in along the z-axis and thus raises the
energy of the dz2 orbital . In Grays MO treatment of transition metal complexes, the
dz2 is the only orbital with symmetry along the metal-ligand axis, and therefore
forms a pair of bonding and anti-bonding MOs with a orbital on the ligand (here a
nitrogen lone pair orbital). The raising of the energy of dz2 corresponds to the antibonding component, and is accompanied by a lowering of the energy of the
occupied ligand donor orbital by an amount equal to the bond energy. If these two
orbitals are the only contributors to the bond, the energy of dz2 is raised by the same
amount, hence E D(M-L) as assumed in Question 5.
In the blue absorption, an electron jumps from one of the lower levels to the dz2.
From there the absorbed quantum is rapidly dissipated as heat. The absorption
band extends over about 100 nm instead of being a sharp line. This occurs because
various vibrations within the complex are excited simultaneously with the electrons
hopping; there are 51 normal mode vibrations in FeSCN(H2O)52+!

54

A further complication which was ignored in the analysis is the reaction of FeSCN2+
with another SCN- to form Fe(SCN)2+. This could be neglected in the present case
because the equilibrium constant for this step is about two orders of magnitude
lower than the first substitution.

55

Experiment #11: Aqueous Complexes of Nickel (II):


Spectroscopy and Stability
(Vernier Version)
Safety
Several of the reagents used in todays experiment are in concentrated from,
including HCl, NH3, and ethylene diamine (H2NCH2CH2NH2); eyes and lungs are at
risk if you are not careful. Please work in the fume hoods and observe all safety
precautions.
Background
The aqueous chemistry of transition metal salts was poorly understood until the early
twentieth century, when Alfred Werner, a chemistry professor at Zurich, and a large
group of his graduate students, established the octahedral complex (ML6) as the
predominate form adopted by ionic transition metal solutes. Here, as we have
already defined in the complex ion equilibrium experiment, M is a metal ion and L
are ligands, which arrange themselves so as to occupy the vertices of an octahedron
(8 sided solid figure). Below is an illustration taken from Werners seminal book on
the subject, New Idea in Inorganic Chemistry, published in 1911. It shows two
possible isomers of an octahedral complex of the form MCl4(NH3)2; the existence of
only two isomers was part of the web of evidence that established the octahedral
geometry.

Later x-ray diffraction experiments on crystals of these complexes confirmed


Werner;s deductions from chemical evidence in every case. Werner was awarded
the Nobel Prize in chemistry in 1913 for this work (as usual, the students did all the
work, but the professor got the glory!)
Since Werners day, we have learned a great deal concerning the relationship
between the striking colors these complexes often display and their stability , or ease
56

of ligand substitution. This in turn how led to the development of a molecular orbital
bonding theory for these complexes known as ligand field theory that links these two
features in a consistent way.
In todays experiment, you will briefly explore the colors, spectra, and reactions of a
series of Ni2+ complexes, and thereby establish a correlation between transition
wavelength (or energy), and reactivity.
Procedure
The ligands to be examined in the experiment will be Cl-, H2O, NH3, and
H2NCH2CH2NH2 (ethylene diamine, abbreviated en). The en is a bidentate (two
toothed) ligand, which occupies two of the six octahedral bonding sites around Ni2+;
the formula of a fully formed Ni-en complex is then Ni(en)32+.
You will be using the SpectroVis units to measure the absorption spectra of your
complexes. After, obtaining this unit along with a LabQuest mini and making the
necessary connections to your laptop, you should calibrate the spectrometer using
distilled water. Make sure that you label each spectrum on the screen and print the
graph when you are finished.
A. Synthesis and Spectra
You will be supplied with solid NiSO46H2O, 12 M HCl, 5 M NH3, and 4% by
volume (0.60 M) en. Used distilled water when needed for all syntheses and
reactions. For each complex you make, estimate the wavelength (max) and
energy of maximum absorbance by first locating the complex on the eyeball
spectroscopy chart at the end of this experiment, and then measuring the
absorption spectrum using the SpectroVis unit. Set the solutions beakers on a
white sheet to enhance visual comparison.
To make the Cl- complex, weigh out about 0.005 moles of nickel (II) sulfate
hexahydrate into a 50 mL beaker, and carefully, under a fume hood, add about
0.1 moles of HCl. Stir with a stirring rod, or with a magnetic stirrer, until the
solution phase shows intense color. Record your observations, and then,
cautiously transferring the clear solution into a cuvette, measure the spectrum.
Pour the solution back into the beaker and save for Part B.
The H2O complex may be made by dissolving the solid in water instead of HCl;
make about 20 mL of 0.5 M NiSO4. Record your observations and measure the
spectrum as above. Then, by diluting a portion of this solution (saving the rest
for Part B), make about 50 mL of 0.1 M NiSO4, and divide this into 5 equal
portions in five 50 mL beakers.
For the NH3 complex, add about a two-fold excess of NH3 with stirring to one of
the five portions. Record your observations during and after the addition, and
record the spectrum. Save this solution for Part B.
57

For the en ligand, you can observe the stepwise formation of Ni(en)2+, Ni(en)22+,
and Ni(en)32+. To one of three 50 mL beakers, add dropwise with stirring a few
drops less than a stoichiometric amount of en solution to make the 1:1 complex,
observe, record the spectrum, and save. To the second beaker, add enough en
for (nearly) a 1:2 complex, and to the third enough to form a 1:3 complex, again
viewing, recording, and saving each. Pause to look over your color and
wavelength data, noting any trends you find.
B. Ligand Substitution reactions
Transfer the contents of the beaker containing the Cl- complex into 250 mL
beaker, including any undissolved solid. Now slowly and cautiously add water
under a fume hood, while stirring and observing, until the color stops changing
(aside from fading due to dilution). Record the volume of water used. Compare
the resulting solution to those saved from Part A, and describe what happened
with a balanced ligand substitution reaction. Transfer 10 mL of this solution to a
150 mL beaker. To this solution add NH3 dropwise while stirring and observing
until the color again stops changing. Again record volume, compare with Part A,
and write a balanced chemical equation describing the reaction. Finally, add en
solution dropwise until reaction is complete, and again record, compare, and
write a balanced chemical equation. Pause to consider that the free energy
changes for each of these reactions had to be negative, and the implications for
the relative stabilities of the various complexes. Compare the order of stability
with any trends from Part A.
C. Further Exploration
Free energy is of course concentration dependent, and it is possible that you
could render the reverse of one or more of the reactions of Part B spontaneous
by changing the concentrations. Devise an experiment to test this idea. Other
ligand solutions may also be available for making different complexes than those
observed here.

58

Report
For Part A, construct a table showing the formula of the synthesized complex ion, its
proper name, masses and/or volumes used, color, estimated wavelength (max) from
hc
eyeball chart, max from your measured spectra, energy of transition E = max
(eV
and kcal/mol), and any pertinent observations. Some of your complexes of Part A
were synthesized by ligand; list these, and give the reactions by which they were
made. Draw structures for each of the six complexes you have made; for the en
complexes show any possible isomers.
For Part B, list the color change observed, identification of the reaction product, and
balanced chemical equation for each reaction. From the directionality of the
reactions, rank the complexes in order of increasing stability, i.e., least stable first.
Under your ranking, indicate the transition energies in eV for each, noting any
correlation. You have measured the relative affinity of these ligands for Ni2+, but the
ranking of ligands is nearly independent of metal ion and is known as the
spectrochemical series.
Connect the results for Part A and B by constructing a series of abbreviated ligand
field energy level diagrams showing the d orbitals of Ni2+, the ligand orbitals, and
the d bonding, the (n) non-bonding, and d* anti-bonding levels for the four
complexes. Since you have measured the transition energy E o, you can
construct quantitative diagrams , assuming the energy lowering of the d in
magnitude to the energy raising of the d*. Indicate the electron occupations of
these levels, and show the absorption transition with a heavy arrow. Briefly discuss
your reactivity results of Part B using these diagrams, assuming that G is
dominated by M-L bond energy differences, and estimate G and K for each
reaction on this basis. Are there any reactions for which you might not be able to
neglect the energy term?

59

The Eyeball Spectroscopy Table for Broadband Absorption


Color You See

Color Absorbed

Approx.
(nm)

Energy Level Spacing (E)


(cm-1)

(eV)

Colorless

Ultraviolet

< 400

> 25,000

> 3.10

Lemon yellow

Violet

410

24,400

3.03

Yellow

Indigo

430

23,200

2.88

Orange

Blue

480

20,800

2.58

Red

Blue-green

500

20,000

2.48

Purple

Green

530

18,900

2.34

Violet

Lemon yellow

560

17,900

2.22

Indigo

Yellow

580

17,300

2.14

Blue

Orange

610

16,400

2.03

Blue-green

Red

680

14,700

1.82

Green

Purple-red

720

13,900

1.72

Green-yellow

Purple

750

13,300

1.65

Colorless

Infrared

> 750

< 13,300

< 1.65

60

You might also like