You are on page 1of 24

Module 2: Quantitative Analysis: Absolute Methods

LESSON 1.0: GRAVIMETRIC ANALYSIS


Gravimetric methods of analysis are used where weights of reactants and products
of chemical reactions are reproducible, stable and reflect the presence of
constituents which are important in the establishment of identity. 

LESSON 1.1: GRAVIMETRIC ANALYSIS:

1. Using Mass as an Analytical Signal


Suppose you are to determine the total suspended solids in the water released by a
sewage-treatment facility. Suspended solids are just that—solid matter that has yet
to settle out of its solution matrix. The analysis is easy. After collecting a sample,
you pass it through a pre-weighed filter that retains the suspended solids and dry
the filter and solids to remove any residual moisture. The mass of suspended solids
is the difference between the filter’s final mass and its original mass. We call this
a direct analysis because the analyte—the suspended solids in this example—is the
species that is weighed.

What if our analyte is an aqueous ion, such as Pb 2+? Because the analyte is not a
solid, we cannot isolate it by filtration. We can still measure the analyte’s mass
directly if we first convert it into a solid form. If we suspend a pair of Pt electrodes in
the sample and apply a sufficiently positive potential between them for a long
enough time, we can force the following reaction to completion.

Oxidizing Pb2+ deposits PbO2 on the Pt anode. If we weigh the anode before and
after applying the potential, the change in its mass gives the mass of PbO 2 and,
from the reaction’s stoichiometry, the amount of Pb 2+ in the sample. This is a direct
analysis because PbO2 contains the analyte.
Sometimes it is easier to remove the analyte and let a change in mass serve as the
analytical signal. Suppose you need to determine a food’s moisture content. One
approach is to heat a sample of the food to a temperature that vaporizes the water,
capturing it in a preweighed absorbent trap. The change in the absorbent’s mass
provides a direct determination of the amount of water in the sample. An easier
approach is to weigh the sample of food before and after heating, using the change
in its mass as an indication of the amount of water originally present. We call this
an indirect analysis because we determine the analyte using a signal that is
proportional its disappearance.

The indirect determination of a sample’s moisture content is done by difference. The


sample’s initial mass includes the water, but its final mass does not. We can also
determine an analyte indirectly without its ever being weighed. For example,
phosphite, PO33–, reduces Hg2+ to Hg22+, which in the presence of Cl– precipitates as
Hg2Cl2.
If we add HgCl2 in excess, each mole of PO 33– produces one mole of Hg2Cl2. The
precipitate’s mass, therefore, provides an indirect measurement of the amount of
PO33– in the original sample.

2. Conservation of Mass
An accurate gravimetric analysis requires that the analytical signal—whether it is a
mass or a change in mass—be proportional to the amount of analyte in our sample.
For all gravimetric methods this proportionality involves a conservation of mass. If
the method relies on one or more chemical reactions, then the stoichiometry of the
reactions must be known. Thus, for the analysis of PO 33– described earlier, we know
that each mole of Hg2Cl2 corresponds to a mole of PO33– in our sample. If we
remove the analyte from its matrix, then the separation must be selective for the
analyte. When determining the moisture content in bread, for example, we know that
the mass of H2O in the bread is the difference between the sample’s final mass and
its initial mass.

3. Why Gravimetry is Important


Except for particulate gravimetry, which is the most trivial form of gravimetry, you
probably will not use gravimetry after you complete this course. Why, then, is
familiarity with gravimetry still important? The answer is that gravimetry is one of
only a small number of definitive techniques whose measurements require only
base SI units, such as mass or the mole, and defined constants, such as
Avogadro’s number and the mass of 12C. Ultimately, we must be able to trace the
result of an analysis to a definitive technique, such as gravimetry, that we can relate
to fundamental physical properties. Although most analysts never use gravimetry to
validate their results, they often verifying an analytical method by analyzing a
standard reference material whose composition is traceable to a definitive
technique. 

LESSON 1.2: TYPES OF GRAVIMETRIC ANALYSIS

1. Precipitation methods
Many metallic elements in their ionic forms react with negative counter ions to
produce stable precipitates. Silver ions form stable and highly insoluble salts with
chloride, bromide, and iodide. Calcium precipitates quantitatively with oxalate and
can be measured reproducibly at any of three temperature dependent plateaus as
the oxalate, the carbonate, and the oxide. Barium precipitates quantitatively as the
sulfate. The reactions often follow the same patterns:

Positive and negative ions in an aqueous solution, otherwise soluble with the
counter ions in their environment, produce highly insoluble precipitates with certain
added reagents.
2. Volatilization methods
An interesting volatilization method which is entirely gravimetric is the one shown by
the equations below.

The analyte can be bicarbonate as shown or a mixture of carbonate and


bicarbonate. The total amount of carbonate in whatever form is found by placing the
analyte in a solution containing an excess of H 2SO4. This solution is in a flask
connected to incoming nitrogen gas gently bubbled through the solution and an exit
tube first to a drying agent to absorb aerosolized water and water vapor and then to
a mixture of NaOH and drying agent to absorb the CO 2 and

3. Electrogravimetric methods
In electrogravimetry, we deposit the analyte as a solid film an electrode in an
electrochemical cell. The deposition as PbO 2 at a Pt anode is one example of
electrogravimetry. The reduction of Cu 2+ to Cu at a Pt cathode is another example of
electrogravimetry.

LESSON 1.3: CONSIDERATIONS FOR THE ISOLATION OF PRECIPITATES


Precipitates ought to be easy to wash free of contaminants without loss of the
precipitate either in solution or through the filter. The particle size of the precipitate
ought to be large enough not to escape through the filter pores. That the precipitate
has a low solubility is paramount. The precipitate ought not to react with the
atmosphere and it must have a known composition which remains stable after
ignition.

Substances of low solubility have the nasty habit of forming colloidal suspensions.
Colloidal particles have diameters from 10 -7 cm to 10-4 cm. That is on the order of
from 10 atomic diameters to 10,000 atomic diameters. Particles in this size range
are still sufficiently jostled about by thermal molecular motion to remain in
suspension. Where they are the result of a process of precipitation brought about by
the addition of ionic species, the particles are surrounded by the excess ionic
species. If Ba2+ is added in excess to SO42- the BaSO4precipitate which is formed is
considered to be surrounded by Ba 2+ ions. If the opposite procedure were being
followed, the precipitate would be surrounded by SO 42- ions. That these particles all
have like charge and therefore repel each other suggests that your technique must
favor the formation of large rather than small precipitate particles and to offer ways
of encouraging the coagulation of particles after they have formed.

This can be done by carrying out the precipitation at a temperature close to the
boiling point of water, in a dilute solution of your analyte and with constant stirring
for the reasons given below. Although analytical chemists still have some
disagreement as to the mechanism of precipitation, there is wide agreement that a
quantity called the relative supersaturation affects the particle size. Relative
supersaturation is given as
where Q is the instantaneous concentration of the species added to effect
precipitation and S is the equilibrium solubility of the substance which precipitates.
Particle size seems to be inversely proportional to Relative Supersaturation because
a high concentration of added reagent increases the probability that oppositely
charged ions will begin the precipitation process at late as well as early stages of
the addition and the resulting particles will be on the order of atomic dimensions,
whereas the maintenance of a value of Q just slightly above S lowers that probability
but offers in any case a layer of the added reagent ions around existing particles for
their further growth.

LESSON 1.4: THE ELECTRIC DOUBLE LAYER


If a particle of precipitate is surrounded by the ion in excess, say Ba 2+ in the case of
the determination of SO42-, any negative ions in the immediate surroundings will be
attracted to that primary positive layer. In the case of the addition of a BaCl 2 solution
to a Na2SO4 solution, the ions Cl- and SO42- are available. As the sulfate is used up
in the precipitation process it is the chloride which is left, and which forms the
second layer. Thus, we have an electric double layer, made up first of barium ions
then of chloride ions. This double layer keeps the colloidal precipitate particles from
coming into contact with each other for further coagulation.

There are two ways


to bring the particles
closer together and
to increase the
probability of
coagulation: (1)
heating increases
overall thermal
motion, affecting both
the mobility of
adsorbed ions and of
the colloidal
precipitate particles
themselves. The
summary effect is
that there are
collisions of particles
which result in the
increase in particle
size due to increased
coagulation; (2)
increasing the
electrolyte
concentration of the
solution, for reasons not entirely clear, results in a decrease in the mean radius of
the electric double layer and encourages further coagulation. Carrying out both of
these operations results in digestion of the precipitate, an unfortunate term because
biological digestion usually refers to the dissolving of food and absorption at the
molecular level through the wall of the intestine.

 Digestion in quantitative analysis refers to the coagulation of a precipitate into a


filterable form. Unfortunately, after successful digestion, some of the primary electric
layer is made up of Na+ ions which must be washed away ultimately for quantitative
results to be achieved. The Ba 2+ ions as well will give a positive error if not removed
and end up being dried with the precipitate as excess BaCl 2. Many coagulated
precipitates do not respond well to washing with distilled water because as the
second electric layer is removed (excess Cl - for example) the first remains on all
particles with an electric charge of the same sign. The result is that there is a return
to the repulsive state and an effective increase in the radius of the particles which
then begin once again to separate as colloidal particles.

The process is called peptization and is to be avoided if some of the precipitate is


not to be lost. One way around this for many precipitates is to encourage digestion
by heating and also by increasing the electrolyte concentration by washing with a
reagent which will go off as a gas during the drying process. Dilute nitric acid,
HNO3 , is effective for washing excess ions from AgCl. In choosing such a wash, it is
imperative that the procedure has been carried out and has been shown to yield
reproducible, quantitative results. Unexpected side reactions, complex formation,
and changes in solubility with added reagents are sufficiently unpredictable to make
intuition in the absence of experience unacceptable.

LESSON 1.5: OTHER FACTORS WHICH CAN AFFECT IN THE ISOLATION OF


PRECIPITATE
During the precipitation procedure a number of other problems can arise to give
erroneous positive or negative results. Among these are surface adsorption, mixed
crystal formation, occlusion, and mechanical entrapment.

Any ions may be carried down during a precipitation as the result of surface
adsorption. Na+ , or Cl- in the case of the determination of SO 42- by the addition of
dilute BaCl2 solution to a NaSO4 solution. Both Ba2+ and Na+ can compete for lattice
positions as the particles form. Likewise, the ions Cl - and SO42- can have the same
effect. In the quantitative determination of some transition metals, iron for example
as Fe(OH)3, zinc, cadmium and manganese may be present as impurities and all
three form sparingly soluble hydroxides as well, though each with greater solubility
than the hydroxide of iron:

Compound Solubility Product


Fe(OH)3 4 x 10-38
Cd(OH)2 2.5 x 10-14
Mn(OH)2 1.9 x 10-9
Zn(OH)2 1.2 x 10-17
Mixed crystal formation can occur if two ions have the same charge if their ionic
diameters are sufficiently close to fit into the same crystal lattice. Ions which
commonly interfere with each other are shown in the table below with their ionic
diameters in picometers given after each.

Interfering Ions
K+, 133 pm NH4 , 148 pm
+

Sr2+, 113 pm Ba2+, 135 pm


Mn2+, 80 pm Cd2+, 97 pm

In cases where one has a known interference of one ion with the other it is
necessary to find methods of removing one before carrying out a precipitation of the
other or using a precipitating reagent in which there is no interference.

Occlusion and mechanical entrapment. If a precipitation procedure is carried out too


quickly, pockets of solvent and spectator ions can form, trapping them within the
precipitate particles and dashing one's hope of removing them during the washing
procedure. This is another reason why the relative supersaturation must be kept as
low as possible so that in principle at least, all precipitation occurs only at the
surface of a growing solid particle, devoid of solvent pockets.

All of these problems of coprecipitation of unwanted ions can lead to positive or


negative errors. In the example above where it pointed out that Na + or Cl- may
coprecipitate in the SO42- determination, surface adsorption will produce a positive
error. In the case of mixed crystal formation, the direction of the error depends on
the relative atomic weight of the ion which replaces that which is desired in the
precipitate. In the case of the precipitation of zinc hydroxide, mixed crystal formation
with manganese would produce a negative error but with cadmium or zinc a positive
error.

Desired precipitate
Compound Mn(OH)2 Zn(OH)2 Cd(OH)2
At. Wt. of M2+ 54.94 65.39 112.41
Direction of error negative ------------- positive

SAMPLE PROBLEMS IN GRAVIMETRIC ANALYSIS


The stoichiometry of a precipitation reaction provides a mathematical relationship
between the analyte and the precipitate. Because a precipitation gravimetric method
may involve several chemical reactions before the precipitation reaction, knowing
the stoichiometry of the precipitation reaction may not be sufficient. Even if you do
not have a complete set of balanced chemical reactions, you can deduce the
mathematical relationship between the analyte and the precipitate using a
conservation of mass. The following example demonstrates the application of this
approach to the direct analysis of a single analyte.

Example 1:
To determine the amount of magnetite, Fe 3O4, in an impure ore, a 1.5419-g sample
is dissolved in concentrated HCl, giving a mixture of Fe 2+ and Fe3+. After adding
HNO3 to oxidize Fe2+ to Fe3+ and diluting with water, Fe 3+ is precipitated as
Fe(OH)3 by adding NH3. Filtering, rinsing, and igniting the precipitate provides
0.8525 g of pure Fe2O3. Calculate the %w/w Fe3O4 in the sample.

Solution:
A conservation of mass requires that all the iron from the ore is found in the Fe 2O3.
We know there are 2 moles of Fe per mole of Fe 2O3 (FW = 159.69 g/mol) and 3
moles of Fe per mole of Fe3O4 (FW = 231.54 g/mol); thus

The % w/w s Fe3O4 in the sample, therefore, is

LESSON 2.0: VOLUMETRIC ANALYSIS


Volumetric or titrimetric analyses are quantitative analytical techniques which
employ a titration in comparing an unknown with a standard. In a titration, a
measured and controlled volume of a standardized solution, a solution containing a
known concentration of reactant «A» from a burette is added incrementally to a
sample solution of known volume (measured by a pipette) containing a substance to
be determined (analyte) of unknown concentration of reactant «B». Th titration
proceeds until reactant «B» is just consumed (stoichiometric completion). This is
known as the equivalence point. (The titration is complete when sufficient titrant has
been added to react with all the analytes.) At this point the number of equivalents of
«A» added to the unknown equals the number of equivalents of «B» originally
present in the unknown.

An indicator, a substance that have distinctly different colors in acidic and basic
media, is usually added to the reaction flask to signal when and if all the analyte has
reacted. The use of indicators enables the end point to be observed. In this, the
titrant reacts with a second chemical, the indicator, after completely reacting with the
analyte in solution. The indicator undergoes a change that can be detected (like
color). The titrant volume required for the detection of the equivalence point is called
the end point. Note that the end point and equivalence point are seldomly the same.
Ideally, we want the equivalence point and the end point to be the same. This
seldom happens due to the methods used to observe end points. As a result, we get
a titration error, the difference between the end point and the equivalence point,
which leads to over titration.
The end point is then the point where sufficient indicator has been converted for
detection. The sequence of events can be demonstrated as below:
followed by

The last step does NOT require that all indicators be converted. In fact, it is best if a
very small percent need to be reacted to make the color change visible.
For volumetric methods of analysis to be useful, the reaction must reach 99%+
completion in a short period of time. In almost all cases, a burette is used to
measure out the titrant. When a titrant reacts directly with an analyte (or with a
reaction the product of the analyte and some intermediate compound), the
procedure is termed a direct titration. The alternative technique is called a back
titration. Here, an intermediate reactant is added in excess of that required to
exhaust the analyte, then the exact degree of excess is determined by subsequent
titration of the unreacted intermediate with the titrant. Regardless of the type of
titration, an indicator is always used to detect the equivalence point. Most
common are the internal indicators, compounds added to the reacting solutions that
undergo an abrupt change in a physical property (usually absorbance or color) at or
near the equivalence point. Sometimes the analyte or titrant will serve this function
(auto indicating). External indicators, electrochemical devices such as pH meters,
may also be used. Ideally, titrations should be stopped precisely at the equivalence
point. However, the ever-present random and systematic error, often results in a
titration endpoint, the point at which a titration is stopped, that is not quite the same
as the equivalence point. Fortunately, the systematic error, or bias may be
estimated by conducting a blank titration. In many cases the titrants not available in
a stable form of well-defined composition. If this is true, the titrant must
be standardized (usually by volumetric analysis) against a compound that is
available in a stable, highly pure form (i.e., a primary standard).

Note that by accurately measuring the volume of the titrant that is added (using a
buret), the amount of the sample can be determined.

For a successful titrimetric analysis, the following need to be true:


 The titrant should either be a standard or should be standardized.
 The reaction should proceed to a stable and well-defined equivalence point.
 The equivalence point must be able to be detected.
 The titrants and sample’s volume or mass must be accurately known.
 The reaction must proceed by a definite chemistry. There should be no
complicating side reactions.
 The reaction should be nearly complete at the equivalence point. In other
words, chemical equilibrium should favor the formation of products.
 The reaction rate should be fast enough to be practical.
Illustration: In the determination of chloride, 50 ml of a 0.1M AgNO 3 solution would
be required to completely react with 0.005 moles of chloride present in solution.

Balanced Equation for the reaction: Ag+(aq) + Cl-(aq) → AgCl(s). Here, 1 mole of


Ag+ ions reacts stoichiometrically with 1 mole of Cl - ions. Therefore 50 ml (0.05L) of
a standard 0.10M AgNO3 which contains 0.005 moles (= 0.10 moles L-1 x 0.050 L)
requires an equivalent number of moles of Cl - ions.

Since the titrant solution must be of known composition and concentration, we


ideally would like to start with a primary standard material, a high purity compound
used to prepare the standard solution or to standardize the solution
with. A standard solution is one whose concentration is known. The concentration
of a standard solution is usually expressed in molarity (mole/liter). The process by
which the concentration of a solution is determined is called standardization.
Because of the availability of some substances known as primary standards, in
many instances the standardization of a solution is not necessary. Primary standard
solutions are analytically pure, and by dissolving a known amount of a primary
standard in a suitable medium and diluting to a definite volume, a solution of known
concentration is readily prepared. Most standard solutions, however, are prepared
from materials that are not analytically pure and they have to be standardized
against a suitable primary standard.

The following are the desired requirements of a primary standard:


 High purity
 Stable in air and solution: composition should be unaltered in the air at
ordinary or moderately high temperatures.
 Not hygroscopic.
 Inexpensive
 Large formula weight: equivalence weight ought to be high in order to reduce
the effects of small weighing error.
 Readily soluble in the solvent under the given conditions of the analysis.
 On titration, no interfering product(s) should be present.
 The primary standard should be colorless before and after titration to avoid
interference with indicators.
 Reacts rapidly and stoichiometrically with the analyte.

The following are also the desired requirements of a primary standard solution:
 Have long term stability in solvent.
 React rapidly with the analyte.
 React completely with analyte.
 Be selective to the analyte.

Apparatus for titrimetric analysis:


The most common apparatus used in volumetric determinations are the pipette,
buret, measuring cylinder, volumetric and conical (titration) flask. Reliable
measurements of volume is often done with the help of a pipet, buret, and a
volumetric flask. The conical flask is preferred for titration because it has a
good“mouth” that minimizes the loss of the titrant during titration.
 

Classification of reactions in volumetric (titrimetric) analysis


Any type of chemical reactions in solution should theoretically be used for trimetric
analysis. However, the reactions most often used fall under two main categories:
(a) Those in which no change in oxidation state occurs. These are dependent on
combination of ions.
(b) Oxidation-reduction reactions: These involve a change of oxidation state (i.e.,
the transfer of electrons).

For convenience, however, these two types of reactions are further divided into four
main classes:
(i) Neutralization reactions or acidimetry and alkalimetry: HA+B⇌HB+ +A-
(ii) Precipitation reactions: M(aq) +nL(aq) ⇌ MLn(s)
(iii) Oxidation-reduction reactions: Ox + Red ⇌ Red' + ox'
(iv) Complex ion formation reactions: M(aq) +nL(aq) ⇌ MLn(aq)

LESSON 2.0: VOLUMETRIC ANALYSIS PRINCIPLES

In titrimetry we add a reagent, called the titrant, to a solution that contains another


reagent, called the titrand, and allow them to react. The type of reaction provides us
with a simple way to divide titrimetry into four categories: acid–base titrations, in
which an acidic or basic titrant reacts with a titrand that is a base or an acid;
complexometric titrations , which are based on metal–ligand complexation; redox
titrations, in which the titrant is an oxidizing or reducing agent; and precipitation
titrations, in which the titrand and titrant form a precipitate.

Before we consider individual titrimetric methods in greater detail, let’s take a


moment to consider some of these similarities. As you work through this chapter,
this overview will help you focus on the similarities between different titrimetric
methods. You will find it easier to understand a new analytical method when you
can see its relationship to other similar methods.

Equivalence Points and End Points


If a titration is to give an accurate result, we must combine the titrand and the titrant
in stoichiometrically equivalent amounts. We call this stoichiometric mixture
the equivalence point. Unlike precipitation gravimetry, where we add the precipitant
in excess, an accurate titration requires that we know the exact volume of titrant at
the equivalence point, Veq. The product of the titrant’s equivalence point volume and
its molarity, MT, is equal to the moles of titrant that react with the titrand.

If we know the stoichiometry of the titration reaction, then we can calculate the
moles of titrand.

Unfortunately, for most titration reactions there is no obvious sign when we reach
the equivalence point. Instead, we stop adding the titrant at an end point of our
choosing. Often this end point is a change in the color of a substance, called
an indicator, that we add to the titrand’s solution. The difference between the end
point’s volume and the equivalence point’s volume is a determinate titration error. If
the end point and the equivalence point volumes coincide closely, then this error is
insignificant and is safely ignored. Clearly, selecting an appropriate end point is of
critical importance.

Volume as a Signal
Instead of measuring the titrant’s volume, we may choose to measure its mass.
Although generally we can measure mass more precisely than we can measure
volume, the simplicity of a volumetric titration makes it the more popular choice.

Almost any chemical reaction can serve as a titrimetric method provided that it
meets the following four conditions. The first condition is that we must know the
stoichiometry between the titrant and the titrand. If this is not the case, then we
cannot convert the moles of titrant used to reach the end point to the moles of
titrand in our sample. Second, the titration reaction effectively must proceed to
completion; that is, the stoichiometric mixing of the titrant and the titrand must result
in their complete reaction. Third, the titration reaction must occur rapidly. If we add
the titrant faster than it can react with the titrand, then the end point and the
equivalence point will differ significantly. Finally, we must have a suitable method for
accurately determining the end point. These are significant limitations and, for this
reason, there are several common titration strategies.

Depending on how we are detecting the endpoint, we may stop the titration too early
or too late. If the end point is a function of the titrant’s concentration, then adding the
titrant too quickly leads to an early end point. On the other hand, if the end point is a
function of the titrant’s concentration, then the end point exceeds the equivalence
point.

1. Direct Titration
A simple example of a precipitation titration is an analysis for Ag + using thiocyanate,
SCN–, as a titrant.
This reaction occurs quickly and with a known stoichiometry, which satisfies two of
our requirements. To indicate the titration’s end point, we add a small amount of
Fe3+ to the analyte’s solution before we begin the titration. When the reaction
between Ag+ and SCN– is complete, formation of the red-colored
Fe(SCN)  complex signals the end point. This is an example of a direct
2+

titration since the titrant reacts directly with the analyte.

2. Back-titration
If the titration’s reaction is too slow, if a suitable indicator is not available, or if there
is no useful direct titration reaction, then an indirect analysis may be possible.
Suppose you wish to determine the concentration of formaldehyde, H 2CO, in an
aqueous solution. The oxidation of H2CO by 

is a useful reaction, but it is too slow for a titration. If we add a known excess of
I3-  and allow its reaction with H2CO to go to completion, we can titrate the unreacted
I3-  with thiosulfate, 

The difference between the initial amount of I 3-   and the amount in excess gives us
the amount of I3-   that reacts with the formaldehyde. This is an example of a back
titration. The type of titration that was discussed previously is an example of redox
titration. 

3. Displacement titration
Calcium ions play an important role in many environmental systems. A direct
analysis for Ca2+ might take advantage of its reaction with the ligand
ethylenediaminetetraacetic acid (EDTA), which we represent here as Y 4–.

Unfortunately, for most samples this titration does not have a useful indicator.
Instead, we react the Ca2+ with an excess of MgY2–

releasing an amount of Mg2+ equivalent to the amount of Ca2+ in the sample.


Because the titration of Mg2+ with EDTA

has a suitable end point, we can complete the analysis. The amount of EDTA used
in the titration provides an indirect measure of the amount of Ca 2+ in the original
sample. Because the species we are titrating was displaced by the analyte, we call
this a displacement titration.

4. In-direct titration
If a suitable reaction with the analyte does not exist it may be possible to generate a
species that we can titrate. For example, we can determine the sulfur content of coal
by using a combustion reaction to convert sulfur to sulfur dioxide
S(s)+O2(g)→SO2(g)
and then convert the SO2 to sulfuric acid, H2SO4, by bubbling it through an aqueous
solution of hydrogen peroxide, H2O2.

Titrating H2SO4 with NaOH

provides an indirect determination of sulfur.

LESSON 2.1: TITRATION CURVE


To find a titration’s end point, we need to monitor some property of the reaction that
has a well-defined value at the equivalence point. For example, the equivalence
point for a titration of HCl with NaOH occurs at a pH of 7.0. A simple method for
finding the equivalence point is to monitor the titration mixture’s pH using a pH
electrode, stopping the titration when we reach a pH of 7.0. Alternatively, we can
add an indicator to the titrand’s solution that changes color at a pH of 7.0.

Suppose the only available indicator changes color at a pH of 6.8. Is the difference
between this end point and the equivalence point small enough that we safely can
ignore the titration error? To answer this question we need to know how the pH
changes during the titration.

A titration curve provides a visual picture of how a property of the titration reaction


changes as we add the titrant to the titrand. The titration curve in Figure 1, for
example, was obtained by suspending a pH electrode in a solution of 0.100 M HCl
(the titrand) and monitoring the pH while adding 0.100 M NaOH (the titrant). A close
examination of this titration curve should convince you that an end point pH of 6.8
produces a negligible titration error. Selecting a pH of 11.6 as the end point,
however, produces an unacceptably large titration error.
Figure 1 . Typical acid–base
titration curve showing how the
titrand’s pH changes with the
addition of titrant. The titrand is
a 25.0 mL solution of 0.100 M
HCl and the titrant is 0.100 M
NaOH. The titration curve is the
solid blue line, and the
equivalence point volume (25.0
mL) and pH (7.00) are shown
by the dashed red lines.
The green dots show two end
points. The end point at a pH of
6.8 has a small titration error,
and the end point at a pH of
11.6 has a larger titration error.
The shape of the titration curve in Figure 1  is not unique to an acid–base titration.
Any titration curve that follows the change in concentration of a species in the
titration reaction (plotted logarithmically) as a function of the titrant’s volume has the
same general sigmoidal shape. Several additional examples are shown in Figure 2 .

Figure 2 . Additional examples of titration curves. (a) Complexation titration of 25.0
mL of 1.0 mM Cd2+ with 1.0 mM EDTA at a pH of 10. The y-axis displays the
titrand’s equilibrium concentration as pCd. (b) Redox titration of 25.0 mL of 0.050 M
Fe2+ with 0.050 M Ce4+ in 1 M HClO4. The y-axis displays the titration mixture’s
electrochemical potential, E, which, through the Nernst equation is a logarithmic
function of concentrations. (c) Precipitation titration of 25.0 mL of 0.10 M NaCl with
0.10 M AgNO3. The y-axis displays the titrant’s equilibrium concentration as pAg.

LESSON 2.2: ACID-BASE TITRATION


Acid-Base titrations are usually used to find the amount of a known acidic or basic
substance through acid base reactions. The analyte (titrand) is the solution with an
unknown molarity. The reagent (titrant) is the solution with a known molarity that will
react with the analyte.

LESSON 2.3: COMPLEX TITRATION


The earliest examples of metal–ligand complexation titrations are Liebig’s
determinations, in the 1850s, of cyanide and chloride using, respectively, Ag + and
Hg2+ as the titrant. Practical analytical applications of complexation titrimetry were
slow to develop because many metals and ligands form a series of metal–ligand
complexes. Liebig’s titration of CN – with Ag+ was successful because they form a
single, stable complex of
, which results in a single, easily identified end point. Other metal–ligand complexes,
such as  are not analytically useful because they form a series of metal–
ligand
complexes (CdI+, CdI2(aq),  that produce a sequence of
poorly defined end points. 

In 1945, Schwarzenbach introduced aminocarboxylic acids as multidentate ligands.


The most widely used of these new ligands—ethylenediaminetetraacetic acid, or
EDTA—forms a strong 1:1 complex with many metal ions. The availability of a
ligand that gives a single, easily identified end point made complexation titrimetry a
practical analytical method.
LESSON 2.4: REDOX TITRATION
Analytical titrations using oxidation–reduction reactions were introduced shortly after
the development of acid–base titrimetry. The earliest redox titration took
advantage of chlorine’s oxidizing power. In 1787, Claude Berthollet introduced a
method for the quantitative analysis of chlorine water (a mixture of Cl 2, HCl, and
HOCl) based on its ability to oxidize indigo, a dye that is colorless in its oxidized
state. In 1814, Joseph Gay-Lussac developed a similar method to determine
chlorine in bleaching powder. In both methods the end point is a change in color.
Before the equivalence point the solution is colorless due to the oxidation of indigo.
After the equivalence point, however, unreacted indigo imparts a permanent color to
the solution.

The number of redox titrimetric methods increased in the mid-1800s with the
introduction of and I 2 as oxidizing titrants, and of Fe2+ and    
as reducing titrants. Even with the availability of these new titrants, redox titrimetry
was slow to develop due to the lack of suitable indicators. A titrant can serve as its
own indicator if its oxidized and its reduced forms differ significantly in color. For
example, the intensely purple   ion serves as its own indicator since its reduced
form, Mn , is almost colorless. Other titrants require a separate indicator. The first
2+

such indicator, diphenylamine, was introduced in the 1920s. Other redox indicators
soon followed, increasing the applicability of redox titrimetry.

LESSON 2.4: PRECIPITATION TITRATION


A reaction in which the analyte and titrant form an insoluble precipitate also can
serve as the basis for a titration. We call this type of titration a precipitation
titration.

One of the earliest precipitation titrations—developed at the end of the eighteenth


century—was the analysis of K2CO3 and K2SO4 in potash. Calcium nitrate, Ca(NO3)2,
was used as the titrant, which forms a precipitate of CaCO 3 and CaSO4. The
titration’s end point was signaled by noting when the addition of titrant ceased to
generate additional precipitate. The importance of precipitation titrimetry as an
analytical method reached its zenith in the nineteenth century when several
methods were developed for determining Ag + and halide ions.
Module 3: Ultraviolet and Visible Spectroscopy

Lesson 3.1: Theory

Principle
The absorption phenomenon is based on the interaction of electromagnetic radiation
and electrons. When a molecule absorbs light, radiation induces electronic
transitions, i.e. the transition of an electron from its fundamental to an excited
electronic state (Perrin-Jablonski diagram, see below):

Ferrin-Jablonski Diagram
The electromagnetic wave is characterized by its energy or its wavelength based on
the following equation: E = hc/λ with:
h = 6.62x10-34 J.s (Planck constant)
c = 3x108 m/s (speed of light)
λ = wavelength (in nm)

 As electronic states of a molecule are quantified, absorption can occur if the


following rule is confirmed: ΔE = Eexcited state – Efundamental state
 It means that the energy of the light radiation is at least above or equal to the
difference between both states, the fundamental and the excited one.
 The energy of molecular orbitals being specific to each group or chemical
function, absorption wavelength is thus correlated to the type of chemical
bonds.

Absorption Spectroscopy
Absorption spectroscopy is a technique based on the interaction between radiation
and matter. It allows the characterization and quantification of liquids, solids, and
gases.

Electromagnetic spectrum

Generally, wavelength ranges probed are:


 200 to 400 nm for ultra-violet (UV)
 400 to 800 nm for the visible region (this range of wavelength is related to the
colour of a molecule)
 > 800 nm for near infra-red

In the UV domain, we may consider 4 categories depending on the wavelength:


 UV-A (315 – 400 nm)
 UV-B (280 - 315 nm)
 UV-C (100 - 280 nm)
 Extreme UV (10 - 100 nm)

UV-visible absorption spectroscopy is often used as a detection method for


separative techniques such as liquid chromatography.

Absorption spectrum
An absorption spectrum represents the absorption of a molecule as a function of
wavelength.

where-in:
X-axis is wavelength in nm and the Y-axis is absorbance (A) or transmittance (T).
   Aλ = log I0/I without unit
    Tλ = 10-Aλ = I/I0
Chemical groups responsible for absorption are called chromophores.
For gases, absorption bands are thin, although for liquids they can be tens of
nanometers wide.

LESSON 3.2: SPECTRA INTERPRETATION

I. Electronic transitions
After excitation by electromagnetic radiation, valence electrons take part in
electronic transitions. Only four different transitions are allowed between σ, π and
lone pair electrons.

Transition σ → σ*
 occurs only for simple bonds (C-H or C-C for example)
 requires a high quantity of energy, and is thus found in the UV region (<200
nm)

Transition n → σ*
 occurs for atoms with lone pair electrons (N, O, S….)
 requires a high quantity of energy, and is thus found between 150 and 250
nm
Transition n → π* and n → π
 occurs only for atoms with lone pair electrons and for unsaturated functions
 there is sometimes a transition n → π when the level of the doublet is below
the level of the π orbital
 these transitions usually take place in the UV domain
Transition π → π* (aromatics)
 occurs only for atoms with unsaturated functions
 are located in the UV and visible domain (200 to 700 nm)
II. Absorption bands position
          The position of an absorption band can be influenced by several factors such
as solvent, steric, inductive or mesomeric effects.

Where-in:

HYPSOCHROMIC SHIFT is a shift towards


lower wavelengths (UV)

BATHOCHROMIC SHIFT is a shift towards


higher wavelengths (IR)

Example:

In chloroform, the absorbance of caffeine is upshifted (bathochromic shift) compared


to its absorption in water. The energy difference between fundamental and excited
states is smaller in chloroform.
 
Woodward-Fieser rules:
The Woodward and Fieser rules allow to predict the wavelength for the maximum of
absorption of conjugated systems.  When the number of delocalized electrons
increases in a molecule, its chromophores face a bathochromic effect.

Molecule A possesses one C=C bond responsible for its absorption at 185 nm.
Molecule B possesses the same chromophore but with delocalized electrons. Its
absorption band is then shifted to higher wavelengths: bathochromic effect.

Molecules A, B and C possess the same chromophore, C=C, but with an increasing
number of delocalized electrons. Their absorption is therefore upshifted
(bathochromic effect).

III. Color-absorption relationship


In the visible range, the color of molecules is the first indication of their absorption
wavelength: a solution is colored, which means that the molecule absorbs light
corresponding to its complementary color i.e. its opposite color in the chromatic
circle.
Example:

IV. Absorption bands intensity


Absorption bands intensity is related to the electronic transition probability, with a
molar absorption coefficient ε λmax which is valid for one wavelength.

ε ≤ 10 forbidden transition
10 ≤ ε ≤ 1000 partly allowed transition
1000 ≤ ε ≤ 100000 allowed transition
ε ≥ 100000 highly allowed transition

Absorption bands intensity is measured with the absorbance Aλ = log I0/I


For a defined wavelength, the Beer-Lambert law links the absorbance to the
concentration of species in solution.
Aλ = ελ.l.c
with ελ the molar absorption coefficient (M-1.cm-1)
l the path length
c the concentration (mol.L-1)
The absorbance of a compound is proportional to the number of
chromophores it possesses.

VALIDITY OF THE BEER-LAMBERT LAW

The Beer-Lambert law is valid only if the exciting light is monochromatic, with a
homogeneous system, isotropic and transparent. Besides, absorbing species must
be non-photosensitive and in relatively low concentration, especially to ensure a
sufficient quantity of light being detected by the spectrophotometer. Other factors,
such as aggregates formation or species change upon dilution, may be responsible
for a loss in linearity. (The absorbance is no longer proportional to concentration.

Additivity of the Beer-Lambert law


For a fixed wavelength, the absorbance of a mixture is the sum of the absorbances
of each species.
As absorption bands position, their intensity can be modified by different factors (pH,
solvent…).

Where-in: 

The hyperchromic effect corresponds to


an increase in absorbance.

The hypochromic effect corresponds to


a decrease in absorbance.

LESSON 3.3: INSTRUMENTATION

A spectrophotometer
A "single-beam" spectrophotometer is made up of:
 a polychromatic source
 a wavelength selector and eventually a source selector
 a sample compartment
 a detector
 a screen and/or a computer.
While, a double-beam spectrophotometers are more sophisticated and able to
simultaneously measure a sample and a reference. I 0 and I are both determined and
compared to obtain a resulting spectrum.

How does a spectrophotometer work?


"White" light is obtained from a polychromatic source, then a monochromator
extracts the radiation at the wavelength requested by the user. Beam intensity is
measured after it has passed through the sample in the cuvette. This intensity is
compared to a reference value – the resulting value displayed by the instrument
reflects this comparison. Thus, for a single-beam spectrophotometer, the reference
spectrum is recorded with the solution (solvent or buffer) but without the sample. For
a double-beam spectrophotometer, one cuvette contains the sample and the other
one only contains the solution: the difference between both absorbances is then
directly done by the system.

You might also like