Download as docx, pdf, or txt
Download as docx, pdf, or txt
You are on page 1of 44

Analytical Chemistry

Kristel M. Salalila, LPT

1
Table of Contents

Module 1: Solution Equilibria


Introduction 33
Learning Objectives 33
Lesson 1. Solubility Equilibrium Concepts 34
Lesson 2. Equilibria in Solutions and The Solubility Product 37
Lesson 3. Calculations involving Solubility Equilibria 41
Assessment Task 43
Summary 44
References 45

Module 2: pH Calculations and Acid Base Equilibria 46


Introduction 46
Learning Objectives 46
Lesson 1. Properties of acids and bases 47
Lesson 2. Bronsted-Lowry Conjugate acid and base 51
Lesson 3. Acid-Base equilibria 54
Assessment Task 57
Summary 59
References 60

Module 3: Gravimetry and Titrimetry 61


Introduction 61
Learning Objectives 61
Lesson 1. Gravimetry 62
Lesson 2. Titrimetry 65
Assessment Task 68
Summary 69
References 70

2
List of Figures
Figure Title Page
4.1 Solubility Product Constants 41
5.1 Conjugate acid base pairings in acetic acid and water 57
reaction
6.1 Titration curves. 65

3
List of Tables

Table Title Page


4.1 Solubility Product Constants 41
5.1 Nomenclature of some of the common acids. 48
5.2 Comparison of acids and bases 48
5.3 Common bases 49
5.4 Relative strengths of acids and bases. 51
6.1 Indicators and the pH range and color change. 64

4
MODULE 4
Solubility Equilibria

Introduction

Solubility, degree to which a substance dissolves in a solvent to make a solution (usually


expressed as grams of solute per liter of solvent). Solubility of one fluid (liquid or gas) in another
may be complete (totally miscible; e.g., methanol and water) or partial (oil and water dissolve
only slightly). In general, “like dissolves like” (e.g., aromatic hydrocarbons dissolve in each other
but not in water). Some separation methods (absorption, extraction) rely on differences in
solubility, expressed as the distribution coefficient (ratio of a material’s solubilities in two
solvents). Generally, solubilities of solids in liquids increase with temperature and those of
gases decrease with temperature and increase with pressure. A solution in which no more
solute can be dissolved at a given temperature and pressure is said to be saturated (Britannica,
n.d.).

Learning Outcomes

At the end of this module, students will be able to:

1. Discuss concepts involving solubility equilibria;


2. Define solubility product constant Ksp; and
3. Solve problems involving solubility equilibria.

33
Lesson 1. Solubility Equilibrium Concepts
The laws of thermodynamics govern the behavior of all species in solution. Every reacti on
depends upon the thermodynamic properties of the species involved. Where those properties
are changed by the solvent by association, by reaction or temperature, the behavior will alter.
Physical and chemical equilibria in solution are most important. The equilibrium amount of
solute which will dissolve in a given amount of solvent at a given temperature and pressure
is called the solubility. The solubility may be quoted in any units of concentration, for
example, mol m-3, molarity, mole fraction, mass per unit volume or parts per million (ppm).
There is a general ‘rule of thumb’ that ‘like dissolves like’. For example, a nonpolar hydrocarbon
solvent such as hexane would be a very good solvent for solid hydrocarbons such as dodecane
or naphthalene. An ester would be a good solvent for esters, and water or other polar solvents
are appropriate for polar and ionic compounds (Kealey & Haines, 2016).

A. Solubility equilibrium for different phases

1. GASES

Gases dissolve in solvents according to Henry’s Law, provided they do not react with the
solvent: pB = xB K where xB is the mole fraction of solute gas B which dissolves at a partial
pressure pB of B, and K is a constant at a given temperature. This is analytically important for
several reasons. For example, nitrogen is bubbled through solutions to decrease the partial
pressure of oxygen in electrochemical experiments. Similarly, air is removed from liquid
chromatography solvents by assing helium through them, or by boiling them, since gas solubility
decreases as the temperature is increased (Kealey & Haines, 2016).

2.LIQUIDS

Liquids. When different liquids are mixed, many types of behavior may occur. If the
molecules in the liquids are of similar size, shape, polarity and chemical nature they may mix in
all proportions. For example, benzene and methylbenzene (toluene) mix completely. In such
ideal
34
solutions, obeying Raoult’s law, the activity coefficient is close to 1: a = p/pn = x If the
component molecules differ greatly in polarity, size or chemical nature (e.g., water and
tetrachloromethane) they may not mix at all. This is an important condition for solvent extraction
(Topic D1). The distribution of a solute between a pair of immiscible liquids depends primarily on
the solubility of the solute in each liquid (Kealey & Haines, 2016).

3. SOLIDS

Solids generally follow the ‘like dissolves like’ rule. Nonpolar, covalent materials dissolve
best in nonpolar solvents. Solid triglycerides such as tristearin are extracted by diethyl ether, but
are nearly insoluble in water. Salts, such as sodium chloride are highly soluble in water, but
virtually insoluble in ether. The solubility of solids in liquids is an important process for the
analyst, who frequently uses dissolution as a primary step in an analysis or uses precipitation as
a separation procedure. The dissolution of a solid in a liquid is favored by the entropy change as
explained by the principle of maximum disorder discussed earlier. However, it is necessary to
supply energy in order to break up the lattice and for ionic solids this may be several hundred
kilojoules per mole. Even so many of these compounds are soluble in water. After break-up of
the lattice the solute species are dispersed within the solvent, requiring further energy and
producing some weakening of the solvent-solvent interactions. (Kealey & Haines, 2016).
The energy needed to bring about this change can only be supplied by the solvation of the
solute species utilizing van der Waals' type ion-dipole interactions and to a lesser extent dipole-
dipole reactions, etc. The total energy available from this source, the solvation energy, may be
in the range –400 to –4000 kj mol–1 for aqueous systems containing ionic solids. On the other
hand non-ionizing solvents will have a much lower ability to produce solvation and insufficient
energy is available to break up ionic lattices and produce a solution. Solvent-solute interactions
between an ionizing solvent and a covalent solid are small but may be large enough to
overcome the low lattice energy for many solids although their solubility will still be very low as a
result of the associated nature of the solvent. Covalent solventsolute systems have minimal
interactions and may approximate in behavior to ideal mixtures (Fifield & Kealey, 2000).

B. SOLUBILITY OF IONS

35
According to Kealey & Haines (2016), the behavior of ions in solution may be summarized
as follows:
(i) Solids whose structure consists of ions held together by electrostatic forces (e.g.
NaCl) must be separated into discrete ions when they dissolve. These ions often
gain stability by solvation with molecules of the solvent. Such solutions are described
as strong electrolytes.

(ii) Some covalent molecules, such as ethanoic acid, may form ions by the unequal
breaking of a covalent bond, followed by stabilization by solvation. This occurs only
partially and these are called weak electrolytes.

Equation 4.1 H: OCOCH3 H+ + - OCOCH3

(iii) In some cases, ions do not separate completely. In concentrated solutions,


oppositely charged ions may exist as ion-pairs, large ions of surfactants may
aggregate into micelles, which are used in capillary electrophoresis.

C. The pX Notation

The concentration of species in solution may range from very small to large. For example, in
a saturated aqueous solution of silver chloride, the concentration of silver ions is about 10-5 M,
while for concentrated hydrochloric acid the concentration of hydrogen and chloride ions is
about 10 M. For convenience, a logarithmic scale is often used:

Equation 4.2. pX = -log (X)

where X is the concentration of the species, or a related quantity. Thus, for the examples
above, pAg = 5 in saturated aqueous silver chloride and pH = -1 in concentrated HCl. Since
equilibrium constants are derived from activities or concentrations as noted below, this notation
is also used for them: (Kealey & Haines, 2016).

36
Equation 4.3. pK = -log (K)

Lesson 2. Equilibria in Solutions


Most reactions will eventually reach equilibrium. That is, the concentrations of reactants
and products change no further, since the rates of the forward and reverse reactions are the
same. From the above arguments concerning solutions, and from the laws of thermodynamics,
any equilibrium in solution involving species D, F, U and V (Kealey & Haines, 2016).

Equation 4.4. D + F U+V

will have an equilibrium constant, KT, at a particular temperature T given by:

Equation 4.5. KT = (aU . aV)/(aD . aF)

where the activities are the values at equilibrium. It should be noted that KT changes
with temperature.

The larger the equilibrium constant, the greater will be the ratio of products to reactants
at equilibrium. There are many types of equilibria that occur in solution, but for the important
analytical conditions of ionic equilibria in aqueous solution, four examples are very important
(Kealey & Haines, 2016).

A. Acid and base Dissociation

According to Haines & Kealey (2016), In aqueous solution, strong electrolytes (e.g.,
NaCl, HNO3, NaOH) exist in their ionic forms all the time. However, weak electrolytes exhibit
dissociation equilibria. For ethanoic acid, for example:

Equation 4.6. HOOCCH3 + H2O H3O+ + CH3COO

Ka = (aH . aA)/(aHA . aW) = 1.75 x 10-5 where HA, W, H and A represent each of the
species in the above equilibrium. In dilute solutions the activity of the water aW is close to 1.

For ammonia:

Equation 4.7. NH3 + H2O NH4 + + OH

37
Equation 4.8. Kb = (aNH4+ . aOH-)/(aNH3 . aW) = 1.76 x 10-5

Water behaves in a similar way:

Equation 4.9. 2 H2O H3O+ + OHKW = (aH3O+ . aOH-) = 10-14

B. Complexation Equilibria

The reaction between an acceptor metal ion M and a ligand L to form a complex ML is
characterized by an equilibrium constant. but a simple example will suffice here:

Equation 4.10. M(aq) + L(aq) ML(aq)

Equation 4.11. Kf = (aML)/(aM . aL)

For example, for the copper-EDTA complex at 25 C: Kf = 6.3 x 1018

C. Solubility equilibria

If a compound is practically insoluble in water, this is useful analytically because it


provides a means of separating this compound from others that are soluble. The technique of
gravimetric analysis has been developed to give very accurate analyses of materials by
weighing pure precipitates of insoluble compounds to give quantitative measurements of their
concentration. For the quantitative determination of sulfate ions, SO4 2- , the solution may be
treated with a solution of a soluble barium salt such as barium chloride BaCl2, when the
following reaction occurs:

Equation 4.12. Ba2+ + SO4 2- BaSO4(s)

Conversely, if solid barium sulfate is put into water:

Equation 4.13. BaSO4 (s) = Ba2+ + SO4 2-

The solubility product, Ksp, is an equilibrium constant for this reaction Ksp = a(Ba 2+ ) .
a(SO4 2- ) = 1.2 x10-10 bearing in mind that the pure, solid BaSO 4 has a = 1. (Haines & Kealey,
2016). This means that a solution of barium sulfate in pure water has a concentration of sulfate
ions of only 1.1x10-5 M. The concentration of the barium ions is the same.

D. Redox Equilibria

38
When a species gains electrons during a reaction, it undergoes reduction and,
conversely, when a species loses electrons it undergoes oxidation. In the total reaction, these
processes occur simultaneously, for example:

Equation 4.14. Ce4+ + Fe2+ = Ce3+ + Fe3+

The cerium is reduced from oxidation state 4 to 3, while the iron is oxidized from 2 to 3.
Any general ‘redox process’ may be written: Ox1 + Red2 = Red1 + Ox2 The equilibrium
constant of redox reactions is generally expressed in terms of the appropriate electrode
potentials but for the above reaction (Haines & Kealey, 2016).

Equation 4.15. K = (a(Ce3+ ) . a(Fe3+ ))/(a(Ce4+ ) . a(Fe2+ )) = 2.2 x 1012

The solubility product, Ksp

In this section, we consider the equilibria associated with solids dissolving in water to
form aqueous solutions. When an ionic solids dissolves in water, we assume that it dissociates
into separated cations and anions (Zumdahl, 1998).

Equation 4.16. CaF2 in H2O Ca2+ + 2F-

When the solid salt is first added to water, no Ca2+ and 2F- are present. However, when
dissolution proceeds, the concentration of Ca2+ and 2F- increase, making it more and more likely
that these ions will collide and reform the solid phase. Thus, two competing processes are
occurring. (Zumdahl, 1998)

Equation 4.17. Ca2+ + 2F- in H2O CaF2

When a dynamic equilibrium is reached: at this point, no more CaF2 is dissolved and the
solution is said to be saturated.

Equation 4.18. CaF2 Ca2+ + 2F-

The equilibrium constant expression of the reaction is:

Ksp= [Ca2+][F-]2 where the concentration [ ] of Ca and F is expressed in mol/L or M.

39
The constant Ksp is called the solubility product constant or simply the solubility product
for equilibrium expression (Zumdahl, 1998).

Table 4.1. Solubility product constants.

Source: https://courses.lumenlearning.com/introchem/chapter/solubility-product/

40
Lesson 3. Calculations involving Solubility Equilibria (Kealey &
Haines, 2016)

Sample Problem 1

Compute for the solubility of the reaction of Ag+ and Br-. The Ksp value of AgBr at 25 ˚C is
5.0x10-13

Step 1: Write the balanced equilibrium reaction:

AgBr Ag+ + Br-

Step 2: Write the Ksp expression.

Ksp= [Ag+][Br-]

Step 3: Make an ICE (Initial, Change and Equilibrium) table

Ag+ Br-
Initial 0 0
Change +x +x
Equilibrium +x +x

Step 4: Substitute from the Ksp equation:

5.0x10-13= [x][x]

5.0x10-13= [x][x]= x2

5.0x10-13= x2

41
Get the square root to obtain x= 7.07 x10-7 M

To check: Ksp= 5.0x10-13= [7.07 x10-7 M] [7.07 x10-7 M]

Sample Problem 2

Compute for the solubility of the reaction of Ca2+ and F-. The Ksp value of CaF2 at 25 ˚C is
4.0x10-11

Step 1: Write the balanced equilibrium reaction:

CaF2 Ca2+ + 2F-

Step 2: Write the Ksp expression.

Ksp= [Ca2+][F-]2

Step 3: Make an ICE (Initial, Change and Equilibrium) table

Ca2+ 2F-
Initial 0 0
Change +x +2x
Equilibrium +x +2x

Step 4: Substitute from the Ksp equation:

4.0x10-11= [x][2x]2

4.0x10-11= [x][4x2]= 4x3

4.0x10-11= 4x3

4 4

Get the cube root to obtain x= 2.15 x10-4 M

42
To check: Ksp= 4.0x10-11= [2.15 x10-4][2(2.15 x10-4)]2

Assessment Task 4

1. Enumerate and describe the different concepts in solubility.

2. Write the solubility product expression of the following:


a. PbSO4
b. Ca3 (PO4)2
c. Fe (OH)3

3. What is the solubility of the reaction of PbCl2 if the Ksp value of PbCl2 is 1.6x10-5

4. Compute for the solubility of the reaction of Fe2+ and CO32-. The Ksp value of FeCO3 at
25 ˚C is 2.1x10-11

5. Compute for the solubility of the reaction of Ca2+ and F-. The Ksp value of CaF2 at 25 ˚C
is 4.0x10-11

6. Compute for the solubility of the reaction of Ag+ and SO42-. The Ksp value at 25 ˚C is
1.2x10-5

43
Summary

The equilibrium amount of solute which will dissolve in a given amount of solvent at a
given temperature and pressure is called the solubility. Solubility can be mathematically
expressed using the solubility product constant or Ksp.

44
References

Britannica. (n.d.).https://www.britannica.com/science/solubility-chemistry.
Fifield, FW., Kealey, D. (2000).Principles and Practice of Analytical Chemistry.
Blackwell Science Ltd.
Kealey, D., & Haines, P.J. (2016). Analytical Chemistry. Instant Notes series. BIOS Scientific
Publishers Ltd.
Zumdahl, S. (1998). Chemical Principles. 1998. Houghton Mifflin Company

45
MODULE 5
pH Calculations and Acid Base Equilibria

Introduction

The acidity or alkalinity of a reaction mixture is most important. It can control the rate of
reaction, the nature of the species present and even subjective properties such as taste. The
original definition of pH (Sorensen, 1909) related it to the concentration of hydrogen ions
(Kealey & Haines, 2016).

Learning Outcomes

At the end of this module, students will be able to:

1. Characterize acids and bases

2. Calculate and express pH, pOH; and

3. Solve for acid-base equilibria problems.

46
Lesson 1. Properties of Acids and Bases

A. Acids

Acids and bases are important substances in health, industry, and the environment. One
of the most common characteristics of acids is their sour taste. Lemons and grapefruits taste
sour because they contain acids such as citric and ascorbic acid (vitamin C). Vinegar tastes
sour because it contains acetic acid. We produce lactic acid in our muscles when we exercise.
Acid from bacteria turns milk sour in the production of yogurt and cottage cheese. We have
hydrochloric acid in our stomachs that helps us digest food. Sometimes we take antacids, which
are bases such as sodium bicarbonate or milk of magnesia, to neutralize the effects of too much
stomach acid. The term acid comes from the Latin word acidus, which means “sour.” We are
familiar with the sour tastes of vinegar and lemons and other common acids in foods. In 1887,
the Swedish chemist Svante Arrhenius was the first to describe acids as substances that
produce hydrogen ions (H+) when they dissolve in water. Because acids produce ions in water,
they are also electrolytes. For example, hydrogen chloride dissociates in water to give hydrogen
ions, H+, and chloride ions, Cl-. The hydrogen ions give acids a sour taste, change the blue
litmus indicator to red, and corrode some metals (Timberlake, C., & Timberlake, W., 2014).

Naming of Acids

Acids dissolve in water to produce hydrogen ions, along with a negative ion that may be
a simple nonmetal anion or a polyatomic ion. When an acid dissolves in water to produce a
hydrogen ion and a simple nonmetal anion, the prefix hydro is used before the name of the
nonmetal, and its ide ending is changed to ic acid. For example, hydrogen chloride (HCl)
dissolves in water to form HCl(aq), which is named hydrochloric acid. An exception is hydrogen
cyanide (HCN), which as an acid is named hydrocyanic acid. When an acid contains oxygen, it
dissolves in water to produce a hydrogen ion and an oxygen-containing polyatomic anion. The
most common form of an oxygen-containing acid has a name that ends with ic acid. The name
of its polyatomic anion ends in ate. If the acid contains a polyatomic ion with an ite ending, its
name ends in ous acid. The halogens in Group 7A (17) can form more than two oxygen-
containing acids. For chlorine, the common form is chloric acid (HClO3), which contains the
chlorate polyatomic ion (ClO3 -). For the acid that contains one more oxygen atom than the
common form,
47
the prefix per is used; HClO4 is named perchloric acid. When the polyatomic ion in the acid has
one oxygen atom less than the common form, the suffix ous is used. Thus, HClO2 is named
chlorous acid; it contains the chlorite ion (ClO2 -). The prefix hypo is used for the acid that has
two oxygen atoms less than the common form; HClO is named hypochlorous acid (Timberlake,
C., & Timberlake, W., 2014).

Table 5.1 Nomenclature of some of the common acids.

B. Bases

You may be familiar with some household bases such as antacids, drain openers, and
oven cleaners. According to the Arrhenius theory, bases are ionic compounds that dissociate
into cations and hydroxide ions (OH-) when they dissolve in water. They are another example of
strong electrolytes. For example, sodium hydroxide is an Arrhenius base that dissociates
completely in water to give sodium ions (Na+) and hydroxide ions (OH-). Most Arrhenius bases
are formed from Groups 1A (1) and 2A (2) metals, such as NaOH, KOH, LiOH, and Ca(OH)2.
The hydroxide ions (OH-) give Arrhenius bases common characteristics, such as a bitter taste
and a slippery feel. A base turns litmus indicator blue and phenolphthalein indicator pink
(Timberlake, C., & Timberlake, W., 2014).

Table 5.2. Comparison of acids and bases.

48
Naming of Bases

Use the name of the metal element followed by hydroxide.

Table 5.3. Common bases

Chemical Formula Name


NaOH Sodium hydroxide
KOH Potassium hydroxide
Mg (OH)2 Magnesium hydroxide
Ca (OH)2 Calcium hydroxide

Definitions of Acids and Bases

1. Arrhenius's acid and base


In 1884, Arrhenius defined that an acid is a substance that gives H+ and a base one that
gives OH-. Namely, if an acid is HA and a base BOH, then
Equation 5.1 HA+BOH → A- B+ + H+ + OH
Therefore, when an acid and a base react, water is formed (Saito, 2010).

2. Brønsted-Lowry's acid and base


In a new theory submitted in 1923 independently by Brønsted and Lowry, an acid is defined
as a molecule or an ion which gives H+ and a molecule or ion that receives H+ A base is not
only molecule or an ion which gives OH- but anything which receives H+. Since the acid HA
gives H+ to water in an aqueous solution and generates an oxonium ion, H3O+, water is also a
kind of base according to this and generates an oxonium ion, H3O+, water is also a kind of
base (Saito, 2010).

Equation 5.2. HA (acid)+H2O (base)→H3O(conjugate acid)+A-(conjugate base)

Here H3O+ is called a conjugate acid and A- a conjugate base. However, since
water gives H+ to ammonia and generates NH4+, it is also an acid, as is shown below.

49
Equation 5.3 H2O (acid) + NH3 (base) → NH4+(conjugate acid) + OH - (conjugate base)

That is, water can be an acid or a base dependent on the co-reactant. Although the
definition of Brønsted-Lowry is not much different from that of Arrhenius for aqueous solutions,
it is more useful because the theory was extended to non-aqueous acids and bases (Saito,
2010).

3. Lewis Acids and Bases

A Lewis acid is any species (molecule or ion) that can accept a pair of electrons, and a
Lewis base is any species (molecule or ion) that can donate a pair of electrons. A Lewis acid-
base reaction occurs when a base donates a pair of electrons to an acid. A Lewis acid-base
adduct, a compound that contains a coordinate covalent bond between the Lewis acid and the
Lewis base, is formed (Lewis Acids and Bases, n.d.)

Strong and Weak acids

In the process called dissociation, an acid or a base separates into ions in water. The
strength of an acid is determined by the moles of H3O+ that are produced for each mole of acid
that dissociates. The strength of a base is determined by the moles of OH- that are produced for
each mole of base that dissolves. Strong acids and strong bases dissociate completely in water,
whereas weak acids and weak bases dissociate only slightly, leaving most of the initial acid or
base undissociated. Strong acids are examples of strong electrolytes because they donate H+
so easily that their dissociation in water is essentially complete. For example, when HCl, a
strong acid, dissociates in water, H+ is transferred to H2O; the resulting solution contains
essentially only the ions H3O+ and Cl-. We consider the reaction of HCl in H2O as going 100%
to products. Thus, one mole of a strong acid dissociates in water to yield one mole of H3O+ and
one mole of its conjugate base. We write the equation for a strong acid such as HCl with a
single arrow. HCl(g)
+ H2O(l) b H3O+(aq) + Cl-(aq) There are only six common strong acids, which are stronger
acids than H3O+. All other acids are weak (Timberlake, C., & Timberlake, W., 2014).

50
Table 5.4. Relative strengths of acids and bases.

Source: https://www.quora.com/What-are-some-examples-of-the-most-commonly-used-weak-acids

Lesson 2. Bronsted-Lowry Conjugate Acid and Base

Conjugate acids and bases

When a species donates a proton, it becomes the conjugate base; when a species gains
a proton, it becomes the conjugate acid. Conjugate acids and bases are in equilibrium in
solution. The form of the two forward and reverse reactions given above, both of which depend
on the transfer of a proton from an acid to a base, is expressed by writing the general Brønsted
equilibrium as shown in the figure below:

51
Figure 5.1. Conjugate acid base pairings in acetic acid and water reaction.

Source: https://sceweb.sce.uhcl.edu/wang/biolab/acid_base/5_conjugate_pairs.html

The species Base1 is called the conjugate base of Acid1 , and Acid2 is the conjugate
acid of Base2. The conjugate base of an acid is the species that is left after a proton is lost. The
conjugate acid of a base is the species formed when a proton is gained. Thus, F– is the
conjugate base of HF and H3O is the conjugate acid of H2 O. There is no fundamental
distinction between an acid and a conjugate acid or a base and a conjugate base: a conjugate
acid is just another acid and a conjugate base is just another base.

Writing Dissociation Constants of acids and bases

An acid dissociation expression, Ka, can be written for weak acids that gives the ratio of
the concentrations of products to the weak acid reactants. As with other dissociation
expressions, the molar concentration of the products is divided by the molar concentration of the
reactants. Because water is a pure liquid with a constant concentration, it is omitted. The
numerical value of the acid dissociation expression is the acid dissociation constant
(Timberlake, C., & Timberlake, W., 2014).

For Acids Ka:

Equation 5.4 HX(aq) + H2O(l) → H3 O(aq) +X– (aq)

Equation 5.5 Ka = [H3O+][X–]


[HX]
For bases Kb:

52
Equation 5.6 B(aq) + H2 O(l) → HB(aq) + OH(aq)

Equation 5.7 Kb = [HB ][ OH ]


[ B]

The strengths of Brønsted acids

The strength of a Brønsted acid is measured by its acidity constant, and the strength of a
Brønsted base is measured by its basicity constant; the stronger the base, the weaker is its
conjugate acid. Throughout this discussion, we shall need the concept of pH, which we assume
to be familiar from introductory chemistry: Since the concentrations of ions in aqueous solution
vary over an enormous range from greater than 1 molar down to 10-14 molar or less, it is
convenient to use a logarithmic function to describe them. For hydrogen ions in aqueous
solution, denoted by H3O+, often called the hydronium ion, this is defined as: pH = -log (a (H3O
+ ) ) -log(c( H3O+ )/mol dm-3 ) (Kealey & Haines, 2016).

Equation 5.8 pH= –log [H3O]

Equation 5.9 [H3O] = 10–pH

The strength of a Brønsted acid, such as HF, in aqueous solution is expressed by its
acidity constant (or ‘acid ionization constant’) Ka

Equation 5.10 HF+ H2 O →H3O+ +

F– Equation 5.11 Ka = [H3O+][F–]


[HF]

More generally:

Equation 5.12 Ka = [H3O+][X–]


[HX]
In this definition, [X– ] denotes the numerical value of the molar concentration of the
species X– (so, if the molar concentration of HF molecules is 0.001 mol dm–3, then [HF] 0.001).
A value Ka << 1 implies that [HX] is large with respect to [X– ], and so proton retention by the
acid is favored. The experimental value of Ka for hydrogen fluoride in water is 3.5 × 10–4,

53
indicating

54
that under normal conditions only a very small fraction of HF molecules are deprotonated in
water. The actual fraction deprotonated can be calculated as a function of acid concentration
from the numerical value of Ka (Atkins et al., 2010).

The strengths of Brønsted acids

The proton transfer equilibrium characteristic of a base, such as NH3, in water can also
be expressed in terms of an equilibrium constant, the basicity constant, Kb:

Equation 5.13 NH3 (aq) H2 O(l) →NH4+ (aq) + -OH(aq)

Equation 5.14 Kb = [NH4+][ -OH]


[NH3]
If Kb << 1, then [HB] << [B] at typical concentrations of B and only a small fraction of B
molecules are protonated. Therefore, the base is a weak proton acceptor and its conjugate acid
is present in low concentration in solution. The experimental value of Kb for ammonia in water is
1.8 × 10–5, indicating that under normal conditions, only a very small fraction of NH3 molecules
are protonated in water (Atkins et al., 2010).

Lesson 3. Acid and Base Equilibria


By omitting the constant concentration of pure water, we can write the water dissociation
expression, Kw.

Equation 5.15 Kw = [H3O+][OH-]

Experiments have determined, that in pure water, the concentration of H3O+ and OH- at
25 °C are each 1.0x 10-7 M. Pure water

Equation 5.16 [H3O+] = [OH-] = 1.0 x 10-7 M

When we place the [H3O+] and [OH-] into the water dissociation expression, we obtain
the numerical value of Kw, which is 1.0x 10 -14 at 25 °C. As before, the concentration units are
omitted in the Kw value.

Kw = [H3O+][OH-] = [1.0 x10-7 ][ 1.0 x10-7] = 1.0 x10-14

55
If we know the [H3O+] of a solution, we can use the Kw to calculate the [OH-]. If we
know the [OH-] of a solution, we can calculate [H3O+] from their relationship in the Kw.

Kw = [H3O+][OH-]

[OH-] = Kw_
[H3O+]

[H3O+] = Kw
[OH-]

Sample Problems:

1. A solution has a [OH-] = 4.0 x 10-12 M at 25 °C. What is the [H3O+] of the solution?

Step 1: Write the Kw

Kw = [H3O+][OH-]= 1.0 x10-14

Step 2: Derive the equation to solve for the unknown

[H3O+] = [H3O+][OH-]
[OH-]
[H3O+] = 1.0 x10-14
[4.0 x 10-12]

Step 3: Perform the operation

[H3O+] = 1.0 x10-14 = 2.5x10-3 M


[4.0 x 10-12]

pH= -log [2. 5x10-3 M]= 2.6

56
2. Calculate the [OH] of orange juice with the [H3O+] of 2.0 x10-4 M

Step 1: Write the Kw

Kw = [H3O+][OH-]= 1.0 x10-14

Step 2: Derive the equation to solve for the unknown

[OH-]= [H3O+][OH-]
[H3O+]

[OH-] = 1.0 x10-14


[2.0 x10-4]

Step 3: Perform the operation

[OH-] = 1.0 x10-14 = 5 x10-11 M


[4.0 x10-2]

pH= -log [5 x10-11 M]= 10.30

3.If the pH of Coke is 3.12, what is the concentration of [H3O+]?

pH = - log [H+]

Because pH = - log [H ] then +

- pH = log [H ] +

Take antilog (10 ) of both


x

sides and get

10 =[H ]
-pH +

[H ] = 10
+ -3.12 = 7.6 x 10 M
-4

*** to find antilog on your calculator, look for “Shift” or “2 function” and then the log button
nd

57
4. What is the pH of 0.0010 M NaOH solution?

[OH-] = 0.0010 (or 1.0 X 10-3 M)

pOH = - log 0.0010

pOH = 3

pH = 14 – 3 = 11

OR Kw = [H3O+] [OH-]

[H3O+] = 1.0 x 10-11 M


pH = - log (1.0 x 10-11) = 11.00

Assessment Task 2

1. Create a graphic organizer showing the characteristics of acids and bases.

2. List at least 10 common acids and bases.

3. Identify the conjugate acid and base in the following reactions:


a.
CH OH(aq) + H O(l) → ← CH O (aq) + H O (aq)
3 2 3

3
+

b. H PO (aq) + H O(l) → ← HPO (aq) + H O (aq)


2 4

2 4
2–
3
+

58
c. HS (aq) + H O(l) → ← S (aq) + H O (aq)

2
2–
3
+

4. Write the ionization constants of the reactions in No. 3

5. Calculate the following:


a.
What is the OH concentration of bile in aqueous solution with a [H3O+] of 7.9x10 M
-9

b.
What is the [H3O+] of aspirin in aqueous solution with a [OH] of 1.8x10 M
-11

59
Summary

The original distinction between acids and bases was based, hazardously, on criteria of
taste and feel: acids were sour and bases felt soapy. A deeper chemical understanding of their
properties emerged from Arrhenius’s (1884) conception of an acid as a compound that
produced hydrogen ions in water. The modern definitions that we consider in this chapter are
based on a broader range of chemical reactions. The definition due to Brønsted and Lowry
focuses on proton transfer, and that due to Lewis is based on the interaction of electron pair
acceptor and electron pair donor molecules and ions (Atkins, et. al., 2010).

60
References

Atkins, P.W. Overton, T.L. Rourke, J.P. Weller, M.T and. Armstrong, F.A. (2010). Shriver’s and
Atkin’s Inorganic Chemistry. 5th Edition.. Oxford University Press.
Fifield, FW., & Kealey, D. (2000). Principles and Practice of Analytical Chemistry. 2000.
Blackwell Science Ltd.
Harvey, D. (2016). Analytical Chemistry 2.1. (Electronic Version). McGraw Hill
Companies.
Kealey, D. & Haines, PJ. (2016). Analytical Chemistry. Instant Notes series. BIOS Scientific
Publishers Ltd.

Lewis Acids and Bases. (n.d.). https://opentextbc.ca/chemistry/chapter/15-2-lewis-acids-and-


bases/

Timberlake, C., & Timberlake W. (2014). Basic Chemistry. 4th Edition. 2014. Pearson Company

61
MODULE 6
Gravimetry and Titrimetry

Introduction

Gravimetry is one of the ‘classical’ techniques of analysis, and although less frequently
used now, it is of value when an accurate reference method is required for comparison with an
instrumental technique (Kealey & Haines, 2016). Titrimetry, in which volume serves as the
analytical signal, first appears as an analytical method in the early eighteenth century. Titrimetric
methods were not well received by the analytical chemists of that era because they could not
duplicate the accuracy and precision of a gravimetric analysis. Not surprisingly, few standard
texts from that era include titrimetric methods of analysis (Harvey, 2016).

Learning Outcomes

At the end of this module, students will be able to:

1. Differentiate titrimetry and gravimetry;

2. Create a concept map on the processes involved in titrimetry and gravimetry; and

3. Research on some applications of gravimetry and titrimetry.

61
Lesson 1. Titrimetry
Titrimetry is an analytical technique for the determination of the stoichiometry of a
reaction by the addition of controlled amounts of a standard reagent (Kealey & Haines, 2016).
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.
Unlike gravimetry, the development and acceptance of titrimetry required a deeper
understanding of stoichiometry, of thermodynamics, and of chemical equilibria. By the 1900s,
the accuracy and precision of titrimetric methods were comparable to that of gravimetric
methods, establishing titrimetry as an accepted analytical technique (Harvey, 2016).

1. Standard Solution
Titrations usually involve the addition of controlled volumes of a standard solution,
whose concentration is known accurately, to a solution of reactant of unknown concentration. In
order to achieve the highest accuracy, it is necessary to use well-established standard materials
as reagents in the primary calibration or standardization of the reacting solutions. The most
important of these are called primary standards, and should be easy to obtain, purify and dry,
should be stable and not hygroscopic, but should be readily soluble and react rapidly and
stoichiometrically. They should ideally have a high relative molecular mass to minimize weighing
errors. The above criteria mean that reagents such as sodium hydroxide, which is hygroscopic
and may react with carbon dioxide from the air, and potassium permanganate, which slowly
decomposes in air, are unsuitable as primary standards. Solutions used for quantitative analysis
need to be checked and calibrated frequently (Kealey & Haines, 2016).

2. Equivalence Points and Endpoints


Titrations usually involve the addition of controlled volumes of a standard solution,
whose concentration is known accurately, to a solution of reactant of unknown concentration.
The end point of a titration is based upon experimental observation, whereas the equivalence
point is the theoretical value dependent upon the reaction equation. In an ideal case, these
should be the same, but a check may be needed to ensure that factors such as blank errors do
not affect the
62
results. In any titration, the end point corresponds to rapid changes in the concentration of
species (Kealey & Haines, 2016).
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.

Equation 6.1 moles titrant = MT x V eq

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

3. Indicators
Titrations usually involve the addition of controlled volumes of a standard solution,
whose concentration is known accurately, to a solution of reactant of unknown concentration. In
general, indicators have two forms, which possess different colors. Indicators for acid-base
titrations are themselves weak acids or bases where the two forms differ in colors. The choice of
indicator depends upon the reaction to be studied. The equilibrium constant of the indicator
must match the pH range, or electrode potential range of the species being titrated. Table 6. 1
shows a selection of indicators for acid-base reactions. As a general rule it is noted that the
color change takes place over the range (Haines, Kealey, 2016).
pH = pKIn ± 1

63
Similarly for other reactions, the concentration at which the indicator changes must
match the concentrations at the end point. For example, in the titration of the weak dibasic
maleic acid with sodium hydroxide, the first end point, corresponding to sodium hydrogen
maleate occurs at pH = 3.5, while the second end point for disodium maleate is at pH = 9. Since
pKIn = 3.7 for methyl orange, this will change at the first end point, while phenolphthalein would
change color around pH = 9.

Table 6.1. Indicators and the pH range and color change.

Source: https://www.chemedx.org/activity/indicator-activity-assigning-roles-students-laboratory

4. Titration Curves

Titration curve provides a visual picture of how a property of the titration reaction changes as
we add the titrant to the titrand. 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 (Harvey, 2016).

64
Figure 6.1. Titration curves.

Source: https://www.chemguide.co.uk/physical/acidbaseeqia/phcurves.html

Lesson 2. Gravimetry
Gravimetry includes all analytical methods in which the analytical signal is a
measurement of mass or a change in mass. When you step on a scale after exercising you are,
in a sense, making a gravimetric determination of your mass. Mass is the most fundamental of
all analytical measurements and gravimetry unquestionably is the oldest quantitative analytical
technique. If an element is present in a mixture, for example, silver in a sample of nickel, one
way of separating it is to dissolve the metal completely in a suitable solvent. In this example, the
metal mixture could be dissolved in concentrated nitric acid and a reagent added that would
react with the silver to produce a precipitate, which for silver might be a sodium chloride
solution: Ag(s) + Ni(s) + HNO3(sol) = AgNO3(sol) + Ni(NO3)2(sol) AgNO3(sol) + NaCl(sol) =
AgCl(s) + NaNO3(sol) The
65
silver chloride is precipitated completely, and may be filtered off since both nickel nitrate and
nickel chloride are very soluble in water. The precipitate will be wet and may contain traces of
nickel in solution, so must be thoroughly washed and dried, as discussed below. Since weighing
may be carried out readily and accurately in almost all laboratories, gravimetry is often used as
a reference method. Analysis of major components of metal samples such as steel, and of
minerals and soils may be carried out by gravimetric methods, but they often involve lengthy
separations and are consuming. Newer instrumental methods may determine several
components simultaneously, rapidly and are generally applicable down to trace levels (Harvey,
2016).

Gravimetric Processes

1. Precipitation

Many elements will form compounds insoluble in water or other solvents. Provided the
compound is stable, or may be converted into a stable compound easily, these insoluble
precipitates may be used for analysis.

The technique for obtaining a precipitate may be summarized as follows according to


Kealey & Haines (2016).:

(i) The sample should be dissolved as completely as possible in a suitable solvent.


Any residue that does not dissolve (for example, silica present in the metal
sample of the above example), may be filtered off at this stage.
(ii) Unless there are undesirable changes when the sample is heated, the solutions
should be warmed. This speeds up reactions and helps to form a more granular
precipitate.
(iii) The precipitating reagent must be chosen to give as insoluble a precipitate as
possible. Preferably, a reagent that will produce the largest mass of precipitate
should be used. For example, aluminum may be precipitated and heated to give
the oxide, Al2O3 when 10.0 mg of aluminum will produce 18.9 mg precipitate. If
‘aluminon’ (8-hydroxyquinoline, C9H6ON) is used, 170.0 mg of precipitate
results.
(iv) The precipitating reagent should be added slowly, with stirring to the warm
66
solution. To check whether precipitation is complete, the precipitate is allowed to

67
settle, and more reagent added. If further precipitate does not form, the reaction
is complete. If the solution appears cloudy, it is possible that a colloidal form of
the solid is present. This may be coagulated by further warming or adding more
reagent.
(v) The reaction mixture is filtered. Various means may be used for this. The
simplest is a quantitative filter paper (ashless), which has been dried and
weighed previously. These may be dried, or burnt (‘ashed’). Another useful filter
is a sintered glass or porcelain crucible, dried and weighed as before. Glass will
withstand heating to about 300∞C and porcelain to over 800∞C.

2. Weighing

Modern balances can readily weigh samples directly, and masses from several grams
down to a few micrograms can be weighed accurately and quickly. It is important that the
conditions are the same for the initial weighing (crucible, filter paper) as for the final weighings.
Temperature is especially important and hot samples should never be placed directly onto a
balance pan (Harvey, 2016).

Types of Gravimetric Methods

The examples in the previous section illustrate four different ways in which a
measurement of mass may serve as an analytical signal. When the signal is the mass of a
precipitate, we call the method precipitation gravimetry. The indirect determination of PO3 3- by
precipitating Hg2Cl2 is an example, as is the direct determination of Cl– by precipitating AgCl. In
electrogravimetry, we deposit the analyte as a solid film on an electrode in an electrochemical
cell. The deposition as PbO2 at a Pt anode is one example of electrogravimetry. The reduction
of Cu2+ to Cu at a Pt cathode is another example of electrogravimetry. When we use thermal or
chemical energy to remove a volatile species, we call the method volatilization gravimetry. In
determining the moisture content of bread, for example, we use thermal energy to vaporize the
water in the sample. To determine the amount of carbon in an organic compound, we use the
chemical energy of combustion to convert it to CO2. Finally, in particulate gravimetry we
determine the analyte by separating it from the sample’s matrix using a filtration or an extraction
(Harvey, 2016).

68
Assessment Task

1. Differentiate gravimetry and titrimetry methods

2. Create a graphic organizer showing how gravimetry and titrimetry are performed.

3. Conduct a research on the specific of the applications of gravimetry and titrimetry.

69
Summary

Gravimetry is the analytical technique of obtaining a stable solid compound, of known


stoichiometric composition so that the amount of an analyte in the sample may be found by
weighing mass of the sample. 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.

70
References

Harvey, D. (2016). Analytical Chemistry 2.1. (Electronic Version). McGraw Hill


Companies.
Kealey, D., Haines, PJ. (2016). Analytical Chemistry. Instant Notes series. BIOS Scientific
Publishers Ltd.

71

You might also like