You are on page 1of 342

GENERAL CHEMISTRY

Third Edition

J. Ahmad
G. Kumar
I. P. Meulenberg
G.S. Singh
This work is licensed under the Creative Commons Attribution 3.0 Unported
License. To view a copy of this license, visit
http://creativecommons.org/licenses/by/3.0/ or send a letter to Creative
Commons, 444 Castro Street, Suite 900, Mountain View, California, 94041,
USA.

The virtual textbook Chem1 General Chemistry Virtual Textbook by


Stephen Lower available here:
http://www.chem1.com/acad/webtext/virtualtextbook.html under the
Creative Common License 3.0 was an important source of ideas for
writing several parts of this book. Some diagrams and tables have been
taken from this e-book.
Among other books and sources consulted while writing this work were: Chemistry
the Central Science by T. L.Brown, H. E. LeMay. Jr., B. E. Bursten, C. J. Murphy;
Chemistry: Molecules, Matter and Change by L. Jones and P. Atkins; Chemistry,
Matter and Its Changes by J. E. Brady and F. Senes; General Chemistry by R.
Chang; and General Chemistry by D. D. Ebbing and S. D. Gammon.

Published by Chemistry Department, University of Botswana, Private Bag


00704, Gaborone, Botswana. 2013.
PREFACE to 3rd Edition

The teams teaching introductory chemistry at the University of Botswana over the
years have felt a need for a textbook that is more aligned to the curriculum
compared to the available textbooks in the market. A typical textbook available
these days is an encyclopedic tome running to over a thousand pages. Even if the
high cost of such a book to students of a middle-income country such as Botswana
were not a factor, the sheer range of topics covered is daunting. The vast scope of
these books is understandable in that their authors want to cater to the diversity of
approaches to the subject followed in various institutions.

This concise and affordable book was written to fulfil that need. It has been made
available to the students taking the two general chemistry courses in the
University, with other books now being additional recommended reading. Each
chapter contains solved examples and ends with exercises. Every attempt has been
made to present material concisely in clear language and is augmented with figures
where necessary.

This book will cater to the needs of undergraduate students taking general
chemistry courses (General Chemistry-I and General Chemistry-II) of the
University of Botswana and other institutions offering similar courses. The first
few chapters deal with fundamental concepts such as structure of the atom, moles,
stoichiometry, periodicity in the properties of the elements, and quantum numbers.
The latter half of the book covers such phenomenological areas as
thermochemistry, solutions, chemical kinetics, chemical equilibrium, and an
introductory treatment of carbon-containing compounds.

J. Ahmad
G. Kumar
I. P. Meulenberg
G.S. Singh
Contents
PREFACE TO 3RD EDITION 7
CHAPTER 1 8
INTRODUCTION: MATTER AND MEASUREMENT 8

1.1 Some Basic Concepts 8

1.2 Units of Measurement 12


1.3 Conversion of Units 15
1.4 Uncertainty in Measurements 17
1.5 Exercises 20

CHAPTER 2 23
ATOMS, MOLECULES AND IONS 23

2.1 Atomic Structure 23


2.2 Atomic Weight 26
2.3 The Periodic Table 27
2.4. Molecular Compounds 29
2.5 Ionic Compounds 30
2.6 Naming of Inorganic Compounds 34
2.7 Exercises 39

CHAPTER 3 42
STOICHIOMETRY: CHEMICAL CALCULTIONS 42

3.1 Chemical Reaction Equations 42


3.2 Formula Weights; Percentage Composition 44
3.3 The Mole and Molar Mass 46
3.4 Empirical Formulas 50

1
3.5 Calculations with Balanced Reaction Equations 53
3.6 Limiting Reactants; Reaction Yield 54
3.7 Exercises 57
CHAPTER 4 60
REACTIONS IN AQUEOUS SOLUTION 60

4.1 Introduction to Aqueous Solutions 60


4.2 Precipitation Reactions 61
4.3 Acid Base Reactions 65
4.4 Oxidation-Reduction Reactions 68
4.5 Solution Concentration 75
4.6 Calculations with Reactions in Aqueous Solution 77
4.7 Exercises 78
CHAPTER 5 81
GASES 81

5.1 Pressure, Pressure Units 81


5.2 Simple Gas Laws 81
5.3 The Ideal Gas Law 84
5.4 Applications of the Ideal Gas Law 85
5.5. Mixtures of Gases 87
5.6 Molecular Motion: Effusion 90
5.7 Kinetic Theory of Gases 92
5.8 Exercises 92
CHAPTER 6 95
ELECTRONIC STRUCTURE OF ATOMS 95

6.1 Electromagnetic Radiation 95

2
6.2 Photons; Quantization of Energy 96
6.3 Line Spectrum of Hydrogen Atom; the Bohr Model 98
6.4 Wave Nature of the Electron 101
6.5 The Quantum Mechanical Model of the Atom 103
6.6 Electron Configuration of Atoms 107
6.7 Electron Configurations of Monoatomic Ions 112
6.8 Exercises 112
CHAPTER 7 115
PERIODIC PROPERTIES OF THE ELEMENTS 115

7.1 Electronic Configuration and the Periodic Table 115


7.2 Size of Atoms and Ions 119
7.3. Ionization Energy 123
7.4 Electron Affinity 126
7.5 Exercises 128

CHAPTER 8 129
BASIC CONCEPTS OF CHEMICAL BONDING 129

8.1 Ionic Bonds 129


8.2 Covalent Bonds 132
8.3 Resonance and Resonance Structures 137
8.4 Formal Charge 139
8.5 Electronegativity and Bond Polarity 141
8.6 Exceptions to the Octet Rule 142
8.7 Exercises 144

CHAPTER 9 146

3
MOLECULAR GEOMETRY 146

9.1 Valence-Shell Electron Pair Repulsion Model 146


9.2 Molecular Shapes 147
9.3 Molecular Shape and Polarity 154
9.4 Exercises 157

CHAPTER 10 158
INTRODUCTION TO ORGANIC CHEMISTRY 158

10.1 Carbon-based Materials 158


10.2 Structural Features of Organic Compounds 160
10.3 A Simple Classification of Organic Compounds 164
10.4 Hydrocarbons 165
10.5 Compounds containing Oxygen and Nitrogen Atoms besides
Carbon and Hydrogen: Functional Groups 190
10.6 Exercises 194

CHAPTER 11 197
THERMOCHEMISTRY 197

11.1 Introduction: the Basics of Energy, Heat and Work 197


11.2 Endothermic and Exothermic Processes 198
11.3 Calorimetry 200
11.4 Enthalpy 203
11.5 Hesss Law and its Applications 207
11.6 Bond Enthalpies 213
11.7 Exercises 216

CHAPTER 12 218
4
SOLUTIONS 218

12.1 Introduction to Solutions 218


12.2 Concentration; Concentration Units 219
12.3 Solution Types; Why Solutions Form (or Dont) 226
12.4 Solubility of Gases 229
12.5 Solubility of Solids 231
12.6 Colligative Properties 235
12.7 Osmotic Pressure and Osmosis 244
12.8 Exercises 251

CHAPTER 13 253
THE RATES OF CHEMICAL REACTIONS 253

13.1 Reaction Rate 253


13.2 Rate Laws and Reaction Order 256
13.3 Determining Reaction Orders 258
13.4 Differential and Integral Rate Laws 260
13.5 Temperature and Rate 267
13.6 THEORIES OF REACTION RATES 269
13.7 CATALYSTS 274
13.8 ENZYMES 275

13.9 Exercises 275

CHAPTER 14 278
INTRODUCTION TO CHEMICAL EQUILIBRIUM 278

14.1 Chemical Equilibrium 278


14.2 The Equilibrium Constant 279

5
14.3 Heterogeneous Equilibrium 283
14.4 Calculations with Kc and Kp 285
14.5 Le Chtelier Principle 291
14.6 Exercises 298

CHAPTER 15 300
ACIDS AND BASES 300

15.1 What is an Acid and what is a Base? 300


15.2 Acidity of a Solution: pH 305
15.3 Calculations involving Ka and pH 309
15.4 Titrations 313
15.5 Exercises 317

CHAPTER 16 318
AQUEOUS EQUILIBRIA 318

16.1 pH of Salt Solutions 318


16.2 Buffers 322
16.3 Solubility Equilibria 324
16.4 Exercises 326

INDEX 332

6
PREFACE to 3rd Edition

The teams teaching introductory chemistry at the University of Botswana


over the years have felt a need for a textbook that is more aligned to the
curriculum compared to the available textbooks in the market. A typical
textbook available these days is an encyclopedic tome running to over a
thousand pages. Even if the high cost of such a book to students of a
middle-income country such as Botswana were not a factor, the sheer range
of topics covered is daunting. The vast scope of these books is
understandable in that their authors want to cater to the diversity of
approaches to the subject followed in various institutions.

This concise and affordable book was written to fulfil that need. It has
been made available to the students taking the two general chemistry
courses in the University, with other books now being additional
recommended reading. Each chapter contains solved examples and ends
with exercises. Every attempt has been made to present material concisely
in clear language and is augmented with figures where necessary.

The authors wish to express their appreciation to their colleagues in the


Department of Chemistry who reposed confidence in them and encouraged
the writing of the book. Special mention should be made of the following:

J. Ahmad
G. Kumar
I. P. Meulenberg
G.S. Singh

7
CHAPTER 1
Introduction: Matter and Measurement
1.1 Some Basic Concepts
Matter, classification of matter
Chemistry studies the properties and structure of matter and the
changes that matter can undergo. Matter is usually defined as
anything that has mass and takes up space. The book you are reading,
the air you are breathing and the food you eat, are all examples of
matter.
Matter can be classified according to its physical state (whether it is a
gas, a liquid, a solid), or its composition (whether it is an element, a
compound, a mixture).

Physical states of matter


We distinguish between three physical states of matter (see fig. 1.1):
A gas (vapour): has no fixed shape or volume. Molecules in a
gas are far apart and move at high speeds, colliding with each
other and objects around them.
A liquid: has a distinct volume, but no fixed shape. Molecules in
a liquid are packed closely together but move rapidly.
A solid: has a definite shape and volume. Molecules in a solid are
held tightly together.

Figure 1.1: Arrangement of molecules in a solid, liquid and gas

Composition of matter
All matter is either a (pure) substance or a mixture. Most matter we
meet in everyday life are mixtures; for example many of our drinks,
our food, the air we breathe are all mixtures. A mixture is not pure,

8
but contains various substances mixed together. It can be separated by
physical means to give its components. (Physical means are
methods which use a difference in physical property for separation,
like a difference in boiling point in the case of distillation.)
Components of a mixture have retained their chemical identity. For
example, in a solution of common sugar (sucrose, C12H22O11) in water
(H2O), both substances are still present as water and sucrose
molecules.
Note: Mixtures are either homogeneous (uniform throughout, even at a
microscopic level) or heterogeneous (not uniform throughout). A
homogeneous mixture is also called a solution, while a suspension is
an example of a heterogeneous mixture.
In contrast to a mixture, a substance is a single pure form of matter. It
cannot be separated by physical means and has a set of properties
which are distinct from those of other substances. Water, sucrose,
sodium chloride (NaCl) and oxygen gas (O2) are examples of
substances.

Substances are either elements or compounds. An element is


composed of only one kind of atom and cannot be separated further by
chemical means. (Chemical means are methods which use chemical
reactivity to decompose a substance, for example electrolysis.)
Sodium (Na), carbon (C), hydrogen (H2), oxygen (O2), iron (Fe) and
chlorine (Cl2) are examples of elements.
A compound is a substance which contains different kinds of atoms; it
is made of two or more elements. A compound can be divided further
by chemical means, i.e. it can be decomposed into simpler compounds
or its constituent elements. Water, sodium chloride and sucrose are
examples of compounds.
Matter

Substance Mixture

Element Compound Homogeneous Heterogeneous

Figure 1.2: Classification of matter

9
Elements
An element is a fundamental substance from which all matter is built.
The smallest particle of an element is an atom. A few elements, the
noble gases like helium and neon, exist naturally as individual atoms.
Some others occur naturally as molecules, for example the halogens
(Cl2, Br2), oxygen (O2), nitrogen (N2), phosphorus (P4), and sulphur
(S8). (A molecule is a particle consisting of two or more atoms, of the
same or different elements, chemically bound together.) Many
elements exist as a giant lattice, a huge regular arrangement of atoms:
carbon, silicon, metals.
Most elements are solids under room conditions, several are gases
(noble gases, halogens, hydrogen, nitrogen, oxygen), while only two
(bromine, mercury) are liquids. The Periodic Table shows all known
elements, arranged in a special way (see chapter 2).
Table 1.1: Names and symbols for selected elements
Name - Symbol Name - Symbol Name - Symbol
Actinium Ac Copper Cu Oxygen O
Aluminium Al Fluorine F Phosphorus P
Antimony Sb Gold Au Platinum Pt
Argon Ar Helium He Potassium K
Arsenic As Hydrogen H Scandium Sc
Barium Ba Iodine I Selenium Se
Beryllium Be Iron Fe Silicon Si
Bismuth Bi Krypton Kr Silver Ag
Boron B Lanthanum La Sodium Na
Bromine Br Lead Pb Strontium Sr
Cadmium Cd Lithium Li Sulphur S
Calcium Ca Magnesium Mg Tin Sn
Carbon C Manganese Mn Titanium Ti
Cerium Ce Mercury Hg Uranium U
Cesium Cs Molybdenum Mo Vanadium V
Chlorine Cl Neon Ne Xenon Xe
Chromium Cr Nickel Ni Zinc Zn
Cobalt Co Nitrogen N Rubidium Rb
_____________________________________________________________________________________________________

Each element has a unique name and chemical symbol. This chemical
symbol consists of 1 or 2 letters; the first letter is always a capital
letter, the second letter is always a small letter. Table 1.1 shows the
names and symbols of selected elements.
10
Example 1.1:
Carbon = C, potassium = K, sodium = Na, gold = Au
Be careful not to write the second letter as a capital: silicon = Si; not SI,
which indicates a combination of the elements sulphur, S, and iodine, I.

Note: Most General Chemistry text books give interesting information


about the derivation of the names and symbols of various elements.

Compounds
There are many more compounds than elements. A compound is a
substance consisting of two or more elements in a definite mass ratio
(Law of Constant Composition). No matter its source or how it was
made, a compound has always the same composition. For example,
NaCl obtained from seawater, a salt mine or prepared in the lab
consists of 39.34% Na and 60.66% Cl. Classification and
nomenclature of compounds are discussed in chapter 2.

Chemical formulas
In chemistry, a substance is represented by its chemical formula. It
shows the number and kind of element(s) present in the basic particle
of the substance (but see ionic compounds in chapter 2). For example,
hydrogen gas consists of hydrogen molecules in which two hydrogen
atoms are chemically bound together. The chemical formula of
hydrogen gas is the formula of the molecule, H2; the subscript 2
indicates that 2 hydrogen atoms have combined to form a hydrogen
molecule. The chemical formula of carbon dioxide gas is CO2; it
consists of molecules in which 1 carbon atom is combined with 2
oxygen atoms.
It was mentioned before that noble gases exist naturally as isolated
atoms. Their chemical formula is the symbol of the atom, for
example, argon gas is Ar. Although elements like carbon and the
metals form a giant lattice in which many atoms are linked, their
chemical formula is simply the symbol of the element; for example,
the chemical formula for graphite, a form of carbon, is C, and that for
iron is Fe.

11
Properties of substances
Its properties characterise and identify a substance. Some properties
could be the same as, or very similar to, those of other substances;
however, there are always some which are different and unique to that
particular substance.
Properties can be divided into physical and chemical properties:
Physical properties are observed without changing the chemical
identity of a substance, e.g. melting point. (Determining the melting
point of water involves a change of state from solid to liquid only, not
a chemical change of water H2O molecules remain H2O molecules.)
Chemical properties are observed by changing the chemical identity of
the substance they concern the chemical reactions of a substance,
e.g. reactivity with oxygen. (Investigating the reactivity of the metal
magnesium with oxygen implies that Mg could react with O2 and
change to magnesium oxide, MgO.)
Note: A physical change is a change of state. A chemical change is the
reaction of one substance to give another substance or substances.

Properties can also be divided into intensive and extensive properties:


Intensive properties do not depend on the amount of substance. They
are used to identify a substance. Density is an example of an intensive
property. The density of water is the same for 1 mL of water and
1000 L of water, if both volumes are under the same conditions.
Extensive properties depend on the amount of substance, for example
volume. A larger amount of water implies a larger volume of water.

Example 1.2:
Colour is a physical property but also an intensive property.
Mass is a physical property as well as an extensive property.
Reactivity with water is a chemical property and an intensive property.

1.2 Units of Measurement


Besides qualitative observations (does a reaction take place, what
is the colour and texture of the precipitate), experimental chemistry
involves quantitative measurements. A measured quantity includes a
number and a unit, e.g. a mass of 25.46 g.
12
SI Units
The units used in science are called SI units, which are based on the
metric system. There are 7 single base units from which all other SI
units are derived (table 1.2).
Table 1.2: The SI base units
Physical property Unit Abbreviation
Mass kilogram kg
Length metre m
Time second s (sec)
Temperature Kelvin K
Amount of substance mole mol
Electric current Ampere A
Luminous intensity candela cd

Note that some non-SI units are still widely used. For example, the
unit degree Celsius (oC) is used for temperature. The English
system of units is still in common use in the USA.

Prefixes
Prefixes are often added to units in order to give more convenient
numbers. They are used to indicate decimal fractions or multiples of
units (see table 1.3). Examples of units are discussed in the following
sections.
Table 1.3: SI prefixes
Prefix Name Meaning
p pico 10-12
n nano 10-9
micro 10-6
m milli 10-3
c centi 10-2
d deci 10-1
k kilo 103
M mega 106

13
Temperature
Units used for temperature are degrees Celsius (oC) and Kelvin (K).
Kelvin is the SI unit, but oC is widely used. The relationship between
these two is:
(K) = (oC) + 273.15 or (oC) = (K) 273.15
Absolute zero is a temperature of 0 K (= - 273.15oC).
The Kelvin scale and degrees Celsius scale have the same size of unit;
a temperature difference in Kelvin is the same as a temperature
difference in degrees Celsius:
T (in K) = T (in oC) ( = difference)

Example 1.3:
A temperature of 25oC will be (25 + 273.15) K = 298 K.
A temperature of 400 K is the same as (400 273.15)oC = 127oC.

Derived SI units: volume, density


Derived SI units are units of properties other than those used to define
base units. Examples in chemistry are volume and density.
Volume (V) can be given as length length length. Its SI unit is m3.
However, this unit is not very convenient and more often used in
chemistry are dm3, cm3 and the non-SI units L, mL. Their relationship
is:
1 dm3 = 1 L = 1000 cm3 = 1000 mL
Density (d) of a substance is mass (m) per unit volume (V):
mass m
density or d
volume V
Units used are usually g/cm3 (= g/mL) for solids and liquids, and
g/dm3 (= g/L) for gases. Density depends on temperature (and for
gases also on pressure); temperature should therefore be specified
when reporting density data.
The expression for density can be rearranged to give:
m
mdV and V
d

14
Example 1.4:
The density of a piece of pure iron with a mass of 117.4 g and volume of
14.9 cm3 is
m 117.4 g
d 7.87 g / cm3
V 14.9 cm3
We included the units in our calculation - you are strongly advised to do
the same. Units are a first indication of a possible error in a calculation.
In the example above, we expected the units for our answer to be units for
density and these were the units we obtained. If mass and volume were
exchanged, we would have obtained the unit cm3 / g, clearly not a density
unit.

Example 1.5:
The mass of 25.0 cm3 of ethanol which has a density of 0.789 g/mL at
20oC is
m = d V = 0.789 g/mL 25.0 mL = 19.7 g
(Volume and density have different volume units, but 1 cm3 = 1 mL.)

Example 1.6:
The density of Br2(l) = 3.12 g/cm3 at 20oC. Volume of liquid bromine that
is required to obtain a mass of 55.0 g is
m 55.0 g
V 17.6 cm3
d 3.12 g/cm3
Note that g/cm3 is the same as g cm-3, or g/mL is the same as g mL-1.

1.3 Conversion of Units


It is regularly required to express the value for a measurement or
property in different units. For example, the density obtained from a
data book might be given in kg/m3 while we require it in g/cm3. In a
unit conversion, we express the same quantity in different units, but
do not change its intrinsic value.
We will use the method of conversion factors:
unit required = unit given conversion factor
unit required
so that the conversion factor = .
unit given

15
We should know the relationship between unit given and unit required
in order to set up the conversion factor. For example, from our
knowledge of prefixes we can state that 1 mg = 10-3 g; possible
conversion factors are then
mg 10 -3 g
and .
10 -3 g mg
Note that the value of a conversion factor is 1.

Example 1.6:
Conversion of a mass of 4.8 g to mg:
The relationship between g and mg and the possible conversion factors
were given above. Since we have to change g to mg we choose the
conversion factor which has g in its denominator.
mg
4.8 g 4.8 g 4.8 103 mg
-3
10 g
A conversion often requires multiple conversion factors, as shown below.

Example 1.7:
A volume of 15 mm3 expressed in dm3:
Required unit relationships are 1 mm = 10-3 m and 1 dm = 10-1 m. Then,

3 3 (10-3 m)3 dm3


15 mm 15 mm
mm3 (10 1 m)3
15 10-9 mm3 m3 dm3
1.5 10-5 dm3
10-3 mm3 m3
We treated units as numbers: units which appeared in both numerator and
denominator were cancelled.
Note that for the conversion of a unit in the denominator, the conversion
factor will be the inverse of that for a unit in the numerator.

Example 1.8:
Conversion of a speed of 10 m/s to km/hr:
The relevant unit relationships are: km = 103 m, hr = 60 min, min = 60 s,
so that

16
10 m 10 m km 60 s 60 min 10 60 60 m km s min

s s 103 m min hr 103 s m min hr
36 km/hr

1.4 Uncertainty in Measurements


Exact and inexact numbers
Some of the numbers we deal with in science are exact numbers,
others are inexact numbers. As the name implies, the value of an
exact number is exactly known. Exact numbers are
counted numbers (32 books on the shelf, 141 students in class),
numbers in definitions (1 dozen = 12; 1 m = 1000 mm).
Inexact numbers have an uncertainty. They are
measured values (mass and volume of a certain amount of water),
calculations based on measured values (the density of water
calculated from the measured mass and volume).

Experimental uncertainty
Quantitative measurements have an uncertainty. This uncertainty has
a human cause (e.g. the way the instrument is operated, estimation of
the last digit on the scale) and an instrumental cause (the way the
device is manufactured, electronic noise). Each instrument has
therefore an uncertainty value (or tolerance), which is indicated on the
device or included in the manufacturers manual. This uncertainty
value:
determines the number of digits in a reading,
gives the range within which the true value for the reading can
be found.
The uncertainty value normally implies that the reading should have
one more digit than the smallest scale division.

Example 1.9:
The 25 mL measuring cylinder has an uncertainty of 0.3 mL.
A reading will then be taken to 1 decimal place, like 14.7 mL. The true

17
value of the liquid volume contained in the cylinder will be in the range
14.7 + 0.3 mL = 14.4 15.0 mL. In the reading, the numbers 1 and 4 are
certain, the 7 is uncertain.

Significant figures
The + notation can be used to indicate the uncertainty in a measured
value. However, very often the concept of significant figures is used
for that purpose.
Measured values should be reported with (only) the last digit
uncertain, all preceding digits are certain. Digits reported are called
significant figures. The rules for counting significant figures are:
All non-zero digits are significant.
5.37 cm has 3 significant figures.
Zeros between non-zero digits are significant.
1.002 g has 4 significant figures.
Zeros to the right of the decimal point at the end of the number
are significant.
A volume of 8.00 mL has 3 significant figures, 8.0 mL has 2 significant
figures, while 8 mL has only 1 significant figure. Note that these three
volumes are not the same but differ in precision.
Zeros preceding the first non-zero digit in a number are not
significant - these indicate the decimal point. (To avoid
confusion, it is advisable to rewrite the number in scientific
notation: 1 10 10n)
0.00751 g has 3 significant figures (7.51 10 -3); 0.751 also has 3
significant figures (7.51 10-1).
Note: The value 7500 is ambiguous, it could have 2 (7.5 103), 3
(7.50 103), or 4 significant figures (7.500 103). It is
therefore advisable to report a value in scientific notation.

Significant figures in calculations


Naturally, calculated values which result from measurements also
have an uncertainty. In reporting such values to the correct number of
significant figures, we often have to round. We will use the following
rules for rounding numbers:

18
round up if the last digit dropped is 5 or above.
16.750 g rounded to 1 decimal place = 16.8 g
round down if the last digit dropped is below 5.
6.94 g/mL rounded to 2 significant figures = 6.9 mL
Rounding is done (only once) in the final step of the calculation.

The following rules apply for reporting a calculated value to the


correct number of significant figures:
In addition and subtraction, round the result to the smallest
number of decimal places in the data.
In multiplication and division, round the result to the smallest
number of significant figures in the data.

Examples 1.10:
36.95 g - 2.366 g = {34.584 g} = 34.58 g (2 decimal places)
13.56 g + 0.9447 g = {14.5047 g} = 14.50 g (2 decimal places)
To avoid confusion, it is convenient to clearly indicate the non-significant
figures, for example as subscripts (see examples above or below).
125 g
142.2556 mL 142 mL (3 significan t figures)
0.8787 g mL-1
(45.1755 - 43.765) g 1.4105 g
2.861.... g/L 2.86 g/L
(0.150 0.3430) L 0.4930 L
(3 significan t figures)

Precision and accuracy of measurements


The same measurement is often repeated because the average value is
more reliable than any individual measurement. Precision indicates
the agreement between multiple measurements (how close are the
repeated measurements together). It is determined by the spread in
the results. Average deviation of the individual results from the
average value (see the Lab Manual), or better, the standard deviation
in the average value are used to give information about the precision.

19
Accuracy indicates the agreement with the true (or accepted) value
(how close is the average to the true value). Certain statistical
methods can be applied to obtain an idea about the accuracy if the
actual true value is not known. Such a discussion is however outside
the scope of this book.

1.5 Exercises
Elements, compounds, mixtures
1.1 Indicate if the following statements are true or false. If false, correct
the statement.
(a) Homogeneous mixtures are also known as suspensions.
(b) A solution is an example of a compound.
(c) Bromine is an element since it exists naturally as molecules.
(d) A pure substance consists of the same kind of atoms.

1.2 By representing atoms of different elements as differently shaded or


coloured balls, sketch
(a) a gaseous mixture of argon (Ar), oxygen (O2), and carbon
dioxide (CO2),
(b) a gaseous mixture of nitrogen (N2), water vapour (H2O) and neon
(Ne).

1.3.A Give the name or chemical symbol, where appropriate, for the
following elements:
(a) calcium (b) Zn (c) phosphorus (d) F
(e) neon (f) Br (g) potassium (h) S

1.3.B Write the name or chemical symbol, where appropriate, for the
following elements:
(a) copper (b) Mg (c) nitrogen (d) Cl
(e) argon (f) Li (g) silicon (h) Ag

1.4 Classify the following properties as intensive or extensive properties,


and chemical or physical properties:
(a) boiling point, density, volume, reactivity with H2,
(b) mass, solubility in water, reactivity with O2, melting point.

Units, Conversion of units


1.5.A The average speed, u, of molecules in a gas can be found using the
equation u2 = 3RT/M. If the units for u = cm/s, for M = g/mol, and
for T = K, what are the units for R?
20
1.5.B The Ideal Gas Law, PV = nRT, gives the relationship between
pressure, temperature, volume and number of moles of an ideal gas.
What are the units for R if the units for P = atm, for V = L, for n =
mol, and for T = K?

1.6.A Perform the following unit conversions.


(a) 2.70 g/mL to kg/m3 (b) 5.0 ng/cm3 to g/L
(c) 14.8 km/hr to dm/s (d) 3.96 mm/s to pm/s
2 2
(e) 6.50 mL/s to L/hr

1.6.B Express:
(a) 1.02 g/cm3 in kg/L, (b) 9.68 cm/s in m/hr,
(c) 3.75 g/dm3 in ng/mL, (d) 2.45 mg/ml in kg/m3.
(e) 6.4 mg/m2 in g/nm2

1.7.A The value of gemstones like diamond is determined by its colour,


clarity, cut, and weight. The weight of a diamond is expressed in
carat (1 carat = 200 mg). What is the volume of a 35 carat diamond
if its density is 3.52 g/cm3?

1.7.B In gold trading, the weight of gold is expressed in troy ounce (1 troy
ounce = 31.1035 g). What is the volume of a piece of gold of 25.0
troy ounce? (The density of gold = 19.32 g/cm3.)

Significant Figures
1.8.A What is the number of significant figures in the following
experimental values?
(a) 7.99 mL (b) 0.0466 kg (c) 1.890 mg/L (d) 10.5 mm
2 -2
(e) 2.3 x 10 g (f) 7.11 x 10 L (g) 30 min (h) 30.0 0C

1.8.B How many significant figures does each one of the following
measurements have?
(a) 215.0 cm (b) 0.0034 L (c) 37.1 oC (d) 550 mg
-3 2
(e) 0.100 mL (f) 1.9 x 10 m (g) 2.6 x 10 g (h) 40.8 sec

1.9.A Give the answer of the following calculation to the correct number of
significant figures.
2.331 m3 (18.1 - 12.20) g
(a) (b)
(0.026 m)(2.55 m) (15.0 cm)(1.9 cm2 )

21
1.9.B Perform the following calculations and give your answer to the
correct number of significant figures.
(84.05 - 75.8) g (6.73 - 0.451) L
(a) (b)
23.05 mL (45.5 - 12.85) min

Experimental errors
1.10 The percentage of carbon in a compound was determined in triplicate
by two students. Their data were as follows:
student 1: 13.16%, 13.14%, 13.15%
student 2: 13.16%, 13.36%, 13.26%
The true value for the percentage of carbon was 13.27%. Which
student had the more accurate results, and which student the more
precise results?

22
CHAPTER 2
Atoms, Molecules and Ions
2.1 Atomic Structure
Protons, neutrons, electrons
Atoms are made up of subatomic particles: protons, neutrons, and
electrons. (More subatomic particles are known, but the three
mentioned here are relevant to Chemistry.) Protons and neutrons form
a compact (very small, dense) central body of the atom, the nucleus.
Electrons are distributed in space like a cloud around the nucleus.
Table 2.1 gives properties of these particles.
Table 2.1: Properties of subatomic particles
Relative Actual charge Mass Mass
Particle Symbol
charge (in Coulomb) (in g) (in amu*)
proton p 1+ +1.602 1019 1.673 1024 1.0073
neutron n 0 0 1.675 1024 1.0087
electron e 1 1.602 1019 9.109 1028 5.486 104
* amu = atomic mass unit = 1.6606 10-24 g

Masses of subatomic particles are extremely small. They are therefore


often expressed in atomic mass unit (amu) to give more convenient
numbers. As can be seen from the data in table 2.1, almost all of the
mass of an atom is contained in the nucleus the mass of an electron
is only about 1/2000 of the mass of a proton or neutron.

Example 2.1:
Which number of electrons has the same mass as a tiny grain of sand with
mass of 15 mg?
Solution:
Convert the mass of the grain of sand to grams since the mass of an
electron is given in the unit grams (table 2.1):
10-3 g
15 mg 15 mg 1.5 10- 2 g
mg
The number of electrons will be the mass of the grain of sand divided by
the mass of one electron:
23
1.5 10 - 2 g
no. of electrons 1.6 10 25 e _
- 28 _
9.109 10 g /e

The charge of an electron, proton or atom is normally given as its


relative charge whereby the electron charge is set at 1. Since atoms
are electrically neutral, the number of protons (in the nucleus) must
equal the number of electrons (around the nucleus). (Note that the
electron cloud around the nucleus has a definite structure see
chapter 6 about electronic structure of atoms.)
The diameter of an atom is of the order of 1010 m (e.g. for a hydrogen
atom, the smallest, the diameter is about 0.75 1010 m; for a cesium
atom, one of the largest, it is about 5.4 10 10 m). The diameter of
the nucleus is a factor 104 smaller and is of the order of 1014 m. An
atom is thus mostly empty space. (A useful, but non-SI unit is the
ngstrom () which equals 1010 m; the diameter of a cesium atom
can then be given as 5.4 .)

Atomic number, mass number, isotopes


The number of protons in atoms of an element is called the atomic
number, Z, of that element. Atomic number identifies an element: all
atoms of the same element have the same number of protons. For
example, for the element copper, Cu, Z = 29. All copper atoms have
29 protons in their nucleus.
Protons and neutrons are called nucleons. The mass number, A, of an
atom equals the number of nucleons (= protons + neutrons in the
nucleus). Neutrons, with no charge, do not affect the number of
protons (Z) or electrons (= Z) of an atom, but do affect mass and mass
number.
Atoms of the same element can have a different number of neutrons in
the nucleus. Isotopes of an element are atoms with the same atomic
number (Z) but different mass number (A) they have the same
number of protons, but a different number of neutrons. The masses of
these atoms are therefore not the same because of the difference in
neutron numbers.
For example, the element oxygen, O, consists of atoms with 8, 9, or 10
neutrons. Since they all have 8 protons (Z for O = 8), their mass numbers
are (8 + 8 =) 16, (8 + 9 =) 17 and (8 + 10 =) 18 respectively.

24
The following symbol is used to indicate an isotope:
A
ZX with superscript A = mass number, subscript Z = atomic
number, and X = chemical symbol of the element.
A possible charge of the isotope is shown as a superscript to the right
of the symbol.

Example 2.2:
What is the number of protons, neutrons and electrons in 129 53 I
(pronounced as 53 iodine 129), an isotope of the element iodine?
Solution:
Z, the atomic number = 53, therefore the number of protons = 53.
Since this is an atom, which is electrically neutral, the number of electrons
= number of protons = 53.
A, the mass number = 129, which equals number of protons + neutrons.
The number of neutrons is then A Z = 129 53= 76.

Example 2.3:
Give the isotopic symbol for a particle with 11 protons, 12 neutrons and
10 electrons.
Solution:
The number of protons = 11, therefore the atomic number Z is 11. We
can conclude that this particle is an isotope of the element sodium, Na.
The mass number = number of protons + neutrons = 11 + 12 = 23.
Finally, the number of electrons = 10. This is one less than the number of
protons. This particle is therefore an ion with charge 1+.

The isotopic symbol will be: 23 11 Na

The atomic number is often omitted as all atoms of a certain element


have the same number of protons, and number of protons is known
once the name or symbol of the element is known.
For example, the isotope of the element neon (Z = 10) with 10 neutrons can
be given as 20 Ne instead of 20
10 Ne . It is referred to as neon-20.

Isotopes of the same element have the same chemical properties


(determined by Z), and similar but not identical physical properties.
Note however that the physical properties for isotopes of hydrogen
25
differ substantially as their differences in mass are relatively large.
Isotopes of hydrogen with mass numbers 2 and 3 have been given
their own chemical symbol and name: 1H = hydrogen, D (2H) =
deuterium (D2O is called heavy water), T (3H) = tritium.
Note: Consult a General Chemistry text book to read more about the
discoveries which lead to the structure of the atom discussed in this
section, for example the experiments of Thompson, Rutherford,
Geiger and Marsden.

2.2 Atomic Weight


Most elements occur naturally as a mixture of isotopes. The atomic
weight of an element is the average atomic mass taking into account
the abundances of the different isotopes. It is the weighted average of
the isotopic masses and is expressed in atomic mass units.
% abundance of isotope i
Atomic weight mass of isotope i
i 100%

Example 2.4:
The element copper has two naturally occurring isotopes: copper-63 with
mass = 62.9298 amu and abundance = 69.09%; and copper-65 with mass
= 64.9278 amu and abundance = 30.91%. Determine the atomic weight
of copper.
Solution:
69.09% 30.91%
At. wt.Cu 62.9298 amu 64.9278 amu
100% 100%
= 43.478 amu + 20.069 amu = 63.55 amu

Note: Although the values are numerically very close (see example
above), the mass number of an isotope is not the same as its atomic
weight. Atomic weight is a mass with mass units (g, amu) while
mass number is a (whole) number without any units.

26
2.3 The Periodic Table
Features of the Periodic Table
All known elements are included in the Periodic Table. It lists the
elements according to increasing atomic number but arranged such
that elements with similar chemical and physical properties occur in
vertical columns. These vertical columns are called groups (or
families) and horizontal rows periods. The position of an element in
the Periodic Table gives a general idea of its properties.
18
1 1 2 H 13 14 15 16 17
Period numbers

2 Group numbers

Noble gases
Alkaline earth metals

3 3 4 5 6 7 8 9 10 11 12
Alkali metals

Halogens
4
5
6
7

Lanthanides
Actinides

Figure 2.1: The Periodic Table of the elements

The minimum information normally given for each element in the


Periodic Table includes atomic number, chemical symbol and atomic
weight (see the example in fig. 2.2). Some versions also include
electron configuration (see chapter 6) and several physical properties
of the element.
12 Atomic number
Mg Chemical symbol
24.305 Atomic weight

Figure 2.2: Information given for the element magnesium

The atomic weight for radioactive elements is given in parentheses.


27
Most of them are artificial, man-made. Elements with atomic number
1 92, with the exception of 43, occur naturally.

Groups in the Periodic Table


Groups are numbered from 1 18, although another numbering
system is also used in the USA. The long groups (groups 1, 2, and 13
18) are called the main groups with the representative elements.
Some groups have been given special names. Elements in group 1 are
called the alkali metals (e.g. sodium Na, potassium K), the most
reactive metals which normally form an ion with charge 1+, M+, in
their reactions with other elements. Group 2 is the family of the
alkaline earth metals (e.g. calcium Ca, barium Ba), also reactive
metals which normally form M2+ in their reactions. Group 17 are the
halogens (e.g. fluorine F, chlorine Cl), the most reactive non-
metals, often forming an ion with charge 1, X, when reacting. The
noble or inert gases (helium He, neon Ne, etc.) are the elements in
group 18; they are not reactive, although under special conditions the
heavier members of this group react with fluorine and oxygen.
The 10 short groups between groups 2 and 13 are called the transition
metals. They contain many important and well-known metals like
iron, chromium, copper, nickel, gold and silver. Two series of
elements, the inner transition metals, are found at the bottom of the
Periodic Table. The first row elements are called the lanthanides or
rare earth elements; the second series are the actinides or trans-
uranium elements. Their correct position is between groups 2 and 3,
and is often indicated by a special sign. However, to save space, these
two series of elements are normally placed at the bottom.
Finally, hydrogen is normally placed on its own, at the top of the
Table, and not in any of the groups. Although some of its properties
are similar to those of the elements in groups 1 and 17, hydrogen does
not fit either group completely.

Metals, non-metals, metalloids


Elements can be divided into metals, non-metals and metalloids (see
figure 2.3).

28
= metals 18
1 1 2 = metalloids H 13 14 15 16 17
2 = nonmetals B
3 3 4 5 6 7 8 9 10 11 12 Si
4 Ge As
5 Sb Te
6 Po
7

Figure 2.3: Metals, non-metals, metalloids in the Periodic Table

Metals are found in the lower left side of the Periodic Table the
majority of elements are metals. They conduct electricity, show
metallic lustre (shine), are malleable (can be hammered into sheets)
and ductile (can be drawn into wires). Non-metals are hydrogen plus
the elements in the upper right section of the Table. They are
electrical insulators, brittle if a solid; some are gases. Metalloids are a
diagonal line of elements, separating metals and non-metals: boron
B, silicon Si, germanium Ge, arsenic As, antimony Sb,
tellurium Te, and polonium Po. They show the physical properties
of metals but chemical properties of non-metals; some, like silicon,
are semiconductors.

2.4. Molecular Compounds


Many more compounds exist than elements new compounds are
being discovered or made daily. We distinguish between molecular
and ionic compounds. A second distinction is that between organic
and inorganic compounds.
Organic compounds are compounds containing carbon. Inorganic
compounds are all other compounds, but include a few carbon
containing compounds like CO, CO2, and the carbonates.
Molecular compounds consist of molecules, two or more atoms
chemically bound together. They are generally composed of non-
metallic elements. (A molecule is the smallest unit which still shows
29
the chemical properties of the substance.)
For example: H2O = water, H2O2 = hydrogen peroxide, NH3 = ammonia,
C12H22O11 = sucrose, SO3 = sulphur trioxide.
Note: Polyatomic molecules are composed of two or more atoms (P4,
H2S). More specifically, diatomic molecules are composed of two
atoms (O2, CO), triatomic molecules of three atoms (O3, SO2).

Chemical formulas of molecular compounds


There are different types of chemical formula for a molecular
compound:
the molecular formula, which shows the actual numbers and types
of atoms in a molecule of the compound;
For example: ethane = C2H6 , benzene = C6H6
the empirical formula, which shows the simplest, whole number
ratio of the atoms making up a molecule;
For example: ethane = CH3 (C : H = 2 : 6 = 1 : 3)
benzene = CH (C : H = 6 : 6 = 1 : 1)
the structural formula, which shows the arrangement of the atoms
in the molecule (chemical bonds, but not the geometry).
For example:

Note: Many text books also show ball-and-stick models (geometry of the
molecule) and space filling models (geometry of the molecule and
relative size of the atoms).

2.5 Ionic Compounds


Ionic compounds consist of positive and negative ions held together
by electrostatic forces in a geometric arrangement (a crystal lattice).
Figure 2.4 shows part of the crystal of sodium chloride, NaCl.

30
Figure 2.4: Crystal structure of NaCl

Ions
An atom which has lost one or more electrons is called a cation. Loss
of an electron leaves an excess positive charge. The charge of the ion
is shown as a superscript.
For example:
Na (atom) e + Na+ (cation)
(11 protons, 11 electrons) (11protons, 10 electrons)

An anion is an atom which has gained one or more electrons and is


thus negatively charged.
For example:
Cl (atom) + e Cl (anion)
(17 protons, 17 electrons) (17 protons, 18 electrons)

Ionic compounds are usually a combination of metals and non-metals,


in which the metal is the cation and the non-metal the anion.
For example: sodium oxide, Na2O (Na+ and O2); zinc floride, ZnF2
(Zn2+ and F).

Common monoatomic anions and cations


Monoatomic ions are single atoms that have gained or lost electron(s)
(e.g. K+, Ba2+, Cl, S2). The charge of a monoatomic ion is often
related to the group number of the element in the Periodic Table: ions
often have the same number of electrons as its nearest noble gas.

31
Example 2.5:
Consider the three elements before and after neon, Ne, the noble gas at the
end of period 2:
Element: N O F Ne Na Mg Al
Group number: 15 16 17 18 1 2 3
No. of e in atom: 7 8 9 10 11 12 13
Common ion: N3 O2 F Na+ Mg2+ Al3+
No. of e in ion: 10 10 10 10 10 10

Table 2.2 shows the general formula of monoatomic ions for each
main group in the Periodic Table. Common charges for ions of the
transition metals are difficult to predict and will have to be
memorized.
Table 2.2.: General formula for common monoatomic ions
Common cations
Group 1 M+ (Li+, Na+, etc.)
Group 2 M2+ (Mg2+, Ca2+, etc.)
Groups 3 - 12 Most transition metals have multiple charges (e.g. Fe2+,
Fe3+, Cu+, Cu2+); a few have only one common charge
(e.g. Ag+, Zn2+)
Group 13 M3+ (Al3+); the heavier elements also form M+ (e.g. In+,
In3+)
Group 14 The heavier elements form two cations, M2+ and M4+
(e.g. Pb2+, Pb4+)
Group 15 The heavier elements form two cations, M3+ and M5+
(e.g. Bi3+, Bi5+)
Common anions
Group 15 M3 for the lighter elements (N3, P3)
Group 16 M2 (O2, S2, etc.)
Group 17 M (F, Cl, etc.)

Example 2.6:
Predict the most likely monoatomic ion for strontium (Sr), bromine (Br),
selenium (Se), and rubidium (Rb).
Sr is in group 2, it is a metal, most likely ion is Sr2+.
Br is a member of group 17, a non-metal, most likely ion is Br .
32
Se belongs to group 16, it is a non-metal, most likely ion is Se2.
Rb is in group 1, it is a metal, most likely ion is Rb+.

Polyatomic ions
A polyatomic ion is a group of two or more atoms bonded together as
a unit that has lost or gained electron(s) and has a net positive or
negative charge. Note that the charge of a polyatomic ion belongs to
the group as a whole and not to one particular atom. An oxyanion is a
polyatomic anion with one or more oxygen atoms around another
central atom.
For example:
NH4+ = ammonium ion; H3O+ = hydronium ion (not an oxyanion);
PO43 = phosphate ion; CO32 = carbonate ion; NO3 = nitrate ion (all three
are oxyanions)

Chemical formula of ionic compounds


The chemical formula of an ionic compound shows the smallest
whole-number ratio of cations and anions. It is an empirical formula
and not a molecular formula since an ionic compound does not
consists of molecules. Ionic compounds must be electrically neutral,
the charges of cations and anions in the formula must balance.

Example 2.7:
Write the chemical formula for the ionic compounds calcium chloride,
lithium nitride, and aluminium sulphate.
Calcium chloride consists of Ca2+ (calcium) and Cl (chloride) ions. Two
Cl ions are required to balance the charge of one Ca2+ ion. The formula
is therefore CaCl2.
Lithium nitride consists of Li+ (lithium) and N3(nitride) ions. Three Li+
are needed to balance one N3. Chemical formula is Li3N.
Aluminium sulphate consists of Al3+ (aluminium) and SO42 (sulphate)
ions. Two Al3+ will balance three SO42. Chemical formula is Al2(SO4)3.
(The sulphate ion is included in brackets to indicate that the subscript 3
multiplies the whole SO42-.)

33
Table 2.3: Names and formulas of selected ions
Monoatomic cations Monoatomic anions
Hydrogen H+ Hydride (rare) H
Lithium Li+ Fluoride F
Sodium Na+ Chloride Cl
Potassium K+ Bromide Br
Rubidium Rb+ Iodide I
Cesium Cs+ Oxide O2
Magnesium Mg2+ Sulphide S2
Calcium Ca2+ Nitride N3
Strontium Sr2+
Barium Ba2+ Polyatomic anions
Aluminium Al3+ Sulphate SO42
Tin Sn2+, Sn4+ Sulphite SO32
Lead Pb2+, Pb4+ Thiosulphate S2O32
Phosphate PO43
Transition metal cations Phosphite PO33
Chromium Cr2+ , Cr3+ Hydroxide OH
Manganese Mn2+, Mn3+ Cyanide CN
Iron Fe2+, Fe3+ Cyanate OCN
Cobalt Co2+, Co3+ Thiocyanate SCN
Nickel Ni2+ Peroxide O22
Copper Cu+, Cu2+ Permanganate MnO4
Zinc Zn2+ Chromate CrO42
Silver Ag+ Dichromate Cr2O72
Cadmium Cd2+ Carbonate CO32
Platinum Pt2+, Pt4+ Hypochlorite ClO
Gold Au+, Au3+ Chlorite ClO2
Mercury Hg22+, Hg2+ Chlorate ClO3
Cerium Ce3+, Ce4+ Perchlorate ClO4
Silicate SiO44
Polyatomic cations Nitrate NO3
Ammonium NH4+ Nitrite NO2
Hydronium H3O+ Acetate CH3COO (C2H3O2)
_________________________________________________________________________________________________________

2.6 Naming of Inorganic Compounds


Although several substances are known by a common name, e.g.
CaCO3 is called limestone, CaO quicklime and Ca(OH)2 slaked lime,
we refer to substances in chemistry by their systematic name. These
are based on a set of rules agreed upon by IUPAC, the International
Union for Pure and Applied Chemistry.
34
Ionic compounds
Naming cations
The name of a monoatomic cation is the name of its element +
ion.
For example: K+ = potassium ion, Sr2+ = strontium ion

If an element can form cations with more than one charge, then
the charge is indicated with a Roman numeral in brackets after
the name of the element. This applies to most transition metals
and the heavier metals in groups 13-15. (The suffix -ous or -
ic to indicate different charges is officially no longer in use.)
For example: Fe2+ = iron(II) ion (not ferrous ion),
Fe3+ = iron(III) ion (not ferric ion)

Some special cases are NH4+, the ammonium ion, and H3O+, the
hydronium ion.

Naming anions
The name of a monoatomic anion is the stem of the element name
+ suffix ide + ion.
For example: F is an ion from fluorine; its name is fluoride ion,
O2 is an ion from oxygen; it is named oxide ion,
S2 is from sulphur; its name is sulphide ion.

Rules for the names of oxyanions (anions with one or more


oxygen atoms around another central atom) are rather more
involved:
# The suffix ate is added to the stem of the name of its
central atom.
For example: CO32 is the carbonate ion (from carbon),
SiO44 is the silicate ion (from silicon)

# If the same central atom can form two oxyanions then:


- use the suffix ate for the ion with more O atoms,
- use the suffix ite for the ion with fewer O atoms.
For example:
NO3 = nitrate ion, NO2 = nitrite ion (from nitrogen)

35
SO42 = sulphate ion, SO32 = sulphite ion (sulphur)
PO43 = phosphate ion, PO33 = phosphite ion (phosphorus)

# If the same central atom can form four oxyanions (e.g. the
halogens) then:
- use the prefix hypo with suffix ite for the ion with the
fewest O atoms,
- use the suffix ite and ate for the ion with the 2 nd and
3rd fewest O atoms respectively,
- use the prefix per with suffix ate for the ion with most
O atoms.
For example: ClO = hypochlorite ion,
ClO2 = chlorite ion,
ClO3 = chlorate ion,
ClO4 = perchlorate ion

Some anions include H+ in their formula and therefore also in the


name.
For example:
HCO3 = hydrogencarbonate ion (also called bicarbonate ion)
HPO42 = hydrogenphosphate ion
H2PO4 = dihydrogenphosphate ion

Examples of special cases are OH (hydroxide ion) and CN


(cyanide ion).

Naming ionic compounds


The name of an ionic compound is the name of the cation + the name
of the anion, written separately. The word ion is obviously omitted.

Example 2.8:
KClO3 consists of K+ (potassium ion) and ClO3 (chlorate ion); its name is
potassium chlorate.
Cu(OH)2 consists of Cu2+ (copper(II) ion, from the transition metal Cu
which can form more than one cation) and OH (hydroxide ion); its name
is copper(II) hydroxide.
36
NaH2PO4 consists of Na+ (sodium ion) and H2PO4 (dihydrogen
phosphate ion); its name is sodium dihydrogenphosphate.
NH4Fe(SO4)2 consists of NH4+ (ammonium ion), SO42 (sulphate ion) and
a cation of Fe. To get a neutral formula, the charge of the Fe ion must be
3+ (iron(III) ion, Fe can form more than one cation). The name of the
compound is ammonium iron(III) sulphate.

Acids
An acid is a hydrogen containing compound that produces H+ (also
referred to as H3O+) when dissolved in water. Although pure acids are
molecular compounds, they ionise in water to give the hydrogen ion
and an anion. The names of acids are based on the name of the anion
produced. (Acids will be discussed in more detail in chapter 4.)
If the name of the anion ends in ide, add hydro as a prefix
and ic acid as a suffix to the stem of the anion name.
For example:
HCl(aq) = hydrochloric acid (from Cl = chloride ion)
HCN(aq) = hydrocyanic acid (from CN = cyanide ion)

If the name of the anion ends in ite, add ous acid as a suffix
to the stem of the anion name.
For example:
HClO(aq) = hypochlorous acid (from ClO = hypochlorite ion)
H2SO3(aq) = sulphurous acid (from SO32 = sulphite ion)

If the name of the anion ends in ate, add ic acid as a suffix to


the stem of the anion name.
For example:
HClO4(aq) = perchloric acid (from ClO4 = perchlorate)
H2SO4(aq) = sulphuric acid (from SO42 = sulphate)

Note: Many compounds that are acids in water solution are molecular
when pure, or decompose outside water solution (see table below
for some examples). Therefore, always include (aq) (from
aqueous) to indicate an acid!

37
For example:
Acid Pure compound
HCl(aq) = hydrochloric acid HCl(g) = hydrogen chloride
H2S(aq) = hydrosulphuric acid H2S(g) = dihydrogen sulphide
H2CO3(aq) = carbonic acid CO2(g) = carbon dioxide
H2SO3(aq) = sulfurous acid SO2(g) = sulfur dioxide

Binary molecular compounds


Binary molecular compounds are amongst the simpler molecular
substances. It is a compound containing two non-metals.
When writing its formula:
give the symbol of the less electronegative atom (most left in the
Periodic Table) first,
give the symbol of the more electronegative atom (most right in
the Periodic Table, except noble gases) last,
O is normally the last symbol in the formula except when it is
combined with F (OF2 , O2F2 ).
The following applies when naming a binary molecular compound:
the name of the first atom in the formula is written first,
the stem of the name of the second atom + suffix ide is written
last,
use Greek prefixes to indicate the number of atoms of each
element:
1 = mono 4 = tetra 7= hepta 10 = deca
2 = di 5 = penta 8 = octa 11 = undeca
3 = tri 6 = hexa 9 = nona 12 = dodeca
Note: 1 is often not indicated with mono, especially with the first
atom.
The last a or o in the Greek prefix may be deleted if the
name of the second element starts with a vowel.

38
Example 2.9:
NF3 = nitrogen trifluoride SF6 = sulphur hexafluoride
N2O4 = dinitrogen tetroxide P4O10 = tetraphosphorus decoxide
N2O = dinitrogen monoxide (or nitrogen oxide)
CO = carbon monoxide

Several binary molecular compounds are better known by their


common name, for example, H2O = water and NH3 = ammonia (the
names dihydrogen oxide and nitrogen trihydride are not used).

Hydrates
A hydrate is an ionic salt that contains H2O molecules as part of the
crystal these belong to the compound and should be included in the
formula. We use a Greek prefix to indicate the number of H 2O
molecules.

Example 2.10:
CuSO4.5H2O = copper(II) sulphate pentahydrate
CoCl3.6H2O = cobalt(III) chloride hexahydrate
Potassium chromium(III) sulphate dodecahydrate = KCr(SO4)2.12H2O

2.7 Exercises
Atomic structure
2.1.A Complete the following table.
Isotopic No. of No. of No. of Atomic Mass
Charge
symbol protons electrons neutrons number number
47 61 47
9 9 19
56
Fe3+
30 31 69
79 117 1+

39
2.1.B Fill in the blanks in the table below.
Isotopic No. of No. of No. of Atomic Mass
Charge
symbol protons electrons neutrons number number
42 42 53
17 35 1-
207
Pb2+
25 30 2+
45 45 103

2.2 Which of the following pairs of species is a pair of isotopes?


(a) 16O2 , 16O3 (b) 40K , 40Ca (c) 56Fe2+ , 56Fe3+ (d) 12C , 14C

Atomic weights
2.3.A Naturally occurring boron consists of two isotopes: boron-10 (mass
10.01 amu) and boron-11 (mass 11.009 amu). The abundance of
boron-11 is 80.22%. What is the atomic weight of the element
boron?

2.3.B The atomic weight of the element lithium is 6.941 amu. It has two
naturally occurring isotopes: lithium-6 with mass 6.015 amu, and
lithium-7 with mass 7.016 amu. Determine the abundance of both
isotopes.

The Periodic Table


2.4. For each of the following elements, give its name and atomic
number, its group number (if the group has a special name, mention
it), and state if it is a metal, non-metal or metalloid.
(a) Kr (b) Br (c) Ge (d) Ca (e) Mo
(f) Rb (g) As (h) Al (i) Au (j) Xe

2.5.A If an ionic compound is formed between element X from group 2


and element Y from group 15, what will be the expected formula?

2.5.B An ionic compound is formed between element X from group 1 and


element Y from group 16. What is the likely formula?

40
Molecular and ionic compounds
2.6. Give the empirical formula for the following compounds:
(a) hydrogen peroxide, H2O2 (b) glucose, C6H12O6
(c) hydrazine, N2H4 (d) oxalic acid, H2C2O4

2.7. Explain if combinations of the following elements are expected to


give a molecular or an ionic compound.
(a) C and S (b) N and Cl (c) Ba and O
(d) Li and F (e) P and H (f) Ni and Br

2.8. Explain if the following formulas for ionic compounds are correct
or incorrect. If incorrect, write the correct formula.
(a) Al3(SO4)2 (b) NaH2PO4 (c) CaHCO3 (d) AgS
(e) Cr3Br (f) KMnO4 (g) Mg(NO3)2 (h) NaH
(i) NaCr2O7 (j) Zn2(CO3)2 (k) Sr(OH)2 (l) RbI2

Naming inorganic compounds


2.9.A Name the following compounds:
(a) Cr(H2PO4)3 (b) HClO2(aq) (c) Co(NO3)2.6H2O
(d) NCl3 (e) CuCl (f) H2SO4(aq)

2.9.B Give the name of the following compounds:


(a) NH4ClO3 (b) Fe(NO3)3.9H2O (c) P4O10
(d) NaHSO4 (e) BaCO3 (f) HBrO(aq)

2.10.A Give the chemical formula for the following compounds:


(a) Calcium hydrogencarbonate (b) Periodic acid
(c) Potassium phosphate (d) Ammonia
(e) Nickel(II) sulphate heptahydrate (f) Nitric acid

2.10.B What is the chemical formula for the following compounds:


(a) Dinitrogen trioxide (b) Bromous acid
(c) Barium hydrogenphosphate (d) Zinc nitrite
(e) Cobalt(III) chloride hexahydrate (f) Iron(II) sulphide

41
CHAPTER 3
Stoichiometry: Chemical Calcultions
3.1 Chemical Reaction Equations
Representing a chemical reaction
Reactions are at the heart of chemistry. A reaction equation shows a
chemical change: atoms from one or more substances are rearranged
to give another substance or substances. Reactants undergo a
chemical change, products result following the chemical change. A
reaction arrow indicates the direction of the reaction, and normally
points from left to right. Reactants are then placed left of the arrow,
products on the right hand side of the arrow:
Reactants Products

Example 3.1:
Combustion of charcoal (carbon) produces carbon dioxide gas. (A
reaction with oxygen gas is called combustion.). The reaction equation is:
C(s) + O2(g) CO2(g)

Balancing chemical reaction equations


A reaction equation must be balanced: the same number and kind of
each atom must appear in the reactants and in the products. (This is a
consequence of the Law of Conservation of Mass which states that
there is no observable change in overall mass during a chemical
reaction. In chemical reactions, atoms are neither created nor
destroyed but only rearranged.)
Use the following points when balancing an equation:
First, write the skeleton equation, i.e. an unbalanced equation
with the correct formulas of reactants and products.
Use coefficients to balance the reaction equation. A coefficient is
a whole number which is placed in front of a substance and
multiplies its whole formula.
# Balance the element which appears in the fewest formulas
first.
# Balance the element which appears in the most formulas last.
42
# Never change, add or delete a subscript in the formula of a
substance. A change in subscript means a different
substance: adding a subscript 2 to N or O in the formula NO
(nitrogen monoxide) will give N2O (dinitrogen oxide) or
NO2 (nitrogen dioxide) different oxides of nitrogen!
# A coefficient of 1 is normally not written, but implied.
# Include the physical state of each substance (if known):
(g) = gas, (l) = liquid, (s) = solid, (aq) = aqueous (dissolved
in water)
Many simple equations can be balanced by visual inspection (example
3.2). Others require a more systematic approach (example 3.3).

Example 3.2:
Calcium (a solid metal) reacts with oxygen gas to give calcium oxide (a
solid). Write the balanced reaction equation.
Solution:
The skeleton equation is: Ca + O2 CaO
Just looking at this equation shows that it will be balanced by adding a 2
in front of both Ca and CaO:
2 Ca + O2 2 CaO
(Remember, a coefficient multiplies the whole formula; the coefficient of
O2 , which is 1, is not written.)
Add the physical states of each substance to get the final balanced
reaction equation:
2Ca(s) + O2(g) 2CaO(s)

Example 3.3:
Butane gas (C4H10) burns in oxygen (from the air) to form carbon dioxide
gas and water vapour. Give the balanced reaction equation.
Solution:
First write the skeleton equation: C4H10 + O2 CO2 + H2O
H and C appear in only two formulas, O in three. Therefore, balance C
and H first. There is 4 C on the left (in C4H10), so we place a 4 in front of
CO2 to give us 4 C on the right as well. We find 10 H on the left (in
C4H10); on the right we find H in H2O, one H2O already contains 2 H so

43
we multiply H2O by 5 to give us also (5 2 =) 10 H on the right.
C4H10 + O2 4 CO2 + 5 H2O

Then balance O. There are 13 O atoms in the products (4 2 = 8 from


CO2 and 5 1= 5 from H2O). In the reactants, O appears in O2, an
oxygen molecule containing 2 O atoms. Multiplying O2 by 13/2 will also
give us 13 O in the reactants.
C4H10 + 13 O2 4 CO2 + 5 H2O
2

However, coefficients should be smallest whole numbers: we multiply all


coefficients by 2. We also add the physical state for each substance. The
balanced equation is then
2C4H10(g) + 13O2(g) 8CO2(g) + 10H2O(g)

3.2 Formula Weights; Percentage Composition


In the following sections, we discuss some concepts which are
required to perform calculations with balanced chemical equations.

Formula weights
The formula weight of a substance is the sum of the atomic weights of
all atoms in the formula. We distinguish three types:
Atomic weight (AW) is the weight of an atom. Atomic weights
for atoms of all elements can be found in the Periodic Table.
Molecular weight (MW) is the weight of a molecule of the
substance as given by the molecular formula. It applies to
molecular substances (compounds and elements which exist as
molecules).
Formula weight (FW) is the weight of the formula unit. Although
its definition includes atomic and molecular weight, formula
weight is especially used for ionic compounds. These
compounds do not exist of molecules, but of ions. Their formula
is an empirical formula giving the simplest whole number ratio of
the ions making up the compound.

44
Example 3.4:
Fructose is a sugar found in fruits. It is a molecular compound and has
the formula C6H12O6. What is its molecular weight?
Solution:
The molecular formula shows that one molecule of fructose consists of 6
C atoms, 12 H atoms, and 6 O atoms. Therefore
MW = 6 AW(C) + 12 AW(H) + 6 AW(O)
= 6 12.01 amu + 12 1.01 amu + 6 16.00 amu = 180.18 amu

Example 3.5:
The detergent (cleansing agent) washing soda is a hydrated ionic
compound with formula Na2CO3.10H2O. Give its formula weight.
Solution:
FW = 2 AW(Na) + 1 AW(C) + (3+10) AW(O) + 10 2 AW(H)
= (2 22.99 + 12.01 amu + 13 16.00 + 20 1.01) amu = 286.19 amu

Note that the units are atomic mass units (amu). The atomic weights
used were given to two decimal places. This number of decimal
places will normally give sufficient precision in the value for the
calculated formula weight. However, include more decimal places if a
higher precision (more significant figures) is required.

(Mass) percentage composition of a compound


The mass percentage composition of a compound gives the mass
percentages of each element making up the compound. Calculating
this composition is one of the steps in identifying a compound.
By definition,
mass of atoms of element X
%element X 100%
mass of compound
so that
no. of atoms of element X in formula AW(X)
%element X 100%
FW of compound

45
Often the word mass is omitted and percentage composition is used
to mean mass percentage composition.

Example 3.6:
Vitamin C is a molecular compound with the official name ascorbic acid
and formula C6H8O6. What is its % composition?
Solution:
MW(C6H8O6) = 6 AW(C) + 8 AW(H) + 6 AW(O)
= {(6 12.01) + (8 1.01) + (6 16.00)} amu = 176.14 amu
6 12.01 amu
%C 100% 40.91%
176.14 amu
8 1.01 amu
%H 100% 4.59%
176.14 amu
6 16.00 amu
%O 100% 54.50%
176.14 amu
Check, the sum of the mass percentages should be 100%:
%C + %H + %O = 40.91% + 4.59% + 54.50% = 100.00%

3.3 The Mole and Molar Mass


The mole, Avogadros number
The coefficients in a balanced equation indicate the ratio in which
numbers of particles (atoms, molecules, ions) react and are produced.
In practice, we do not count number of particles when we want to mix
reactants. Instead, we weigh out masses. The mole provides the link
between particle numbers and masses of substances.
The mole is defined as the amount of matter that contains Avogadros
number of particles; its unit is mol. Avogadros number, NA, equals
6.022 1023 mol-1. The particle referred to must be specified, e.g. 1
mol of H atoms, 0.1 mol of Na+ ions , 3 mol of H2O molecule, 0.05
mol of electrons.
We then have the following relationship between number of moles
and number of particles:

46
no. of particles of X
no. of moles of X
N A
no. of particles of X = NA no. of moles of X

Example 3.7:
What is the number of C atoms in 0.200 mol acetic acid, CH 3COOH, the
active ingredient in vinegar?
Solution:
No. of CH3COOH molecules = no. of moles CH3COOH NA
= 0.200 mol 6.022 1023 mol-1 = 1.204 1023
Each CH3COOH molecule contains two C atoms, therefore
No. of C atoms = 2 no. of CH3COOH molecules
= 2 1.204 1023 = 2.41 1023

Since the mole can be equated to a number, comparing number of


moles is the same as comparing number of particles example 3.8.

Example 3.8:
Natural gas is methane, CH4; propane, C3H8, is a component in LPG,
bottle gas. Which contains the larger number of hydrogen atoms, 2.0
mol of CH4 molecules or 5.0 1023 C3H8 molecules?
Solution:
Comparing numbers of H atoms is the same as comparing numbers of
moles of H atoms. We therefore express both in either number of moles
of H atoms, or in number of H atoms.
Comparing number of moles of H atoms:
Each CH4 molecule contains 4 H atoms. Therefore, no. of moles of H
atoms in 2.0 mol CH4
4 H atoms
= 2.0 mol CH 4 8.0 mol H atoms
CH 4

Each C3H8 molecule contains 8 H atoms, so that no. of moles of H atoms


in 5.0 1023 C3H8 molecules

47
5.0 10 23 C3 H8 molecules 8 H atoms
= 6.6 mol H atoms
23 -1
6.022 10 molecules mol C 3 8
H
Comparing both number of moles of H atoms shows that 2.0 mol of CH4
contains the larger number of H atoms.

Molar mass
The mass of one mole of a substance is called its molar mass. Units
are g/mol. Molar mass of a substance can be calculated from the
molar masses of the elements it contains. The mass (in g) of 1 mol of
atoms of an element is numerically equal to the mass of 1 atom (in
amu) of that element.

Example 3.9:
The Periodic Table shows us that the mass of 1 copper (Cu) atom (its
AW) = 63.55 amu. Therefore, the mass of 1 mol of Cu atoms (its molar
mass, MM) = 63.55 g/mol.
Since the mass of an electron is negligible compared to that of an atom,
we use the same AW, and thus MM, for an ion and its parent atom: MM
of Cu = MM of Cu+ = MM of Cu2+ = 63.55 g/mol.

Example 3.10:
Sodium hydrogencarbonate (NaHCO3) is used in baking powder (under
the common name, bicarbonate of soda).
Its molar mass is MM (Na) + MM (H) + MM (C) + 3 MM (O)
= (22.99 + 1.01 + 12.01 + 3 16.00) g/mol = 84.01 g/mol

Conversions: mass, number of moles and number of particles


The relationships between mass, number of moles and molar mass are:
mass of X (g)
By definition : molar mass of X (g/mol)
no. of moles of X

Then: mass X (g) = molar mass X (g/mol) no. of moles X (mol)


mass of X (g)
And: no. of moles of X (mol)
molar mass of X (g/mol)

48
Example 3.11:
What is the mass of 0.250 mol Na2SO4 (sodium sulphate)?
Solution:
Mass (Na2SO4) = MM (Na2SO4) no. of mol Na2SO4
Therefore, we must first find the molar mass (MM) of Na2SO4:
MM (Na2SO4) = 2 MM (Na) + MM (S) + 4 MM (O)
= (2 22.99 + 32.06 + 4 16.00) g/mol = 142.04 g/mol
Then: mass (Na2SO4) = 142.04 g/mol 0.250 mol = 35.5 g

Example 3.12:
What is the number of moles of CH3OH (methanol) in 4.96 g of this
compound?
Solution:
mass of CH 3OH
No. of moles of CH 3OH
molar mass of CH 3OH
Therefore, first calculate the MM of CH3OH. Addition of the molar
masses of the atoms in the molecule gives a value of 32.05 g/mol.
4.96 g
Then : no. of moles of CH 3OH 0.155 mol
32.05 g/mol

Example 3.13:
A sample of 0.0150 mol ZnBr2 (zinc bromide) has a mass of 3.38 g. What
is its molar mass?
Solution:
mass ZnBr2 3.38 g
molar mass ZnBr2 225 g/mol
no. of moles ZnBr 2 0.0150 mol

The mole is the link between mass and number of particles. Using
molar mass, we convert mass to number of moles. Number of moles
can be converted to actual number of particles (molecules, atoms)
using Avogadros number.

49
Example 3.14:
What is the number of carbon atoms in 5.00 g sucrose (C12H22O11,
sugar), about the mass of a tea spoon full?
Solution:
Mass divided by molar mass will give the no. of moles of sucrose:
Molar mass C12H22O11 = 342.34 g/mol (check).
5.00 g
No. of moles C12 H 22 O11 0.01461 mol
342.34 g/mol
Number of C12H22O11 molecules is its no. of moles Avogadros number:
= 0.01461 mol 6.022 1023 mol1 = 8.798 1021
One C12H22O11 molecule contains 12 C atoms. Then, the number of C
atoms will be the number of C12H22O11 molecules multiplied by 12:
= 12 8.798 1021 = 1.06 1023

Figure 3.1 shows the relationships we used in the above examples.


MM NA

Mass No. of moles No. of particles

MM NA
Figure 3.1: Converting mass, no. of moles and no. of particles

3.4 Empirical Formulas


After calculating the mass percentage composition, finding the
empirical formula is one of the next steps in determining the identity
of a compound. The empirical formula of a compound gives the
simplest whole number ratio of the atoms of each element in the
formula. In determining the empirical formula we use the fact that the
ratio of the actual number of atoms = the ratio of the number of moles
of atoms. We can find the empirical formula from the mass
percentage composition or from the actual masses of each element in a
sample of the compound.

50
Example 3.15:
A chlorinated organic compound contained 49.02% C, 2.74% H, and
48.23% Cl by mass. What is its empirical formula?
Solution:
It is easiest to consider exactly 100 g of the compound. Then the masses
of the elements (in g) equal their mass percentages:
In exactly 100 g of compound there is 49.02 g C, 2.74 g H and 48.23 g Cl.
Next, we must calculate the number of moles of each element in the 100 g
of compound:
no. of mol C = mass / MM = 49.02 g / 12.011 g mol1 = 4.0813 mol
no. of mol H = 2.74 g / 1.008 g mol1 = 2.718 mol
no. of mol Cl = 48.23 g / 35.453 g mol1 = 1.3604 mol
We then have the ratio of the no. of moles of each element:
no. of mol C : H : Cl = 4.0813 : 2.718 : 1.3604
We set the smallest number of moles at 1 by dividing each by the smallest
number (in this case, divide by 1.3604):
4.0813 2.718 1.360 4
ratio of no. of moles C : H : Cl = : :
1.3604 1.360 4 1.360 4
= 3.0001 : 1.998 : 1.0000
Rounded to whole numbers, the ratio of the number of moles C : H : Cl
=3:2:1
The empirical formula is then C3H2Cl.
Note that we used molar masses to three decimal places in order to
minimize rounding errors.

Because of experimental errors in determining the percentage


composition, the calculated numbers of moles for each element are
rarely whole numbers. After dividing by the smallest number of
moles to set this number at 1, we have to round to get whole numbers
(see example above). However, be aware of the possibility of
fractions (, , , , ) in the other numbers in the ratio before
rounding:
3.49 : 1.99 : 1.00 3 : 2 : 1 = (multiply by 2 = ) 7 : 4 : 2
2.98 : 1.35 : 1.00 3 : 1 : 1 = (multiply by 3 = ) 9 : 4 : 3
51
Example 3.16:
An oxide of phosphorus contains 43.64% P by mass. What is its
empirical formula?
Solution:
This compound contains only two elements, P and O. The mass% of O is
then (100.00 43.64)% = 56.36%.
After going through the steps shown in example 3.15, we find the
following ratio:
no. of moles P : O = 1.0000 : 2.5003 (check)
Obviously, the number 2.5003 is very close to 2, so that the ratio
becomes:
2
no. of moles P : O = 1.0000 : 2.5003 = 1 : 2 = 2 : 5
The empirical formula will be P2O5.

Molecular formula
The next step in identifying a molecular compound is finding its
molecular formula (the real formula). (Remember, for ionic
compounds we cannot go further than the empirical formula.) The
molecular formula is normally a multiple of the empirical formula:
molecular formula = (empirical formula)n (n = whole number)
The value for n can be found if the molar mass of the compound is
known:
molar mass compound
n
molar mass empirical formula

Example 3.17:
Consider the halogenated organic compound from example 3.15, with
empirical formula C3H2Cl. Its molar mass was found to be 147 g/mol.
Molecular formula (MF) = (empirical formula (EF))n
The molar mass (MM) of the empirical formula = 73.50 g/mol. Then:
MM compound 147 g/mol
n 2.00 2
MM empirical formula 73.50 g/mol
The molecular formula = (C3H2Cl)2 = C6H4Cl2

52
There are various methods to find the molecular formula of a
compound. One method, which can be used for volatile substances, is
discussed in chapter 5.

3.5 Calculations with Balanced Reaction Equations


The coefficients in a balanced reaction equation show the ratio in
which reactant particles react and product particles are produced.
Since the mole contains a specified number of particles (Avogadros
number), a balanced reaction equation also shows the ratio of the
number of moles of reactants and products.
For example:
Hydrogen gas and chlorine gas react to give hydrogen chloride gas:
Reaction equation: H2(g) + Cl2(g) 2HCl(g)
Coefficients: 1 molecule 1 molecule 2 molecules
NA: NA molecules NA molecules 2NA molecules
No. of moles: 1 mol 1 mol 2 mol
Ratio of the number of moles of H2 : Cl2 : HCl = 1 : 1 : 2. This ratio is
used to set up conversion factors (see chapter 1):
1 mol H 2 1 mol Cl 2 1 mol H 2 (or Cl 2 ) 2 mol HCl
and ; and
1 mol Cl 2 1 mol H 2 2 mol HCl 1 mol H 2 (or Cl 2 )
These conversion factors make it possible to relate actual number of
moles, or masses, of reactants and products. We can then predict the
mass of a certain reactant required, or the mass of a certain product
produced. (Obviously, these conversion factors apply to one
particular reaction.)

Example 3.18:
In the complete combustion of a sample of propane gas (C3H8), 10.0 g
CO2 gas is collected. What mass of propane burned, and what mass of O2
was required?
Solution:
The balanced reaction equation is
C3H8(g) + 5O2(g) 3CO2(g) + 4H2O(g)
giving a ratio of the no. of moles C3H8 : O2 : CO2 : H2O = 1 : 5 : 3 : 4.
53
Molar masses: C3H8 = 44.11 g/mol, O2 = 32.00 g/mol, CO2 = 44.01 g/mol.
No. of moles CO2 collected:
= mass CO2 / MM CO2 = 10.0 g / 44.01 g mol1 = 0.2272 mol
Therefore, the no. of moles C3H8 which reacted:
1 mol C3H8
= 0.2272 mol CO 2 0.07573 mol C3H8
3 mol CO 2
The mass of C3H8 which reacted = no. of mol C3H8 MM C3H8
= 0.07573 mol 44.11 g/mol = 3.34 g
Note: we could have combined the previous three steps into one:
10.0 g CO 2 1 mol C3 H8
mass C3H8 44.11 g/mol 3.34 g
44.01 g/mol 3 mol CO 2

no. of moles ratio molar mass


CO2 C3H8 : CO2 C3H8

no. of moles
C3H8
We will do the calculation of mass of O2 required in one step:
10.0 g CO 2 5 mol O 2
mass O 2 32.00 g/mol 12.1 g
44.01 g/mol 3 mol CO 2

3.6 Limiting Reactants; Reaction Yield


In practice, it is impossible to mix the reactants in exactly the amounts
required; balances cannot weigh to the molecule or atom precise.
Therefore, reactants are normally not available in stoichiometric
ratios. One reactant will be limiting, while the others are in excess.
Furthermore, the actual yield of the product(s) is normally less than
the theoretically possible maximum amount.

Limiting reactant
Consider the reaction of hydrogen with nitrogen to give ammonia:
Reaction equation: 3H2(g) + N2(g) 2NH3(g)
Imagine that 2.0 mol H2 and 2.0 mol N2 were mixed.
54
If all 2.0 mol H2 reacts, then no. of moles of N2 needed will be
1 mol N 2
2.0 mol H 2 0.67 mol N 2
3 mol H 2
More N2 (2.0 mol) is available, thus N2 is in excess and H2 is
limiting. Some N2 will be left after the reaction, but all H2 reacts
the number of moles of H2 determines the mass of product.
Therefore, 2.0 mol H2 and 0.67 mol N2 react.
No. of moles of N2 left = no. of moles (available - needed)
= (2.0 0.67) mol = 1.3 mol
No. of moles of NH3 produced
2 mol NH3
= 2.0 mol H 2 1.3 mol NH3
3 mol H 2
We reach the same conclusion if we had started our calculation with
the number of moles of N2:
If all 2.0 mol N2 reacts, then no. of moles of H2 needed will be
3 mol H 2
2.0 mol N 2 6.0 mol H 2
1 mol N 2
Less, only 2 mol, H2 is available, thus H2 is limiting and N2 is in
excess.

Example 3.19:
Which reactant is limiting when 50.0 g of water is added to 50.0 g
calcium carbide, CaC2(s)? What mass of acetylene, C2H2(g), is produced,
and what mass of excess reactant will be left?
Reaction equation: CaC2(s) + 2H2O(l) Ca(OH)2(aq) + C2H2(g)
Solution:
Required molar masses are:
CaC2 = 64.10 g/mol, H2O = 18.02 g/mol, C2H2 = 26.04 g/mol
No. of moles CaC2 available = 50.0 g / 64.10 g mol1 = 0.7800 mol
No. of moles H2O available = 50.0 g / 18.02 g mol1 = 2.775 mol
If all 0.7800 mol CaC2 reacts, then no. of moles H2O required will be

55
2 mol H 2O
0.7800 mol CaC 2 1.560 mol H 2O
1 mol CaC 2
More H2O is available (2.775 mol). Therefore, H2O is in excess, and CaC2
is limiting. All CaC2 reacts, it will determine the mass of C2H2 formed.
No. of moles of C2H2 produced:
1 mol C 2 H 2
0.7800 mol CaC 2 0.7800 mol C 2 H 2
1 mol CaC 2
Mass of C2H2 produced = 0.7800 mol C2H2 26.04 g/mol = 20.3 g

No. of moles of H2O left = no. of moles (available required)


= (2.775 1.560) mol = 1.215 mol
Mass of H2O left = 1.215 mol 18.02 g/mol = 21.9 g

Yield of the reaction


The amount of product collected is normally less than what is
predicted. Side reactions of the main reaction could occur and use up
some of the reactant(s). A product will generally have to be purified -
some of it could be lost during this process.
The theoretical yield is the maximum amount of product theoretically
possible, based on 100% reaction of the limiting reactant. The actual
yield is the real amount of product obtained. The percent yield is the
actual yield expressed as a percentage of the theoretical yield:
actual yield
% yield 100%
theoretica l yield
Obviously, the percent yield should be 100%.

Example 3.20:
Consider the reaction from example 3.19:
CaC2(s) + 2H2O(l) Ca(OH)2(aq) + C2H2(g)
We calculated that mixing 50.0 g of each reactant would produce 20.3 g
C2H2. This is the theoretical yield.
If 18.5 g of C2H2 was actually obtained, then the percent yield is
(18.5 g / 20.3 g) 100% = 91.1%

56
3.7 Exercises
Chemical equations
3.1. Balance the following reaction equations:
(a) _ MgCl2(aq) + _ KOH(aq) _ Mg(OH)2(s) + _ KCl(aq)
(b) _ Cu(NO3)2(s) _ CuO(s) + _ NO2(g) + _ O2(g)
(c) _ NH3(g) + _ NO(g) _ N2(g) + _ H2O(l)
(d) _ C2H6(g) + _ O2(g) _ CO2(g) + _ H2O(l)

Mole calculations
3.2 A. Perform the following mole calculations:
(a) Determine the molar mass of copper(II) sulphate pentahydrate,
CuSO4.5H2O.
(b) What is the number of moles of aspirin, C9H8O4, in a 500 mg
sample of this compound?
(c) Find the number of hydrogen atoms in 4.67 g of ethanol,
C2H6O?

3.2 B. What is the mass % composition of the following compounds?


(a) Na2CO3.10H2O (b) C12H20NS (c) C18H12Br2(OH)2

3.3. Which of the following amounts contains the largest number of


atoms: 27 g of phosphorus (P4) or 94 g of sulphur (S8)?

3.4. Explain if the following statements are true or false.


(a) One molecule of Br2 has a mass of 159.80 g.
(b) The molar mass of NaCl is the mass of 1 mol of NaCl
molecules.
(c) Since 1 molecule of CO2 contains 1 C atom and 2 O atoms,
one-third of its mass is due to carbon.
(d) One mole of NH3 contains Avogadros number of molecules.
(e) If the mass percentage of an element in a compound is 25%,
then 25% of the atoms in the compound belong to that
element.

Empirical formulas
3.5.A A compound of nitrogen and oxygen only contains 63.65% N by
mass. What is the empirical formula of this compound?

3.5.B A mass of 1.76 g tin reacted with excess fluorine to give 2.88 g of a
tin fluoride. Calculate the empirical formula of the compound.

57
3.6.A The empirical formula for an organic compound is C3H4. Its molar
mass is 120.2 g/mol. Determine its molecular formula.

3.6.B A certain compound has the empirical formula CH2O and a molar
mass of 60.0 g/mol. Find its molecular formula.

3.7.A Determine the empirical formula and molecular formula for a


compound containing 44.3% C, 43.5% Cl, 2.42% H, and 9.78% O.
Its molar mass is 325.9 g/mol.

3.7.B Calculate the empirical formula and the molecular formula for a
compound containing 5.26% H, 12.2% N, 55.6% O, and 26.9% P.
Its molar mass is 115 g/mol.

3.8. Explain if the following statements are true or false.


(a) The limiting reactant is the reactant with the smallest mass.
(b) Since C3H6O3 and C6H12O6 have the same empirical formula,
they represent the same compound.
(c) The compound Co4(CO)12 has the same percentage composition
as Co12(CO)4.

Calculations using reaction equations


3.9. Iron can be produced using reduction of one of its ores, Fe 2O3
iron(III) oxide, by carbon monoxide, CO, at high temperatures:
Fe2O3(s) + 3CO(s) 2Fe(s) + 3CO2(g)
Assuming 100% yield,
(a) what mass of iron(III) oxide is required to produce 10 kg of
iron,
(b) and how many grams of carbon dioxide are released?

3.10. A mass of 100 g of pentane, C5H12, is burned completely according


to the equation
C5H12(g) + 8O2(g) 5CO2(g) + 6H2O(g)
(a) Find the mass of carbon dioxide, CO2, produced.
(b) What mass of oxygen is required?
(Assume a yield of 100%.)

3.11. The following reaction equation shows the decomposition of nitrous


acid, HNO2(aq).
3HNO2(aq) HNO3(aq) + 2NO(g) + H2O(l)
If 55.0 g of nitrous acid reacts,
(a) what mass of nitric acid, HNO3(aq) is produced,

58
(b) what volume of water results (density of H2O = 1.00 g/mL)?
(Assume 100% yield.)

Limiting reactant, reaction yield


3.12. Sodium metal reacts with water to give sodium hydroxide and
hydrogen gas. (The heat released could set of an explosive reaction
of the hydrogen with oxygen from the air.) The reaction equation is
Na(s) + 2H2O(l) 2NaOH(aq) + H2(g)
When 10.00 g Na is added to 30.00 mL H2O (density = 1.00 g/mL),
(a) what is the limiting reactant,
(b) which mass of hydrogen, H2, is produced,
(c) how much of the excess reactant is left?

3.13. Lithium can react with nitrogen gas to form lithium nitride:
6Li(s) + N2(g) 2Li3N(s)
When 10.5 g of Li is made to react with 25.0 g of N2,
(a) what is the limiting reactant,
(b) which mass of lithium nitride, Li3N, is produced,
(c) how much of the excess reactant is left?

3.14. Upon heating, sulphur dioxide and hydrogen sulphide react to form
elemental sulphur and water:
SO2(g) + 2H2S(g) 3S(s) + 2H2O(l)
If 1.00 L of SO2 is mixed with 1.00 L of H2S,
(a) what is the limiting reactant (the density of SO2 is 2.81 g/L and
that of H2S is 1.49 g/L),
(b) what mass of sulphur is formed,
(c) how much of the excess reactant is left after the reaction?

3.15. Magnesium hydroxide, Mg(OH)2, is soluble in hydrochloric acid:


Mg(OH)2(s) + 2HCl(aq) MgCl2(aq) + 2H2O(l)
When 19.44 g of magnesium hydroxide were added to excess
hydrochloric acid, 28.75 g of magnesium chloride, MgCl2, were
produced. What is the percentage yield of the reaction?

3.16. Hydrogen bromide can be prepared in a reaction of sodium


bromide, NaBr, with excess phosphoric acid, H3PO4:
3NaBr(s) + H3PO4(l) 3HBr(g) + Na3PO4(s)
If 25.0 g of HBr is required and the yield is 87.5%, how many
grams of NaBr should be used?

59
CHAPTER 4
Reactions in Aqueous Solution
4.1 Introduction to Aqueous Solutions
Solutions
A solution consists of a solvent and one or more solute(s). The
solvent is the substance which does the dissolving and the solute the
substance which dissolves. Often, but not always, is the solvent
present in a larger amount.
We commonly mean by solution a liquid solution, a solution in which
the solvent is a liquid. However, air (a homogeneous mixture of
different gases) and some metal alloys (a homogeneous mixture of
different metals, like copper and zinc in brass) can also be considered
examples of a solution.

Aqueous solutions; electrolytes


An aqueous solution is a solution which has water as the solvent. One
way of classifying solutes is based on the electrical conductivity of its
aqueous solution:
A solute is called an electrolyte if its aqueous solution is able to
conduct electricity. The solution contains mobile ions which are
responsible for the electrical conductivity. These mobile ions
result when the solute breaks apart into individual ions (=
dissociates or ionizes) upon dissolving in water.
For example, when sodium chloride dissolves in water:
NaCl(s) Na+(aq) + Cl-(aq)
Dissociation (or ionization) of NaCl in water gives the hydrated
ions Na+(aq) and Cl(aq), single Na+ and Cl ions which are
surrounded by a layer of H2O molecules (see figure 12.3 in
chapter 12).
Almost all soluble ionic compounds and some molecular
compounds (acids, ammonia) are electrolytes.
A nonelectrolyte is a solute which does not produce mobile ions
upon dissolving in water. Its aqueous solution contains hydrated
solute molecules and does not conduct electricity.

60
For example, when sucrose dissolves in water:
C12H22O11(s) C12H22O11(aq)
Almost all molecular compounds (except acids, ammonia) are
nonelectrolytes.

Electrolytes can be classified as strong and weak electrolytes:


A strong electrolyte is completely (~ 100%) ionized in aqueous
solution.
For example, a solution of HCl, called hydrochloric acid:
HCl(aq) H+(aq) + Cl(aq)
Strong acids and strong bases and almost all soluble salts are
strong electrolytes.
A weak electrolyte is only slightly ionized in aqueous solution.
This is indicated by a double arrow.
For example, a solution of acetic acid, CH3COOH:
CH3COOH(aq) CH3COO(aq) + H+(aq)
In commonly used solutions of acetic acid, it is ionised for only
about 1% the exact degree of ionisation depends on its
concentration.
Weak acids and weak bases are weak electrolytes.
Note: The double arrow indicates a chemical equilibrium: both
reactants and products are present in the reaction mixture as
both forward and backward reactions take place
simultaneously and at the same rate see chapter 14 on
chemical equilibrium.

4.2 Precipitation Reactions


Reactions in aqueous solution are maybe the most important reactions
in chemistry. Precipitation reactions are reactions in water whereby at
least one of the products formed is insoluble.
For example, when mixing aqueous solutions of BaCl2 and Na2CO3:
BaCl2(aq) + Na2CO3(aq) 2NaCl(aq) + BaCO3(s)
The solid formed is called a precipitate.

61
Solubility in water
Based on its solubility in water we distinguish between:
soluble solutes: dissolves readily in water, 0.1 mol or more can
dissolve per litre solution,
insoluble solutes: do not dissolve significantly, less than 0.01 mol
can dissolve per litre,
slightly (or sparingly) soluble solutes: have an intermediate
solubility (between 0.01 0.1 mol can dissolve per litre).
Note that no substance is really completely insoluble in water;
however, the amount dissolved could be so small that, for all practical
purposes, we call the substance insoluble.

Solubility rules for common ionic compounds in water


The solubility of common ionic compounds in water can be predicted
using the following general solubility rules.
Soluble compounds are:
Ionic salts and hydroxides of group 1 cations.
Ammonium (NH4+) salts.
Chlorides (Cl), bromides (Br), and iodides (I), except when
combined with silver (Ag+), mercury(I) (Hg22+) and lead(II)
(Pb2+).
Nitrates (NO3), acetates (CH3COO), chlorates (ClO3) and
perchlorates (ClO4).
Sulphates (SO42), except when combined with strontium (Sr2+),
barium (Ba2+), lead(II) (Pb2+) and mercury(I) (Hg22+).

Insoluble compounds are:


Carbonates (CO32), chromates (CrO42), oxalates (C2O42) and
phosphates (PO43), except with group 1 cations and ammonium
ion.
Sulphides (S2), except with group 1 and 2 cations and
ammonium ion.
Hydroxides (OH) and oxides (O2, which becomes OH when
dissolved in water), except with group 1 and 2 cations (but not

62
magnesium). Note that the solubility of the group 2 hydroxides
increases going down the group:
Mg(OH)2 < Ca(OH)2 < Sr(OH)2 < Ba(OH)2
insoluble - - - - - - - - - - - - - - -> soluble

Compounds with a borderline solubility are calcium sulphate (CaSO4),


silver sulphate (Ag2SO4) and lead(II) chloride (PbCl2).

Example 4.1:
Predict if a precipitate forms when aqueous solutions of the following
strong electrolytes are mixed.
(a) (NH4)2S + CuSO4 (b) NaClO4 + AgNO3
Solution:
First, we must determine which ions are present in each solution, noting
that all compounds are strong electrolytes and thus completely dissociated
in water.
Then we must determine if possible combinations of cations with anions
result in the formation of an insoluble compound.
(a) (NH4)2S + CuSO4
The (NH4)2S solution contains NH4+(aq) and S2(aq) ions; the CuSO4
solution contains Cu2+(aq) and SO42(aq) ions.
Check if the combination of the cation from the first salt + anion of the
second salt produces an insoluble compound:
NH4+ + SO42 could give (NH4)2SO4. However, this compound is
soluble since all NH4+ salts are soluble.
Do likewise for the anion from the first salt + cation from the second salt:
S2 + Cu2+ could give CuS. This compound is insoluble since, with a
few exceptions which do not include Cu2+, all sulphides are insoluble.
Therefore, a precipitate will form; this precipitate is CuS.

(b) NaClO4 + AgNO3


The NaClO4 solution contains Na+(aq) and ClO4(aq) ions; the AgNO3
solution contains Ag+(aq) and NO3(aq) ions.
Cation (first salt) + anion (second salt) = Na+ + NO3 which could give
NaNO3. However, this compound is soluble on two counts: all group 1
compounds and all nitrates are soluble.

63
Cation (second salt) + anion (first salt) = Ag+ + ClO4 which could give
AgClO4. However, this compound is soluble as all perchlorates are
soluble.
Conclusion: no precipitate forms (no reaction takes place).

Net ionic equations


The net ionic equation for a reaction shows only those species which
actually take part in the reaction. It therefore gives a better picture of
what happens in a reaction than a molecular equation.
Use the following steps to write the net ionic equation:
First, determine the balanced molecular equation: give the
complete formulas of all substances which reacted and were
produced.
Then write the full ionic equation: this equation shows all the
ions for the soluble strong electrolytes which are completely
dissociated. Leave non-electrolytes and weak electrolytes, which
are not or hardly dissociated, as molecules. Gases and
precipitates are also not written in a dissociated form.
Finally, cancel spectator ions (= the ions remaining unchanged in
the reaction) to obtain the net ionic equation.

Example 4.2:
Mixing aqueous solutions of BaCl2 and Na2CO3 produces a precipitate.
Write the net ionic equation for this precipitation reaction.
Solution:
Added together were the following ions: Ba2+(aq), Cl(aq), Na+(aq) and
CO32(aq).
Using the solubility rules, only a combination of Ba 2+ and CO32 will give
an insoluble compound, BaCO3.
The balanced molecular equation is then:
BaCl2(aq) + Na2CO3(aq) 2NaCl(aq) + BaCO3(s)
All soluble compounds are strong electrolytes, the full ionic equation is:
Ba2+(aq) + 2Cl(aq) + 2Na+(aq) + CO32(aq)
2Na+(aq) + 2Cl(aq) + BaCO3(s)

64
Cancel species which appear unchanged on the left and right hand side of
the reaction equation: Cl(aq) and Na+(aq). These are the spectator ions.
The result is the net ionic equation:
Ba2+(aq) + CO32(aq) BaCO3(s)
Note that BaCO3, although an ionic compound, was not written in its
dissociated form since it is insoluble.

4.3 Acid Base Reactions


Acids
An acid is a molecular compound that dissolves in water to produce
H+(aq) (also described as H3O+(aq) = hydronium ion).
Note: We will see in chapter 15 that dissolving of an acid in water can
actually be described as a reaction with water, in which the acid is a
proton donor.

We can classify acids based on different criteria. One uses the number
of hydrogen ions which can be given off per acid molecule:
A monoprotic acid gives one hydrogen ion per acid molecule.
For example, nitric acid:
HNO3(aq) H+(aq) + NO3(aq) (strong acid)
A diprotic acid can give, stepwise, two hydrogen ions per acid
molecule.
For example, sulphuric acid:
H2SO4(aq) H+(aq) + HSO4(aq) (strong acid)
HSO4(aq) H+(aq) + SO42(aq) (weak acid)
Finally, a triprotic acid can give three hydrogen ions.
For example, phosphoric acid:
H3PO4(aq) H+(aq) + H2PO42(aq) (weak acid)
H2PO42(aq) H+(aq) + HPO42(aq) (weak acid)
HPO42(aq) H+(aq) + PO43(aq) (weak acid)

A second classification is that between mineral acids (= inorganic


acids) and organic acids. Organic acids are acids containing carbon,
but H2CO3, carbonic acid, is considered an inorganic acid. All other
65
acids are mineral acids.
Examples of organic acids are H2C2O4 = oxalic acid, HC2H3O2 (also
written as CH3COOH) = acetic acid, and H3C3H5O7 = citric acid.

Finally, we can divide acids into strong and weak acids:


A strong acid is effectively totally ionized in water.
Common strong acids are HCl(aq), HBr(aq), HI(aq), (not
HF(aq)!), HNO3(aq), HClO4(aq), H2SO4(aq) (first H only!), while
HClO3(aq) is borderline.
A weak acid is only slightly ionized in water.
All acids other than the strong acids mentioned above can be
considered weak acid.

Bases
A base is a molecular or ionic compound that produces OH(aq) when
dissolved in water, and reacts with H+(aq). Bases can be classified
using the same criteria as those for acids.

Examples of organic bases are the amines with general formula:


RNH2, R2NH, and R3N, in which R = alkyl group, e.g. CH3 = methyl.
For example, aminomethane
CH3NH2(aq) + H2O(l) CH3NH3+(aq) + OH(aq) (weak base)

A strong base is effectively totally dissociated (or ionized) in water.


For example, sodium hydroxide
NaOH(s) Na+(aq) + OH(aq) (strong base)
Common strong bases are the hydroxides and oxides of group 1 and
heavier group 2 metals. All other hydroxides are insoluble.
Group 1: general formula is MOH (M+ = Li+, Na+, K+, Rb+, Cs+).
Group 2: general formula is M(OH)2 (M2+ = Ca2+, Sr2+, Ba2+).
These bases show a decreasing solubility moving up the group:
Ba(OH)2 > Sr(OH)2 > Ca(OH)2
(Ca(OH)2 is a sparingly soluble strong base, while Mg(OH)2 is
insoluble, and thus not classified as a strong base.)

66
The oxide ion, O2-, reacts with water to give OH.
For example, adding lithium oxide to water:
Li2O(s) + H2O(l) 2Li+(aq) + 2OH(aq)
{ O2- + H2O(l) 2OH(aq) }

A weak base is only slightly ionized in water.


For example, a solution of ammonia in water:
NH3(aq) + H2O(l) NH4+(aq) + OH(aq) (weak base)
Although ammonia does not contain hydroxide ions in its formula, it
does produce OH when dissolved in water. An aqueous ammonia
solution is sometimes called ammonium hydroxide, because of the
presence of NH4+ and OH ions together in this solution. However, a
compound like NH4OH cannot be isolated from the solution

Neutralization reactions
A reaction between an acid and a base is called a neutralization
reaction. It has the general form:
acid + base salt + water
For example, the reaction between hydrochloric acid and aqueous
sodium hydroxide:
HCl(aq) + NaOH(aq) NaCl(aq) + H2O(l)
The salt is formed from the cation of the base and anion from the acid.

The net ionic equation for the reaction of any strong acid with any
strong base is:
H+(aq) + OH(aq) H2O(l)

If the reaction involves a weak acid or weak base, then the net ionic
equation includes the weak acid or weak base in its molecular form
(since that is the major species of the weak electrolyte present in
solution).
For example, acetic acid and potassium hydroxide solutions:
molecular: CH3COOH(aq) + KOH(aq) KCH3COO(aq) + H2O(l)
net ionic: CH3COOH(aq) + OH(aq) CH3COO(aq) + H2O(l)

67
or ammonia and hydrochloric acid solutions:
molecular: NH3(aq) + HCl(aq) NH4Cl(aq)
net ionic: NH3(aq) + H+(aq) NH4+(aq)
Note that the reaction equation for ammonia only shows a (soluble)
salt but not water.

The net ionic equation for the reaction of acids with insoluble metal
hydroxides or oxides includes the formula of the solid.
For example, addition of hydrochloric acid to magnesium oxide:
MgO(s) + 2HCl(aq) MgCl2(aq) + H2O(l)
MgO(s) + 2H+(aq) Mg2+(aq) + H2O(l)

Some reactions of acids with salts produce gases: H2S from S2-, CO2
from CO32-, SO2 from SO32-.
For example (molecular equations are given, write the net ionic
equations yourself):
NiS(s) + 2HCl(aq) NiCl2(aq) + H2S(g)
CaCO3(s) + 2HNO3(aq) Ca(NO3)2(aq) + { H2CO3(aq) }
H2CO3 is unstable, then: { H2CO3(aq) } CO2(g) + H2O(l)
2HI(aq) + K2SO3(aq) 2KI(aq) + { H2SO3(aq) }
H2SO3 is unstable, then: { H2SO3(aq) } SO2(g) + H2O(l)

4.4 Oxidation-Reduction Reactions


Oxidation and reduction
In oxidation-reduction reactions, in short called redox reactions,
electrons are transferred between reactants.
For example, the reaction between magnesium and hydrochloric acid:
molecular: Mg(s) + 2HCl(aq) MgCl2(aq) + H2(g)
net ionic: Mg(s) + 2H+(aq) Mg2+(aq) + H2(g)
In this reaction, Mg(s) is oxidised:
Mg(s) Mg2+(aq) + 2e-
and H+(aq) is reduced:
2H+(aq) + 2e- H2(g)
68
Oxidation is defined as the loss of electrons, or an increase in
oxidation number; reduction is the gain of electrons, or a decrease in
oxidation number. (Note that oxidation and reduction take place
simultaneously!)

Oxidation numbers
The oxidation number is the real or hypothetical charge which can be
assigned to an atom. It is used to recognize a redox reaction, and to
distinguish between the oxidation and reduction part in such a
reaction.
The rules for assigning oxidation numbers to atoms are the following:
For atoms in free (uncombined) elements: oxidation number is
zero.
For example, the atoms in He(g), H2(g) and S8(s) all have an
oxidation number = 0.
For atoms in monoatomic ions: oxidation number (ox. no.) is the
ionic charge.
For example, in Mg2+: ox. no. Mg = +2; in Br : ox. no. Br = 1
The sum of the oxidation numbers of all atoms in a species = total
charge of that species:
For example,
in H3PO4: (3 ox. no. H) + (ox. no. P) + (4 ox. no. O) = 0
in HSO4: (ox. no. H) + (ox. no. S) + (4 ox. no. O) = 1
Oxidation number for alkali metals (group 1) in compounds is
always +1: Li+, Na+, K+ etc.
Oxidation number for alkaline earth metals (group 2) in
compounds is always +2: Be2+, Mg2+, Ca2+ etc.
F in compounds always has oxidation number 1: F.
The oxidation number for H in compounds is usually +1, except
in binary compounds with active metals when it is 1 (hydride
ion).
For example, in HF: ox. no. F = 1 and H = +1; in NaH: ox. no.
Na = +1, thus H = 1.

69
Oxidation number for O in compounds is usually 2, except in:
peroxides (O22, ox. no. O = 1), superoxides (O2 , ox. no. O =
), or in binary compounds with F (OF2 : ox. no. F = 1, thus O
= +2).
The oxidation number for the halogens Cl, Br and I, is usually 1,
except in oxyanions or when combined with a more active
halogen (F is the most active, F > Cl > Br > I):
For example, in CaBr2: ox. no. Ca = +2 and Br = 1; in IF5: ox.
no. F = 1, therefore I = +5)
Aluminium and some transition metal cations have one charge:
Al3+, Zn2+, Cd2+, Sc3+.

Example 4.3:
Give the oxidation number of Cr in K2Cr2O7, P in H2PO4.
Solution:
Cr in K2Cr2O7: (2 ox. no. K) + (2 ox. no. Cr) + (7 ox. no. O) = 0
(2 +1) + (2 ox. no. Cr) + (7 2) = 0
ox. no. Cr = (+ 14 2) / 2 = +6
P in H2PO4: (2 ox. no. H) + (ox. no. P) + (4 ox. no. O) = 1
(2 +1) + (ox. no. P) + (4 2) = 1
ox. no. P = (1 + 8 2) / 2 = +5

Oxidising and reducing agents


The oxidising agent (oxidant) in a redox reaction accepts electron(s)
from the species which is oxidised, thereby causing the oxidation
step. It is therefore the species which is reduced and contains the
element that undergoes a decrease in oxidation number.
The reducing agent (reductant) in a redox reaction donates electron(s)
to the species which is reduced, thereby causing the reduction step.
It is therefore the species which is oxidized and contains the element
that undergoes an increase in oxidation number.

70
Example 4.4:
Identify the oxidising and reducing agents in the following reaction:
3Cu(s) + 2NO3(aq) + 8H+(aq) 3Cu2+(aq) + 2NO(g) + 4H2O(l)
Solution:
Since an element in both oxidising and reducing agent undergoes a change
in oxidation number, we first assign oxidation numbers to each atom in
the reaction equation (we will write the oxidation number below each
atom):
3Cu(s) + 2NO3(aq) + 8H+(aq) 3Cu2+(aq) + 2NO(g) + 4H2O(l)
(0) (+5)(2) (+1) (+2) (+2)(2) (+1)(2)

The oxidation number of Cu increases from 0 in Cu(s) to +2 in Cu 2+(aq).


Cu(s) is oxidised, and is therefore the reducing agent.
The oxidation number of N decreases from +5 in NO3(aq) to +2 in
NO(g). NO3(aq) is reduced, and therefore the oxidising agent.

Balancing redox reactions


Besides balancing the numbers of each kind of atom, the second
principle which must be applied to redox reactions is that all electrons
lost in the oxidation step must be gained in the reduction step
(conservation of electrons).
For simple redox reactions, visual inspection is often sufficient:
atoms must balance but also charges must balance. For example,
when copper is added to a solution containing the silver ion, copper
goes into solution giving it a blue colour, while solid silver (grey in
colour) is produced:
Cu(s) + Ag+(aq) Cu2+(aq) + Ag(s)
The number of atoms of each element is balanced, however, the
charges are not: the overall charge on the left is +1 and on the right
+2. This can be balanced by multiplying the reactant Ag+(aq) by 2.
We must then also multiply the produced Ag(s) by 2 to keep the
number of Ag balanced:
Cu(s) + 2 Ag+(aq) Cu2+(aq) + 2 Ag(s)

71
The half-reaction method
More complex redox reactions are not that easily balanced by visual
inspection. For these reactions, we use the half-reaction method. The
principle of this method is:
Oxidation and reduction are written as two separate half-reactions
(including electrons gained and lost).
Each half-reaction is balanced separately, then combined such
that number of electrons lost = number of electrons gained.

The detailed procedure for the half-reaction method takes the


following steps:
1. Write unbalanced half-reactions for the oxidation and reduction.
2. Balance all atoms except H and O.
3. Balance O by adding H2O to the appropriate side of the half-
reaction.
4. Balance H by adding H+ to the appropriate side of the half-
reaction.
5. In this step, we differentiate between an acidic and a basic
solution:
Acidic solution: H+ and H2O can appear in the half-reaction.
Normally at this stage, the half-reaction should already be
balanced.
Basic solution: OH and H2O can appear in the half-reaction.
Neutralize H+ from step 4 by adding the appropriate number
of OH so that all H+ are converted to H2O: (H+ + OH
H2O). Add the same number of OH to the other side of the
half-reaction. Simplify numbers of H2O.
6. Balance overall charges by adding electrons to the appropriate
side of the half-reaction.
(An easy check if this is done correctly is: electrons lost =
increase in oxidation number, electrons gained = decrease in
oxidation number.)
7. Equalize the number of electrons lost and gained in both half-
reactions by multiplying the half-reactions with appropriate
whole numbers.

72
8. Add the two half-reactions and simplify (cancel species,
coefficients should be smallest whole numbers).

Example 4.5:
Balance the following reaction in acidic solution: iron(II) ion reacts with
the permanganate ion to give the iron(III) and manganese(II) ions.
Solution:
First write the skeleton equation (the unbalanced overall equation), and
assign oxidation numbers to each atom in order to identify the oxidation
and reduction step:
Skeleton equation:
Fe2+(aq) + MnO4(aq) Fe3+(aq) + Mn2+(aq)
ox. no.: (+2) (+7)(-2) (+3) (+2)
The oxidation is Fe2+(aq) Fe3+(aq) since the oxidation number of Fe
increases fom +2 to +3. (Fe2+ is the reducing agent.)
The reduction is MnO4-(aq) Mn2+(aq) since the oxidation number of
Mn decreases from +7 to +2. (MnO4 is the oxidising agent.)

In the next steps we will balance each half-reaction. The numbers


preceding each step refer to the detailed procedure discussed on the
previous pages.
Oxidation half-reaction:
1. Fe2+ Fe3+
2 5 are not required.
6. Fe2+ Fe3+ + e

Reduction half-reaction:
1. MnO4 Mn2+
2. not required.
3. MnO4 Mn2+ + 4H2O
4. 8H+ + MnO4 Mn2+ + 4H2O

+7 +2

6. 5e + 8H+ + MnO4- Mn2+ + 4H2O (an electron has a 1 charge)

73
(Note that step 5 was not required since we are in acidic solution.)

We can now combine both balanced half-reactions.


7. Multiply the oxidation by 5 to have the same number of electrons as in
the reduction, and add: 5 oxidation + 1 reduction (5e- in each).
8. Overall reaction:
5Fe2+ + 5e + 8H+ + MnO4- 5Fe3+ + 5e + Mn2+ + 4H2O
Simplify and add the physical state to each species:
5Fe2+(aq) + 8H+(aq) + MnO4-(aq) 5Fe3+(aq) + Mn2+(aq) + 4H2O(l)

Example 4.6:
Solid bismuth(III) oxide reacts with the aqueous hypochlorite ion to give
the aqueous bismuthate ion (BiO3) and chloride ions (in basic solution).
Give the balanced reaction equation.
Skeleton equation:
Bi2O3(s) + ClO(aq) BiO3(aq) + Cl(aq)
ox.no.: (+3)(-2) (+1)(2) (+5)(2) (1)
Oxidation: Bi2O3(s) BiO3(aq) (Bi2O3 is the reducing agent)
(+3) (+5)
Reduction: ClO(aq) Cl(aq) (ClO is the oxidising agent)
(+1) (1)
Oxidation half-reaction:
1. Bi2O3 BiO3
2. Bi2O3 2BiO3 (balance Bi)
3. 3H2O + Bi2O3 2BiO3
4. 3H2O + Bi2O3 2BiO3 + 6H+
5. The solution is basic, so the H+ on the right is neutralized with 6OH;
6OH should also be added to the left to keep O and H balanced:
6OH + 3H2O + Bi2O3 2BiO3 + 6H+ + 6OH
6OH + 3H2O + Bi2O3 2BiO3 + 6H2O
6OH + Bi2O3 2BiO3 + 3H2O (number of H2O simplified)

6 2
6. 6OH + Bi2O3 2BiO3 + 3H2O + 4e
74
Reduction half-reaction:
1. ClO Cl
3. ClO Cl + H2O
4. 2H+ + ClO Cl + H2O
5. 2OH + 2H+ + ClO Cl + H2O + 2OH
2H2O + ClO Cl + H2O + 2OH
H2O + ClO Cl + 2OH (number of H2O simplified)

1 3
6. 2e + H2O + ClO Cl + 2OH
7. 1 oxidation + 2 reduction (then each half-reaction has 4e-)
8. Overall reaction:
6OH + Bi2O3 + 4e + 2H2O + 2ClO
2BiO3 + 3H2O + 4e + 2Cl + 4OH
(simplify and add physical states)
2OH(aq) + Bi2O3(s) + 2ClO(aq) 2BiO3(aq) + H2O(l) + 2Cl(aq)

4.5 Solution Concentration


Molarity
The concentration of a solute in solution is most often expressed in
terms of its molarity. Molarity of a solute in solution is defined as the
number of moles of solute per litre of solution (units = mol/L or M):
no. of moles of A (mol)
By definition: molarity of A (mol/L)
Volume of solution (L)
then: no. of moles solute A = molarity of A Volume solution (L)
no. of moles of A (mol)
and: Volume solution (L)
molarity of A (mol/L)
Example 4.7:
What is the molarity of Ca(NO3)2 in 500 mL of a solution which contains
25.0 g of this salt?
Solution:
The molar mass of Ca(NO3)2 = 164.10 g/mol

75
No. of moles of Ca(NO3)2 = 25.0 g / 164.10 g mol1 = 0.1523 mol
Molarity Ca(NO3)2 = 0.1523 mol / (500 103) L = 0.305 mol/L

Example 4.8:
What is the molarity of the aqueous ions in 0.150 M aq. Na3PO4?
Solution:
Na3PO4 is a soluble salt, a strong electrolyte and dissolves as:
Na3PO4(s) 3Na+(aq) + PO43-(aq)
The ratio no. of moles Na3PO4(s) : Na+(aq) : PO43(aq) = 1 : 3 : 1
Molarity Na+(aq) = 3 molarity Na3PO4 = 3 0.150 M = 0.450 M
Molarity PO43(aq) = molarity Na3PO4 = 0.150 M

Example 4.9:
What is the number of moles of Na3PO4 in 250 mL of a 0.150 M solution?
Solution:
No. of moles Na3PO4 = molarity Na3PO4 volume (L)
= 0.150 mol/L 250 103 L = 0.0375 mol

Example 4.10:
What volume of 0.200 M Ca(NO3)2 solution contains 0.0500 mol of
Ca(NO3)2?
Solution:
Volume needed (L) = no. of moles Ca(NO3)2 / molarity Ca(NO3)2
= 0.0500 mol / 0.200 mol L1 = 0.250 L (250 mL)

Dilutions
Commercially available solutions of ammonia and acids like nitric
acid are concentrated solutions. They often have to be diluted to
prepare solutions of the required molarity. A dilution is the
conversion of a concentrated solution to a more dilute solution by
adding solvent.
Since only solvent is added, the number of moles of solute in the
concentrated solution (before dilution) = the number of moles of

76
solute in the dilute solution (after dilution). If M = molarity and V =
volume, then:
M1 V1 (before dilution) = M2 V2 (after dilution)

Example 4.11:
What volume of 11.0 M HNO3 is required to prepare 1.00 L of 1.00 M
HNO3?
Solution:
M1 = 11.0 M, V1 = unknown, M2 = 1.00 M, V2 = 1.00 L, then:
11.0 M V1 = 1.00 M 1.00 L
V1 = (1.00 M 1.00 L) / 11.0 M = 0.0909 L (90.9 mL)

4.6 Calculations with Reactions in Aqueous Solution

Molarity calculations and mole calculations are combined to calculate


amounts (masses, volumes) of substances reacting and produced in
aqueous reactions. Titrations, which you will meet in the practical
component of this course, are well known examples.

Example 4.12:
If 45.75 mL of 0.0513 M HCl(aq) neutralizes a 20.00 mL sample of
Ba(OH)2(aq), what is the concentration of the aq. Ba(OH)2? And what is
the mass of Ba(OH)2 in the 20.00 mL?
Solution:
The reaction equation is:
2HCl(aq) + Ba(OH)2(aq) BaCl2(aq) + 2H2O(l)
No. of moles HCl reacting
= 0.0513 mol/L 45.75 103 L = 2.347 103 mol
Using the ratio of the no. of moles HCl and Ba(OH)2 in the reaction
equation,
1 mol Ba(OH) 2
no. moles Ba(OH) 2 in 20.00 mL 2.347 10-3 mol HCl
2 mol HCl
1.173 10-3 mol

Molarity Ba(OH)2 = no. of moles / V (L)

77
= 1.173 103 mol / 20.00 103 L = 0.0587 mol/L
The mass of Ba(OH)2 in the 20.00 mL sample
= no. of moles in 20.00 mL MM = 1.173 103 mol 171.35 g/mol
= 0.201 g (201 mg)

4.7 Exercises
Electrolytes
4.1. Explain if the following statements are true or false.
(a) Acetic acid is a weak acid because it is only sparingly soluble in
water.
(b) A solution of 1 mol of glucose (a sugar, C6H12O6) is a better
conductor of electricity than a solution of 1 mol of potassium
chloride, KCl, because it contains more ions.
(c) Aqueous NaCl is a good conductor of electricity since it
contains anions, Cl(aq), which will give electrons to the
solution.

Solubility, precipitation reactions


4.2. Write the net ionic equation and name the spectator ions for the
following reaction:
(a) Ni(NO3)2(aq) + Na2S(aq) NiS(s) + 2NaNO3(aq)
(b) CuSO4(aq) + BaCl2(aq) CuCl2(aq) + BaSO4(s)
(c) 2K3PO4(aq) + 3CaBr2(aq) Ca3(PO4)2(s) + 6KBr(aq)
(d) Fe(NO3)3(aq) + 3NaOH(aq) Fe(OH)3(s) + 3NaNO3(aq)

4.3. Which of the following pairs of reagents could you use to prepare
solutions that will give a precipitate of BaSO4 upon mixing?
(a) K2SO4 and BaCO3 (b) Ba(OH)2 and PbSO4
(c) Ba(NO3)2 and MgSO4 (d) BaCl2 and SrSO4

4.4. Write the molecular equation and the net ionic equation for the
reaction, if any, between aqueous solutions of:
(a) ammonium carbonate and calcium iodide
(b) silver nitrate and magnesium bromide
(c) copper(II) chloride and potassium acetate
(d) magnesium bromide and sodium hydroxide

78
Acids, bases, neutralisation reactions
4.5.A Indicate which of the following acids are weak acids.
Hydrofluoric acid (HF(aq)), hydrochloric acid (HCl(aq)),
hydrobromic acid (HBr(aq)), hydroiodic acid (HI(aq)), sulphuric
acid (H2SO4(aq)), sulphurous acid (H2SO3(aq)).

4.5.B Which of the following acids are strong acids?


Perchloric acid (HClO4(aq)), chloric acid (HClO3(aq)), chlorous
acid (HClO2(aq)), hypochlorous acid (HClO(aq)), nitric acid
(HNO3(aq)), nitrous acid (HNO2(aq)).

4.6. Write the molecular equation and the net ionic equation for the
reaction, if any, between aqueous solutions of:
(a) sodium hydroxide and hydrobromic acid
(b) barium hydroxide and sulphuric acid
(c) ammonia and nitric acid
(d) hydrofluoric acid and potassium hydroxide

Oxidation-reduction reactions
4.7. Give the oxidation number of
(a) Cr in K2Cr2O7 (b) Mo in MoO42 (c) P in CaHPO4
(d) S in Na2S2O3 (e) C in H2C2O4 (f) S in S4O62

4.8. Indicate for each of the following unbalanced half-reactions if it is


an oxidation or a reduction.
(a) Cr(OH)3(s) CrO42(aq)
(b) NO3(aq) NO(g)
(c) Cl2O7(l) ClO2(aq)
(d) H2S(g) S(s)

4.9. Balance the following half-reactions:


(a) NO(g) NO3(aq) (in acidic solution)
(b) ClO(aq) Cl(aq) (in basic solution)
(c) SO42(aq) SO2(g) (in acidic solution)
(d) Br-(aq) BrO3(aq) (in basic solution)

4.10. Balance the following redox reactions, using the half-reaction


method. Identify the oxidising and reducing agents.
(a) MnO4(aq) + HSO3(aq) Mn2+(aq) + SO42(aq) (acidic)
(b) Fe2+(aq) + Cr2O72(aq) Fe3+ + Cr3+(aq) (acidic)
(c) Pb(OH)42(aq) + ClO(aq) PbO2(s) + Cl(aq) (basic)
(d) MnO4(aq) + C2O42(aq) MnO2(s) + CO2(g) (basic)

79
Solution concentrations
4.11. (a) What mass of Li2SO4 is required to prepare 100 mL of an
aqueous solution which is 0.150 M Li2SO4.
(b) What volume of concentrated (= 12.0 M) hydrochloric acid is
needed to prepare 1.00 L of 0.200 M aq. HCl?

4.12. A solution is prepared by dissolving 2.44 g of barium chloride


dihydrate (BaCl2.2H2O) in water to a total volume of 250 mL.
What are the molarities of Ba2+(aq) and Cl-(aq) in the solution?

4.13. A volume of 100 mL of 0.225 M Zn(NO3)2 is diluted with distilled


water to a final volume of 300 mL. What are the molarities of
Zn2+(aq) and NO3-(aq) in the final solution?

4.14. What is the molarity of K+(aq) in a solution prepared by mixing 200


mL of 0.150 M KCl(aq) and 150 mL of 0.200 M K2SO4(aq)?

Solution stoichiometry
4.15. What volume of 0.150 M hydrochloric acid, HCl(aq), is required to
neutralise 2.58 g of calcium carbonate, CaCO3? Reaction equation:
CaCO3(s) + 2HCl(aq) CaCl2(aq) + CO2(g) + H2O(l)

4.16. Iodine, I2, can be determined in a titration with a standard solution


of aqueous thiosulphate, S2O32-(aq):
I2(aq) + 2S2O32-(aq) 2I-(aq) + S4O62-(aq)
What was the mass of iodine present in solution if 21.95 mL of
0.0506 M aq. S2O32- were required to just complete the reaction?

4.17. What is the molarity of an oxalic acid solution (H2C2O4(aq)) if


25.00 mL required 23.88 mL of a 0.205 M aq. NaOH solution? The
reaction equation is
H2C2O4(aq) + 2NaOH(aq) Na2C2O4(aq) + 2H2O(l)

80
CHAPTER 5
Gases
5.1 Pressure, Pressure Units

One property which is characteristic of a gas is its pressure. A gas


exerts a pressure because of the collisions of its particles with the
objects around them (for example, the walls of the container in which
it is kept). Pressure, P, is defined as force (F) exerted per unit area
(A):
F
P
A
Its SI unit is the pascal, Pa, which equals N / m2 (where Newton, N, =
kg m / s2). More common units for gas pressure used in chemistry are
the atmosphere (atm), Torricelli (Torr) and millimeter mercury
(mmHg):
1 atm = 1.01325 105 Pa (101.325 kPa)
1 atm = 760 Torr = 760 mmHg
Conversions between these units are often required.

Example 5.1:
Express 0.690 atm in (i) mmHg, and (ii) kPa.
760 mmHg
(i). 0.690 atm 0.690 atm 524 mmHg
1.00 atm
101.325 kPa
(ii). 0.690 atm 0.690 atm 69.9 kPa
1.00 atm

5.2 Simple Gas Laws


Four variables express the state of a gas: its pressure (P), volume (V),
temperature (T), and number of moles (n). The simple gas laws give
the relationship between two of these variables.

Boyles Law
We know from experience that the volume of a gas decreases if its
pressure increases. This is more precisely given in Boyles law which
81
states that the volume (V) of a fixed amount (n) of gas, kept at a
constant temperature (T), is inversely proportional to its pressure (P).
In terms of an equation:
1
V constant (at constant n and T)
P
or: P V = constant (at constant n and T)
so that: P1V1 = P2V2 (comparing the gas under two different
conditions at constant n and T)
Figure 5.1 shows a graphical representation of Boyles law.

100 100

75 75
Volume, V (L)
Volume, V (L)

50 50

25 25

0 0
0 1 2 3 0 1 2 3
Pressure, P (atm) 1 / Pressure, 1/P (atm-1)

(a) (b)
Figure 5.1: Boyles law: V vs. P (graph (a)) and V vs. 1/P (graph (b))
for a fixed amount of gas at constant temperature.

Charless Law
Volume of a sample of gas increases with temperature. This is stated
in (a reworded) Charless Law: the volume (V) of a fixed amount (n)
of gas at constant pressure (P) is directly proportional to the absolute
(Kelvin) temperature (T).
V = constant T (at constant n and P)
V
or: constant (at constant n and P)
T
V1 V2
so that: (the gas under two different conditions)
T1 T2

82
A graph showing Charless law is given in figure 5.2.

35 Figure 5.2: Charless law: V vs. T


30 for a fixed amount of gas at
25 constant P
Volume (L)

20
15 The line in figure 5.2 is
10 extrapolated to absolute zero, a
5 temperature of 0 K. At this
0 temperature, the volume of the gas
0 200 400 600
is predicted to be zero, which is
Temperature (K) not realistic.

Combined gas law


Boyles law and Charless law can be combined into one law, the
combined gas law:
PV
constant (at constant n)
T
P1V1 P2 V2
so that: (comparing two differerent conditions )
T1 T2

Example 5.2:
A sample of argon gas is stored in a cylinder with volume of 100 mL at
695 Torr and 15 0C. The gas is allowed to expand into a total volume of
1.00 L and stored at 27 0C. What is the final gas pressure?
Solution:
Call the original condition 1, and the new condition 2. Then
P1 = 695 Torr, V1 = 100 mL = 0.100 L, V2 = 1.00 L
T1 = 10 0C = (10 + 273.15) K = 283.15 K,
T2 = 27 0C = (27 + 273.15) K = 300.15 K
P1V1 P2 V2 PVT
so that P2 1 1 2
T1 T2 V2T1
695 Torr 0.100 L 300.15 K
P2 73.7 Torr
1.00 L 283.15 K

83
Gay-Lussac / Avogadros Law
The relationship between volume of a gas and its number of moles is
given by the (restated) law of Avogadro Gay-Lussac: the volume
(V) of a gas at constant temperature and pressure is directly
proportional to the number of moles (n) of the gas.
V = constant n (at constant P and T)

The molar volume, VM, of a gas is the volume occupied by one mole
of that gas:
V
VM (units are L/mol)
n
It is about the same for all gases at the same T and P. Since the
volume of a sample of gas depends on temperature and pressure, these
will have to be stated when a molar volume is given. For example:
VSTP = 22.4 L/mol (STP = 0 0C and 1 atm)
(STP stands for standard temperature and pressure.)
Note: Molar volumes for gases are much larger (~1000 times) than for
solids and liquids.

5.3 The Ideal Gas Law


The Ideal Gas Law combines the gas laws of Boyle, Charles and
Avogadro Gay-Lussac into one equation:
PV = nRT (where R is called the universal gas constant.)
The constant R has the same value for all gases, but can be expressed
in different units:
R = 8.20578 102 L atm K1 mol1
= 62.364 L Torr K1 mol1
= 8.31451 J K1 mol1
The value of R to be used depends on the units in which the pressure
is, or should be, expressed.

Example 5.3:
Find the value for the molar volume of a gas under room conditions
(which is 25 0C and 1 atm).

84
Solution:
The molar volume is the volume of 1 mol of gas, thus n = 1.00 mol
P = 1.00 atm, T = 25 0C = (25 + 273.15) K = 298.15 K
Since P is expressed in atm, we use R = 0.08206 L atm mol1 K1
Then:
nRT 1.00 mol 0.08206 L atm mol-1 K -1 298.15 K
VM 24.5 L
P 1.00 atm

Example 5.4:
A 10.0 L bottle contains oxygen gas at a pressure of 25.0 atm and
temperature of 24 0C. What is the mass of oxygen in the bottle?
Solution:
We can calculate the number of moles of O2 gas in the bottle using the
ideal gas law. Multiplying this number of moles with the molar mass of
O2 gives the mass of O2.
V = 10.0 L, P = 25.0 atm, T = (24 + 273.15) K = 297.15 K
PV 25.0 atm 10.0 L
n 10.25 mol
RT 0.08206 L atm mol-1 K -1 297.15 K
Mass of O2 = 10.25 mol x 32.00 g/mol = 328 g

Any gas that obeys the Ideal Gas Law under all conditions is called an
ideal gas; it shows ideal behaviour. Real gases approximate this
behaviour at low pressures and high to moderate temperatures (room
conditions).

5.4 Applications of the Ideal Gas Law


Density, molar mass
The density of a gas and its molar mass can be determined using the
ideal gas law.
Rearrange the expression for the ideal gas law:
n P
PV = nRT (eq.1)
V RT

85
Multiplying both sides by the molar mass (MM) will give:
n MM P MM
(eq.2)
V RT
but: n MM = mass (eq.3)
mass
and: d ( density) (eq.4)
V
P MM
substitute (eq.3) and (eq.4) in (eq.2): d (eq.5)
RT
d RT
rearrange (eq.5): MM (eq.6)
P
Equation 5 will be used to find the density of a gas when its molar
mass is known; equation 6 can be applied when the density is known
and the molar mass has to be determined.

Example 5.5:
At a pressure of 715 mmHg and temperature of 23 0C, a pure gas in a
flask of volume 504 mL had a mass of 1.83 g. Find the molar mass of the
gas.
Solution:
The density of the gas can be calculated since the mass and volume of the
gas are known:
m 1.83 g
d 3.631 g/L (express volume in litre)
V 0.504 L
Knowing density, we can then find the molar mass using:
d RT
MM
P
P = 715 mmHg = (715 mmHg / 760 mmHg) 1.00 atm = 0.9408 atm
T = 23 0C = (23 + 273.15) K = 296.15 K
3.631 g/L 0.08206 L atm mol-1 K -1 296.15 K
MM 93.8 g/mol
0.9408 atm

Gases in reaction equations


Combining the ideal gas law with mole and/or molarity calculations

86
(see chapters 3 and 4) allows us to include volumes of gases in our
calculations with reactions equations.

Example 5.6:
Potassium superoxide can be used to purify air. It reacts with carbon
dioxide and releases oxygen according to the balanced equation:
4KO2(s) + 2CO2(g) 2K2CO3(s) + 3O2(g)
What mass of KO2 is needed to react with 1.00 L of CO2(g) at 20 0C and
1.00 atm ?
Solution:
Convert the volume of carbon dioxide to number of moles using the ideal
gas law. Then use the balanced reaction equation to determine the
number of moles of potassium superoxide required. Multiply this number
of moles by the molar mass of the superoxide to find the mass required.
PV 1.00 atm 1.00 L
n (CO 2 ) 0.04157 mol
RT 0.08206 L atm mol-1 K -1 293.15 K

4 mol KO 2
No. of moles KO 2 0.04157 mol CO 2 0.08314 mol KO 2
2 mol CO 2
Mass KO2 = 0.08314 mol 71.10 g/mol = 5.91 g

5.5. Mixtures of Gases


Daltons Law; Partial pressure
Gases respond in a similar way to changes in P, V and T (i.e., they
generally follow the Ideal Gas Law, PV = nRT). Hence, a mixture of
gases which do not react chemically with each other behaves like a
single pure gas. As a result, the pressure of a gas mixture can be
found as the sum of the pressures of its components. This is stated in
Daltons law of partial pressures: the total pressure of a mixture of
gases is the sum of the partial pressures of its components.
Ptotal = P1 + P2 + . + Pn = Pi

Partial pressure of a gas is the pressure a gas would exert if it were


alone in the container. If 1, 2, ., n are the components in the
mixture, then the partial pressures are:

87
n RT n RT n RT
P1 1 , P2 2 , .. , Pn n
V V V
In a gas mixture, all gases have the same volume V and temperature
T, so that
n RT n 2 RT n RT RT
Ptotal 1 ....... n (n1 n 2 ..... n n )
V V V V
but: (n1 + n2 + . + nn) = ni = ntotal
n totalRT
so that: Ptotal
V

Example 5.7:
Ignoring the small amounts of carbon dioxide and noble gases, dry air
consists of 76% N2 by mass and 24% O2 by mass. What are the partial
pressure of each gas, and the total pressure of a sample of 1.00 g of dry air
with a volume of 1.00 L at 30 0C?
Solution:
First calculate the masses of nitrogen and oxygen in the sample of air:
mass N2 = (76% / 100%) 1.00 g = 0.760 g
mass O2 = (24% / 100%) 1.00 g = 0.240 g
We can then find the number of moles of both gases in the air sample:
no. of moles N2 = 0.760 g / 28.02 g mol1 = 0.0271 mol
no. of moles O2 = 0.240 g / 32.00 g mol1 = 0.00750 mol
Using the ideal gas law will give the partial pressure of each:
0.0271 mol 0.08206 L atm mol-1 K -1 303.15 K
P(N2 ) 0.67 4 atm
1.00 L
0.00750 mol 0.08206 L atm mol-1 K -1 303.15 K
P(O2 ) 0.187 atm
1.00 L
Ptotal = Pdry air = P(N2) + P(O2) = (0.674 + 0.187) atm = 0.86 atm

Mole fraction
The mole fraction, X, of component A in a mixture is the number of
moles of A expressed as a fraction of the total number of moles of all

88
components in the mixture:
XA = nA / ntotal
Since all gases in a mixture have the same temperature and volume,
the ratio of partial pressure of component A and the total pressure will
equal the mole fraction of A:
PA n RT / V n
A A A
Ptotal n totalRT / V n total
n
then: PA A Ptotal A Ptotal
n total
Note that the sum of the mole fractions of all components in the
mixture equals 1. For example, for a mixture of two components:
n1 n2 (n1 n 2 )
1 2 1
n total n total n total

Example 5.8:
The mole fraction of NO2 in a mixture of NO2 and CO2 kept at 760 Torr is
0.64. Calculate the mole fraction of CO2, and the partial pressure of NO2
and CO2.
Solution:
The sum of the mole fractions equals 1, therefore:
X(CO2) = 1.00 X(NO2) = 1.00 0.64 = 0.36
We can calculate the partial pressures knowing the mole fractions and
total pressure:
P(CO2) = X(CO2) Ptotal = 0.36 760 Torr = 2.7 x 102 Torr
P(NO2) = X(NO2) Ptotal = 0.64 760 Torr = 4.9 x 102 Torr

Collecting a gas over water


A gas produced in a reaction can be collected over water if it is
insoluble. However, the gas collected also contains some water
vapour. The total pressure of the gas mixture collected is then the sum
of the partial pressures of the gas and water vapour. Assuming that
the gas is collected at atmospheric pressure, then:
Patm = Ptotal = Pgas + Pwater

89
Example 5.9:
When heated, ammonium nitrite decomposes into nitrogen and water:
NH4NO2(s) N2(g) + 2H2O(l)
At 28 0C and an atmospheric pressure of 0.908 atm, 355 mL of N2 was
collected over water. What mass of NH4NO2 decomposed? (The water
vapour pressure at 28 0C = 28.35 Torr)
Solution:
Number of moles of N2 and NH4NO2 are related through the balanced
equation. We can find the number of moles of N2 using the ideal gas law,
once we know the partial pressure of N2.
The partial pressure of the nitrogen gas is the difference between the total
pressure of the gas collected (which equals atmospheric pressure) and the
water vapour pressure.
Pwater = 28.35 Torr = 28.35 Torr (1.00 atm/760 Torr) = 0.0373 atm
Ptotal = Patm = 0.908 atm
P(N2) = Ptotal Pwater = (0.908 0.0373) atm = 0.8707 atm
T = (28 + 273.15) K = 301.15 K , V = 355 mL = 0.355 L
0.8707 atm 0.355 L
n(N 2 ) 0.01251 mol
0.08206 L atm mol-1 K -1 301.15 K
In the reaction equation, 1 mol of N2 is produced by 1 mol of NH4NO2 ,
therefore the no. of moles NH4NO2 = no. of moles N2 = 0.01251 mol.
Mass NH4NO2 = 0.01251 mol 64.04 g/mol = 0.801 g

5.6 Molecular Motion: Effusion


Effusion is the escape of a gas through a very small hole (a pin hole)
into a vacuum, or into an area of lower pressure. This phenomenon
can be used to find the molar mass of a gas, as the rate of effusion
depends on molar mass. Grahams Law states that the rate of effusion
of a gas is inversely proportional to the square root of its molar mass:
1
rate where MM = molar mass (eq.7)
MM
Comparing two gases A and B which effuse under identical
conditions:

90
rateA MM B
(eq.8)
rateB MM A
The effusion rate has the units of volume (of gas) over time (for
example, mL/min), so that:
1
rate (eq.9)
time
Combining (eq.7) and (eq.9) will give:
1 1
(eq.10)
time MM
or comparing two gases A and B under identical conditions:
time A MM A
(eq.11)
time B MM B

Example 5.10:
A certain amount of neon gas takes 25 s to effuse through a porous
barrier. How long does it take for the same amount of methane gas
(CH4(g)) to effuse under the same conditions ?
Solution:
MM (Ne) = 20.18 g/mol, MM (CH4) = 16.05 g/mol
time (CH 4 ) MM (CH 4 )
Use (eq.11):
time (Ne) MM (Ne)
time (CH 4 ) 16.05 g/mol
Substitute values: 0.8918
25 s 20.18 g/mol
time (CH4) = 0.8918 25 s = 22 s

Remember that rate and time are inversely proportional. A heavier


molecule (larger MM) will move slower (lower rate) and therefore
require more time than a lighter molecule (smaller MM). Combining
(eq.8) and (eq.11):
rateB time A MM A


rateA time B MM B

91
5.7 Kinetic Theory of Gases
The gas laws discussed in this chapter can be explained by the kinetic
theory of gases. This theory is summarised as follows:
1. A gas is a collection of particles (atoms for noble gases,
molecules for other gases) that are in continuous, random motion.
2. The gas particles move in a straight line path until they collide
with each other and with objects around them. These collisions
are elastic - although the kinetic energy of individual particles
will change, there is no net loss of kinetic energy. No kinetic
energy is converted to heat.
3. The volume of gas particles themselves is ignored. They are
considered to be infinitely small and are treated as point masses.
4. Gas particles do not attract or repel each other.
5. The average translational kinetic energy of the particles is
directly proportional to the absolute temperature:
Ekin T
The average kinetic energy of all particles of a gas is the same at
any given temperature. This implies that a heavier particle will
have a lower average speed:
T
average speed
MM

As was mentioned before, under normal conditions (relatively low


pressure, moderate temperature) all gases obey the Ideal Gas Law
reasonably well. However, at high pressure and low temperatures,
deviations from the Ideal Gas Law become larger. Under these
conditions, the gas particles are closer together. Forces of attraction
and repulsion between them and their own volume cannot be ignored
any longer.

5.8 Exercises
Simple gas laws
5.1. (a) A gas sample of 16.5 mL at 14.5 oC is heated to 70.0 oC at
constant pressure. What is the new volume of the gas?

92
(b) Calculate the mass of Cl2 that has the same volume as 8.59 g of
SO2 gas if both are at the same temperature and pressure.
(c) A 1.0 L gas sample at 35 0C and 760 mmHg is compressed to a
final volume of 100 mL. What is the new pressure?

5.2. A sample of nitrogen gas was kept in a 3.00 L container at a


temperature of 35.0 oC and pressure of 746 Torr. What is the final
gas pressure if the temperature and volume are reduced to 15.0 oC
and 1.50 L respectively?

Ideal Gas Law


5.3. Sketch a graph showing the relationship between
(a) Pressure P and temperature T for a sample of an ideal gas kept
at a constant volume.
(b) Pressure P and number of moles n for an ideal gas kept at
constant volume and temperature.

5.4. (a) What is the molar volume of a gas at 1.00 atm and 100 oC.
(b) A sample of Ne gas with a mass of 0.500 g has a volume of 1.00
L at 25 oC. What is the temperature of a 0.500 g sample of Ar if
it has the same volume and pressure as the Ne sample?

Density, molar mass


5.5. (a) Find the density of ethylene, C2H4, at 760 mmHg and 0 oC.
(b) A sample of a gas with mass 2.24 g has a volume of 348 mL at
STP. What is the molar mass of the gas?
(c) A sample of a gas with mass of 3.53 g occupies a volume of 900
mL at 680 Torr and 32 0C. What is the density of the gas at
1.00 atm and 0 oC.

5.6. A gas containing 79.9 mass% carbon and 20.1 mass% hydrogen has
a density of 1.34 g/L at STP conditions.
(a) Find the molar mass of the gas.
(b) Determine the empirical formula of the gas.
(c) What is the molecular formula of the gas?

Gases in reaction equations


5.7. Addition of hydrochloric acid to calcium carbonate produces carbon
dioxide gas. Excess HCl(aq) was added to a 1.15 g sample of
impure calcium carbonate and the mixture heated to drive the CO2
out of the solution.
(a) Write the balanced equation for this reaction.
93
(b) If the mass% of CaCO3 in the sample was 95.0%, what volume
of CO2 was released at 25 0C and 1.00 atm.
(c) Why is carbon dioxide not collected over water?
(Assume that impurities in the sample do not interfere with the
reaction.)

5.8. Small quantities of hydrogen gas can be prepared in the lab by the
reaction of zinc metal with excess hydrochloric acid.
(a) Write the balanced equation for the reaction.
(b) If 250 mL of H2 gas is required and collected at 22 0C and 0.968
atm, what mass of zinc has to be added to the hydrochloric
acid?
(c) Explain if the hydrogen gas can be collected over water.

Gas mixtures
5.9. A gas mixture consisting of 0.200 g of He, 0.0228 mol of Ne and
0.0551 mol of Ar, has a volume of 600 mL at 30 oC.
(a) What is the partial pressure of each gas?
(b) What is the total pressure of the mixture?

5.10. The mole fraction of argon in air is about 0.009. What is the partial
pressure of Ar in air at an atmospheric pressure of 715 mmHg?

Grahams Law
5.11. The rate of effusion of an unknown gas is 0.79 times that of Ar
under the same conditions.
(a) Find the molar mass of the unknown gas.
(b) The unknown gas is an oxide of sulphur, SOx. Determine the
value for x, and thus the molecular formula of the gas.

5.12. An unknown gas takes 1.42 times longer to effuse than oxygen
under the same conditions.
(a) What is the molar mass of the gas?
(b) The unknown gas is a mono-halogenated ethane, i.e. ethane in
which one hydrogen atom has been replaced by an atom of a
halogen; it therefore has the formula C2H5X, where X is the
halogen. Find the identity of X, and then give the molecular
formula of the gas.

94
CHAPTER 6
Electronic Structure of Atoms
6.1 Electromagnetic Radiation
Wavelength, frequency
Electromagnetic radiation is a form of energy that sometimes acts like
a wave, and other times acts like a particle. Visible light is a well-
known example. All forms of electromagnetic radiation have two
inversely proportional properties: wavelength () and frequency ().
Wavelength is the distance from one wave peak to the next, which can
be measured in meters (1nm = 109 m) see figure 6.1. Frequency is
the number of waves that passes by a given point per second. It is
measured in Hertz (1 Hz = 1 s1).

Figure 6.1: Wavelength

Since wavelength and frequency are inversely related, their product


always equals a constant specifically, 3.00 108 m/s, which is
better known as the speed of light (c). Therefore, the relation between
wave length, frequency and the speed of light is = c.

Example 6.1:
The frequency of violet light is 7.31 1014 s1. What is its wavelength?
Solution:
Rearranging the expression = c gives = c / .
Therefore, = 3.00 108 m s1 / 7.31 1014 s1 = 4.10 107 m

95
The electromagnetic spectrum
Based on its wavelength (or frequency), we can differentiate between
different types of electromagnetic radiation. These are shown in the
electromagnetic spectrum (see figure 6.2). As can be seen, visible
light is only a tiny part of it. (Note that a logarithmic scale is used for
the wavelength.)

Figure 6.2: The electromagnetic spectrum

UV is the abbreviation for ultraviolet light, IR for infrared light.

6.2 Photons; Quantization of Energy


Photons
Instead of a wave, electromagnetic radiation is sometimes better
considered as a stream of very tiny particles or discrete energy packets
(amounts) called photons. (We call this the dualistic (= double) nature
of electromagnetic radiation). The energy of a photon of light is given
by E = h , where h is Plancks constant (6.63 1034 Js). Energy is
directly proportional to frequency, doubling its frequency will double
the energy of a photon.

Example 6.2:
What is the energy of a photon of light of wavelength 500 nm (which
appears blue-green to the human eye)?
Solution:
The frequency of this light is
= c / = (3.00 108 ms1) / (5.00 107 m) = 6.00 1014 s1
The energy carried by a single photon of this frequency is
E = h = 6.63 1034 Js 6.00 1014 s1 = 4.00 1019 J
96
Example 6.3:
Consider the light with wavelength of 500 nm which we discussed in
example 6.2. What is the number of photons emitted if a total energy of
20J is given off by the source of this light?
Solution:
Number of photons emitted will be the total energy divided by the energy
of 1 photon:
= 20 J / 4.00 1019 J photon1 = 5.0 1019 photons

Quantization of energy
Energy can be absorbed or emitted by atoms or subatomic particles
(electrons, protons or neutrons) only as fixed discrete amounts. A
quantum is the smallest amount of energy that can be emitted or
absorbed. It could be considered as the energy (equivalent) of a
photon of radiation.
Electrons in atoms can only be in certain discrete energy states their
energy is quantized (restricted). Therefore, atoms have a series of
precisely defined energy states only. (The energy of all moving
objects is quantized but it is only observable for atoms or subatomic
particles because the allowed energies are so close together.)

The photoelectric effect


A beam of light, directed onto a piece of clean metal, can cause
electrons to be ejected from its surface (see figure 6.3). Evidently, the
energy associated with the light overcomes the energy that binds the
electrons in the metal (this energy is also called work function).
Any energy the light supplies in excess of this binding energy appears
as kinetic energy of the emitted electrons.
What seems peculiar, however, is that the energy of the ejected
electrons does not depend on the intensity of the light (number of
photons). Instead, the energy of the photoelectrons (as they are
called) varies with the frequency, or wavelength, of the light. A
higher frequency of the radiation used (a shorter wavelength) gives a
greater kinetic energy to the ejected electrons.
In order to be knocked out of the metal, an electron must be given a
minimum amount of energy (the work function). Only radiation with
photons having at least this minimum energy can eject electrons from
97
the metal. Any energy of the photon in excess of this minimum will
show up as kinetic energy of the photoelectrons:
Ekin (photoelectron) = (Ephoton work function) with Ephoton = h

Figure 6.3: The photoelectric effect

6.3 Line Spectrum of Hydrogen Atom; the Bohr Model


Emission and absorption spectra
The emission spectrum of an atom is the characteristic range of
electromagnetic radiation (colours) it emits after absorbing energy
(for example, by being heated, being bombarded by electrons, or by
absorbing light/photons). The energy absorbed brings the atom in a
higher energy state; the absorbed energy is almost immediately lost, in
one or several steps, in the form of radiation.
The absorption spectrum of a substance is the opposite of an emission
spectrum. It shows a characteristic range of electromagnetic radiation
(colours) which is absorbed by the substance from light/radiation
directed onto it.

Line and continuous spectra


A line spectrum shows radiation at only certain distinct wavelengths.
Emission spectra of gaseous atoms are line spectra. A continuous
spectrum shows all possible wavelengths in a certain range; for
example, white light is a continuous spectrum of all possible colours.

The atomic spectrum of hydrogen


The emission spectrum of gaseous atomic hydrogen is a line spectrum
showing wavelengths in the UV, visible and IR regions. The various

98
lines (wavelengths) that are observable in the visible and near-UV
region of the spectrum are called the Balmer series, named after the
person who first observed them. Other examples of sets of related
lines in the hydrogen spectrum are the Lyman series (in the ultraviolet
region) and the Paschen, Brackett and Pfund series (in the infrared
region).
A simple mathematical formula shows the relationship of the lines in
each series. This formula is called the Rydberg equation:
1 1
1
RH
n 2 n 2
1 2
where RH (= Rydberg constant) = 1.097 107 m1
n1 and n2 are integers, with n2 > n1
This equation reproduces all the experimental line series for the
hydrogen atom in the IR, visible and UV regions see table 6.1.

Table 6.1: Values of n1 and n2 for the various line series in the
emission spectrum of atomic hydrogen

Name of series Value for n1 Values for n2 Region of spectrum


Lyman 1 2, 3, 4, 5, ... UV
Balmer 2 3, 4, 5, 6, ... Visible
Paschen 3 4, 5, 6, 7, ... Infra-red
Brackett 4 5, 6, 7, 8, ... Infra-red
Pfund 5 6, 7, 8, 9, Infra-red

Example 6.4:
Calculate the wavelength (in nm) for the second Balmer line.
Solution:
For the Balmer series (see table 6.1):
n1 = 2; for the 1st line, n2 = 3; for the 2nd line, n2 = 4
Use the Balmer equation to find wavelength in m, then convert to nm.

1 1 1 1 1
RH 1.097 107 m -1 0.20569 107 m -1
n 2 n 2 22 42
1 2

99
1
4.862 10- 7 m
0.20569 107 m -1
nm
4.862 10-7 m 486.2 nm (visible region)
-9
10 m

The Bohr model of the hydrogen atom


The Danish scientist Niels Bohr proposed the following model of the
hydrogen atom as an explanation for the observed line spectrum:
1. The electron moves in circular orbits of fixed radii and discrete
energies around the nucleus (see figure 6.4). The energy of the
electron does not change as long as it remains in the same orbit.
A principal quantum number (n; n = 1, 2, 3, 4, ) is used to
distinguish increasing orbit size and energy. When the electron is
in orbit with n = 1, it has the lowest possible energy; this is called
the ground state. States with n = 2, 3, 4, are called allowed
excited states of increasing energy.

Figure 6.4: Bohr model of the


hydrogen atom

2. The electron is normally found in orbit with n = 1 (the ground


state) but can be excited (absorbs energy) into a higher orbit. It
emits energy (radiation), when it relaxes into a lower orbit:
energy difference, E = Eexcited state Eground state = h
The energy of an electron in an atom is quantized: the electron
must be in one of the allowed orbits; each orbit has a defined
energy, En:
hR
En with hR 2.18 10 - 18 J
n2
The negative sign for En indicates that the energy of an electron

100
in an atom is lower than that of a free electron (when n = ).
After absorbing energy, the electron moves to an excited state,
say n2. When it then emits energy, it drops to a lower energy
state, say n1. (This lower energy state may be either another
excited state or the ground state.) The difference between the
energies of the initial (n1) and final (n2) states is:

hR hR 1 1
E E n E n
hR
2 1 2 2 2 2
n2 n1 n
1 n 2

Figure 6.5: Relationship between electronic transitions, shown by


arrows, and the different series of emission lines in the spectrum of
the hydrogen atom

6.4 Wave Nature of the Electron


We saw that electromagnetic radiation, normally regarded as a wave,
also shows the properties of a particle (photon). Similarly, de Broglie
proposed that an electron, which is normally regarded as a particle,
can also possess the properties of a wave (dualistic nature of matter).

101
Consider Einsteins relationship between mass and energy of matter:
E = mc2 (eq.1)
where E = energy of the particle, m = mass of the particle,
c = speed of light.
The energy of a photon is given by the equation:
c
E h h (eq.2)

Equating (eq.1) and (eq.2):
hc h
mc 2 which gives (eq.3)
mc
For all moving particles (objects), (eq.3) can be written as:
h
where v = velocity (eq.4)
mv
This equation which relates mass and wavelength of a moving object
is called the de Broglie equation. It shows that a particle in motion
can be treated as a wave, and a wave can exhibit the properties of a
particle.

Example 6.5:
Calculate the de Broglie wavelength of an electron travelling at 1% of the
speed of light (mass of an electron = 9.11 1031 kg).
Solution:
The speed of light is 3.00 108 m/s; 1% of that is 3.00 106 m/s.
Masses in the de Broglie equation are expressed in kg. Therefore we
change the units for Plancks constant (Js) to include kg as one of the
units: 1 J = 1 kg m2 s2, so that Js = kg m2 s2 s1 = kg m2 s1.
6.63 10 34 kg m 2 s 1
2.43 10 10 m (0.243 nm)
h

mv 9.11 10 31 6 -1
kg 3.00 10 m s

The wave nature of the electron was confirmed experimentally, when


it was shown that fast moving electrons are diffracted by a crystal in
the same way as X-rays.

102
6.5 The Quantum Mechanical Model of the Atom
Although the Bohr model was successful in explaining the spectrum
of the hydrogen atom, it failed to do so in a satisfactory manner for
atoms with more than one electron. This required a new model of the
atom which also took the wave properties of the electron into account.

The Schrdinger equation


The Schrdinger equation incorporates both the particle and wave
properties of the electron. For the hydrogen atom, the simplest atom,
it has the following form:
mv 2 e 2
E
2 r

Here we will not discuss the equation itself, but rather focus on the
information given by its solutions:
1. The allowed energies, E, of an atom can have only certain values.
2. The value of the wave function, , depends on the location of the
electron in space with respect to the nucleus of the atom. The
square of this function, 2, evaluated at any given point in space,
represents the probability of finding the electron at that particular
location.
3. The function defines a definite three-dimensional region which
is called an orbital.

Quantum numbers
An electron in an atom is at some location relative to the nucleus, and
is associated with some energy. According to quantum mechanics,
there is a finite probability of finding an electron around the nucleus.
The region of space, in which the probability of finding the electron is
large, is called an orbital or atomic orbital.

Each electron in an atom is described by a set of four numbers called


quantum numbers. Each electron in an atom has a unique set of
quantum numbers - these numbers can change if bonding occurs or if
an electron absorbs energy and moves to a higher energy state.

103
Principal quantum number (n)
Principal quantum number n determines the shell of the orbital in
which the electron can be found. It is the main determinant of the
energy of the electron (higher n corresponds to higher energy), as well
as nuclear distance (higher n means farther from the nucleus).
The principal quantum number can have any positive integral value: n
= 1, 2, 3, 4, ... . Maximum number of electrons in a shell is given by
2n2. The number of the period in which an element is found in the
Periodic Table indicates the number of electron-containing shells, for
example, helium (n = 1), neon (n = 2), argon (n = 3).

Azimuthal or angular momentum quantum number ()


The azimuthal (or angular momentum) quantum number, ,
determines the subshell the orbital is in. Values for depend on n and
range from 0 to (n 1). For example, if
n =1, there is only one possible value for : = 0;
n = 2, two values for are possible: = 0, 1;
n = 3, can have three possible values: = 0, 1, 2
n = 4, four values for are possible: = 0, 1, 2, 3.
Orbitals with the same value of n and form a subshell.

Subshells
Each type of subshell is represented by the letters s, p, d, f, .. :
a subshell with = 0 is called s-subshell;
for = 1 we have a p-subshell;
a d-subshell has = 2;
while the subshell with = 3 is called an f-subshell.

The number of possible values for a certain value of n indicates the


number of subshells:
the first shell, n = 1, has one subshell: s-subshell;
the second shell, n = 2, has two subshells: s- and p-subshells;
the third shell, n = 3, has three subshells: s-, p- and d-subshells;

104
the fourth shell, n = 4, has four subshells: a s-, p-, d- and f-
subshells.
(Note that the value for n is also the number of subshells in that shell.)

Magnetic quantum number (m)


The magnetic quantum number identifies a particular orbital in a
subshell; it determines the orbital in which the electron can be found.
Values for m are determined by : for a certain , m can have the
values , (1), , 0, , (1), . For example, subshell with =
2 has possible m values of: 2, 1, 0, 1, 2. (A certain value of gives
a maximum number of m values of (2 + 1).)
The number of m values which are possible for a value of
determines the number of orbitals in the subshell. Each type of orbital
has a unique shape.
When = 0, m can only be 0: m = 0. Therefore, the s-subshell
has only one orbital. It is shaped like a sphere see figure 6.6.
For = 1, m can have three values: m = 1, 0, 1. A p-subshell
has three orbitals (px, py and pz). They are shaped like a dumb
bell. (The pz orbital is oriented along the z-axis, the px orbital
along the x-axis and py along the y-axis see figure 6.6)
When = 2, m can have five values: m = 2, 1, 0, 1, 2. The d-
subshell has five orbitals.
When = 3, m can have seven values: m = 3, 2, 1, 0, 1, 2, 3.
Therefore, an f-subshell has seven orbitals.

Figure 6.6: Representation of an s- and a pz-orbital

105
The values of the three quantum numbers n, and m are summarized
in the following table (table 6.2).

Table 6.2: Summary of relationship between quantum numbers n, ,


m and shell, subshell and orbitals

Shell Subshell Orbitals Max. no. of e-


n=1 =0 1s m =0 1 1s 1 2e- = 2
n=2 =0 2s m =0 1 2s 1 2e- = 2
=1 2p m =+1,0,-1 3 2p 3 2e- = 6
n=3 =0 3s m =0 1 3s 1 2e- = 2
=1 3p m =+1,0,-1 3 3p 3 2e- = 6
=2 3d m =+2,+1,0,-1,-2 5 3d 5 2e- = 10
n=4 =0 4s m =0 1 4s 1 2e- = 2
=1 4p m =+1,0,-1 3 4p 3 2e- = 6
=2 4d m =+2,+1,0,-1,-2 5 4d 5 2e- = 10
=3 4f m =+3,+2,+1,0,-1,-2,-3 7 4f 7 2e- = 14

Example 6.6:
Determine the total number of orbitals in the shell with n = 4.
Solution:
For n = 4, there are four subshells: 4s, 4p, 4d, and 4f.
Each s-subshell has 1 orbital, each p-subshell has 3 orbitals, an d-
subshells has 5 orbitals while an f-subshell has 7 orbitals.
The total number of orbitals = 1 + 3 + 5 + 7 = 16

Spin quantum number (ms)


The spin quantum number is not related to values for n, , or m, but
can have only two possible values: + and . We can imagine that
it determines the spin of the electron: clockwise and anti-clockwise.

Pauli Exclusion Principle


Each electron in an atom can be assigned a set of values of four
106
quantum numbers, n, , m, and ms. Paulis Exclusion Principle states
that no two electrons in the same atom can have an identical set of
four quantum numbers. If two electrons have the same values of n,
and m (i.e. will be in the same orbital), then they will differ in ms
value.
It follows that each orbital can accommodate up to two electrons.
These two electrons have different values for ms, will show opposite
spin, and are said to be paired. For example, the two electrons in a
full 1s orbital (n = 1, = 0 and m = 0) will have the following sets of
quantum numbers: one with (n = 1, = 0, m = 0, ms = +); the second
with (n = 1, = 0, m = 0, ms = ).

6.6 Electron Configuration of Atoms


Energy of electrons in multi-electron atoms
The energy of an electron in a hydrogen atom is determined by its
principal quantum number. In a multi-electron atom, the energy of an
electron depends on both the principal quantum number and the
azimuthal quantum number. The stability of an electron in a multi-
electron atom reflects both the attraction between the electron and the
nucleus, and the repulsion between that electron and the rest of the
electrons present. Attraction as well as the repulsion depends on the
shape of the orbital in which the electron resides.

Aufbau Principle
The German word Aufbau means "building up". This term has
traditionally been used to describe the manner in which electrons are
assigned to orbitals as we carry out the imaginary task of constructing
atoms of elements with successively larger atomic numbers. In doing
so, we are effectively "building up" the Periodic Table of the
elements.
Electrons occupy preferably orbitals with the lowest possible
energy; lower-energy orbitals are filled with electrons before
orbitals with higher energy.
No more than two electrons can occupy any orbital.
The order of filling up subshells is as follows: 1s < 2s < 2p < 3s < 3p
< 4s 3d < 4p < 5s 4d < 5p < 6s 4f 5d < 6p < 7s 5f 6d < 7p.
107
(See fig. 6.7) It shows how the energies of the various subshells in the
same shell differ as a result of electron-electron repulsion. Energies of
d- and f-orbitals depend on the number of electrons they contain: once
they are filled their energy is lowered and then: 3d <4s ; 4d < 5s ; 4f <
5d <6s ; 5f < 6d < 7s.

Figure 6.7 : The order of filling orbitals in multi-electron atoms

Electron configuration
An electron configuration describes the arrangement of electrons in
the subshells of an atom. For a certain atom, the electron
configuration associated with the lowest energy levels corresponds to
the ground state and all others correspond to excited states.
Hydrogen is a single electron atom. This electron will be in the 1s
orbital in the ground state. Its electron configuration can be written
as: 1s1.

108
Orbital diagram
In an orbital diagram, the number of electrons in each orbital is also
shown. An electron is represented by an arrow; a box or horizontal
line represents an orbital. The arrow can point upwards or
downwards, depending on the spin of the electron:
(ms = +) and (ms = ).
Therefore, the electron configuration for the hydrogen atom in the
form of an orbital diagram is:

1s1

Electron configuration for multi-electron atoms


The electron configuration and orbital diagram for the elements which
follow hydrogen is given below.
Atom (At. no.) Electron configuration Orbital diagram
He (2) 1s2 (closed shell configuration)
1s2

Li (3) 1s2 2s1


1s2 2s1
A closed shell is a shell which contains the maximum number of
electrons. For example, for Li, electrons in the 1s (closed) shell are
called inner or core electrons and are not involved in the chemistry of
the element; electrons in the 2s subshell are called outer or valence
electrons and determine the chemistry of element.
Be (4) 1s2 2s2
1s2 2s2

B (5) 1s2 2s2 2p1


1s2 2s2 2px1
C (6) 1s2 2s2 2p2
Using the orbital diagram of boron, the sixth electron of carbon can be
placed in the following manners in the 2p subshell:
a. b. c. d. e. .
2px 2py 2pz 2px 2py 2pz 2px 2py 2pz 2px 2py 2pz 2px 2py 2pz
109
Orbital diagram (a) and (b) have the same, lowest energy (ground
state). The two electrons have the same spin (parallel spin) in two
different orbitals of the same subshell. Diagrams (c) and (d) are
allowed higher energy states (excited states). Orbital diagram (e) is
forbidden because it violates the Pauli Exclusion Principal.
This is summarized in Hunds rule which states that electrons must
occupy all degenerate (same energy) orbitals of a given subshell
singly before pairing starts. Therefore, the correct arrangement of
electrons in carbon is configuration (a).
The electron configuration of the elements next in period 2 can then
be written as follows:
N (7): 1s2 2s2 2p3
1s2 2s2 2pz1 2py1 2px1

O (8): 1s2 2s2 2p4


1s2 2s2 2pz1 2py1 2px2
F (9): 1s2 2s2 2p5
1s2 2s2 2pz1 2py2 2px2
Ne (10): 1s2 2s2 2p6 (closed shell)
1s2 2s2 2pz2 2py2 2px2

Condensed electron configurations


An electron configuration can be condensed by using the abbreviation
[ noble gas configuration] once the orbitals of a shell are filled - the
elements in the following period have the same core.
For example, the configuration of lithium is 1s2 2s1. Because it
contains a single 2s electron outside the helium-like 1s2 core, its
configuration can be written as: [He] 2s1.
The shell outside the noble gas core is called a valence shell and the
electrons in it are called valence electrons. The electron configuration
of atoms containing a noble gas core abbreviation is also called
condensed electron configuration.

The first two elements in period 4 will have the following electron
configuration:

110
K (19): 1s2 2s2 2p6 3s2 3p6 4s1 or [Ar] 4s1
Ca (20): 1s2 2s2 2p6 3s2 3p6 4s2 or [Ar] 4s2
The ten elements after calcium form the first transition series (Sc
through to Zn). These elements fill the 3d subshell. Their electron
configuration and orbital diagram can be written using the Aufbau
Principle and Hunds rule. For example,
Sc (21): 1s2 2s2 2p6 3s2 3p6 4s2 3d1 or [Ar] 4s2 3d1
Zn (30): 1s2 2s2 2p6 3s2 3p6 4s2 3d10 or [Ar] 4s2 3d10
However, there are two irregularities:
Cr (24): 1s2 2s2 2p6 3s2 3p6 3d5 4s1
and not: 1s2 2s2 2p6 3s2 3p6 3d4 4s2 as expected.
Cu (29): 1s2 2s2 2p6 3s2 3p6 3d10 4s1
and not: 1s2 2s2 2p6 3s2 3p6 3d9 4s2 as expected.
These irregularities are partly a consequence of the reduced electron-
electron repulsion when electrons remain unpaired (Hund's rule), as is
evident in chromium, which contains six unpaired electrons. In
addition, a subshell is particularly stable if it is half full (s 1, d5) or full
(s2, d10). This also applies to atoms in the same group as chromium
and copper.
The orbital diagrams for the first series of transition elements are
shown below:

3d orbital 4s
21
Sc
22
Ti
23
V
24
Cr
25
Mn
26
Fe
27
Co
28
Ni
29
Cu
30
Zn

111
6.7 Electron Configurations of Monoatomic Ions
When atoms form a cation, the highest energy electrons will be lost.
Often, the number of electrons lost is such that a closed shell
configuration results. Similarly, when atoms accept one or more
electrons to form an anion, the number of electrons gained is often
such so as to give a closed shell configuration. For example,
Na: 1s2 2s2 2p6 3s1 ([Ne] 3s1) will form Na+: 1s2 2s2 2p6 (=[Ne])
F: 1s2 2s2 2p5 ([He] 2s2 2p5) will form F: 1s2 2s2 2p6 (=[Ar])

If a transition metal forms a cation, then the electrons lost will first
come from the outermost s subshell, and then from the outermost d
subshell. For example, the electron configuration of Cu + is [Ar] 3d10
(4s1 electron lost); the electron configuration of Cu2+ is [Ar] 3d9 (4s1
electron and one 3d10 electron lost).

6.8 Exercises
Electromagnetic radiation, photons
6.1. Calculate the wavelength of radiation with the following frequency;
indicate to which part of the electromagnetic spectrum it belongs.
(a) 2.3 1015 Hz (b) 1.2 1010 Hz (c) 6.4 1019 Hz

6.2. (a) Find the energy of a photon of light with wavelength 540 nm.
(b) Calculate the energy (in J) of 2.0 mol of photons that has a
wavelength of 755 nm.
(c) A lamp emits a total energy of 30 J of light with a wavelength
of 465 nm. What is the number of photons emitted by the
lamp?

Line spectrum of hydrogen; the Bohr model


6.3. What is the wavelength (in nm) corresponding to an electronic
transition of n = 4 to n = 2 in the hydrogen atom?

6.4. Calculate the wavelength for the second line of the Paschen series
in the emission spectrum of the hydrogen atom. (For the Paschen
series, n1 = 3)

6.5. Determine the energy needed to ionize a hydrogen atom in the


ground state. (Ionization implies that the electron moves to n2 = .)

112
Wave nature of the electron
6.6. Calculate the wavelength corresponding to
(a) an electron having a velocity of 5.0 106 m/s (the mass of an
electron = 9.109 1028 g).
(b) a neutron with a velocity of 5.0 107 m/s (the mass of a
neutron = 1.674 1024 g).

The quantum mechanical model of the atom


6.7. What is the number of subshells, orbitals and electrons in the shell
with n = 3?

6.8. Explain if the following statements are true or false:


(a) The 4th shell contains 4 subshells.
(b) For the 5d orbitals, the possible values for m are 4, 3, 2, 1, 0,
1, 2, 3, 4.
(c) There are 4 atomic orbitals in the shell with n = 2.
(d) The quantum numbers (n = 3, = 3, m = 1, ms = ) could
describe an electron in a 3f orbital.

6.9. Which set of quantum numbers is allowed for an electron in an


atom?
(a) n = 4, = 4, m = 0, ms = +
(b) n = 2, = 1, m = 2, ms =
(c) n = 3, = 2, m = 0, ms = +1
(d) n = 0, = 0, m = 0, ms = 0

Electron configurations
6.10. Explain which electron configuration is possible:
(a) 1s2 2s3 2p5 (b) 1s2 2s2 2p5
(c) 1s2 2s2 2p6 3s2 3p7 3d9 (d) 1s2 2s2 2p6 3s2 3p5 3d10 4s2

6.11. Which of the following electron configurations would correspond to


a ground state and which to an excited state?
(a) 1s2 2s2 2p1 (b) [Ar] 4s2 4p1 (c) 1s2 2s2 2p1 3s1
(d) [Ar] 3d5 4s1 (e) 1s2 2p1 (f) 1s2 2s2 3s2

6.12. Write the ground state condensed electron configuration of the


following species:
(a) Mg (b) Si (c) As (d) S2 (e) Ti4+
(f) Kr (g) P (h) Cl (i) Ag (j) Co2+

113
6.13. Write the orbital diagram of the condensed ground state electron
configuration for the following species, and give the number of
unpaired electrons.
(a) S (b) Fe (c) F (d) Ge (e) Fe3+

6.14. In the ground state electron configuration of Br, how many


electrons have = 1 as one of their quantum numbers?

6.15. A cation with the condensed ground state electron configuration:


[Ar] 3d10 has a charge of 2+. An anion with a charge of 2 has the
condensed ground state electron configuration: [He] 2s 2 2p6. What
is the formula of the compound formed from these ions?

114
CHAPTER 7
Periodic properties of the Elements
7.1 Electronic Configuration and the Periodic Table
The relative orbital energies (discussed in chapter 6) and the Pauli
Exclusion Principle constitute the fundamental basis of the Periodic
Table of the elements. The Periodic Table consists of horizontal rows
known as periods and vertical columns known as groups.

Periods
There are seven periods (1 to 7) in the Periodic Table. The first period
has two elements, hydrogen and helium. The second and third periods
have eight elements which are, respectively, lithium (Li) to neon (Ne)
and sodium (Na) to argon (Ar). The fourth, fifth and sixth periods
consist of 18 (K to Kr), 18 (Rb to Xe) and 32 (Cs to Rn) elements
respectively. The seventh period starts from francium (Fr) and is
incomplete. Two rows of elements from La to Yb and Ac to No are
placed below the main body of the periodic table to save space but
they actually belong to the main body of the Periodic Table. (Note the
jump in atomic number going from Ba (56) to Lu (71) where the
elements La (57) Yb (70) should be placed, and from Ra (88) to Lr
(103) where the elements Ac (89) No (102) belong.)

Groups
In the past, two different systems of Roman numerals and letters were
used to denote the various groups. In the USA, the letter B denoted
the d-block groups while the letter A was used for the others. The rest
of the world used the letter A for the d-block elements and B for the
others. In 1985, a new international (IUPAC) system was adopted in
which the columns were simply labelled 1-18. This is what is shown
in the Periodic Table on the next page.

The elements on the left side of the periodic table are metals and those
on the right side are non-metals. Elements at the boundary between
metals and non-metals are known as metalloids.

115
Fig. 7.1: The Periodic Table of the elements

116
Periodic Table and electron configuration
Each period commences with two s-block elements (electron
configuration of the valence shell is ns1 and ns2) and continues
through the p block. At the end of each row corresponding to the
principal quantum number n > 1, there is an element that has an np6
configuration for the valence shell, a so-called noble gas element. At
n values of 3 and 4, d- and f-block element sequences are added.

Due to the periodicity in the properties, the elements that have similar
properties appear in the same group. Elements of groups 1 and 2 are
called alkali and alkaline earth elements respectively or s-block
elements. These elements have a filled inner shell and have one or
two electrons in their outer-most orbital. The general electron
configuration of s-block elements is ns1-2. Elements from groups 13 to
17 are called main group elements or p-block elements. In these
elements, p-orbitals are successively filled up. Their inner orbitals are
completely filled up. The general electron configuration of p-block
elements is ns2 np1-5 and they are placed towards the right of the
Periodic Table. It must be noted here that in noble gases, the last
electron goes into the p-orbital and therefore, they could be considered
as p-block elements. However, as they acquire a stable configuration
and have different properties, they are classed separately as inert or
noble gases. Both s- and p-block elements are collectively called
representative elements.

Elements from group 3 to 12 are the transition elements. They are


also called d-block elements. The general electron configuration of
transition elements is (n-1)d1-10 ns2 (where n = 4 to 6). In these
elements, a d-subshell is successively filled up. There are three series
of ten transition elements. Transition elements occupy the position
between s- and p-block elements.

Elements in the two rows placed below the main part of the Periodic
Table are called lanthanides and actinides, or f-block elements. They
are also referred to as inner transition elements. These elements are
filling a f-subshell; general electron configuration is (n-2)f1-14 ns2, but
note that there are several exceptions whereby an electron can also be
found in the (n-1)d subshell.
117
Elements in group 18 are called noble or inert gases. Except helium
(1s2), all noble gases have completely filled p-orbitals and have a
general electron configuration ns2 np6.

Summary
In summary, where n indicates the period number (and the shell
number), the general electron configuration of the groups is:
Groups 1-2: [preceding noble gas] ns1-2
Groups 3-12, Periods 4-5: [preceding noble gas] (n-1)d1-10 ns2
Periods 6-7: [noble gas] (n-2)f14 (n-1)d1-10 ns2
Groups 13-18, Periods 1-3: [noble gas] ns2 np1-6
(but note He with 1s2 only)
Periods 4-5: [noble gas] (n-1)d10 ns2 np1-6
Periods 6-7: [noble gas] (n-2)f14 (n-1)d10 ns2 np1-6
Lanthanides, actinides (Period 6-7): [noble gas] (n-2)f1-14 ns2
Note that exceptions to these general rules occur, especially with the
heavier elements.

18
1 1 2 H 13 14 15 16 17
2
3 3 4 5 6 7 8 9 10 11 12
s block

4
p block
5
d block
6
7

f block

Fig. 7.2: The different blocks in the Periodic Table

118
7.2 Size of Atoms and Ions
The concept of "size" is somewhat ambiguous when applied to the
scale of atoms and molecules. The reason for this is that an atom has
no definite boundary. There is a finite (but negligibly small)
probability of finding the electron of a hydrogen atom, for example, 1
cm, or even 1 km from the nucleus. It is not possible to specify a
definite value for the radius of an isolated atom; the best we can do is
to define a spherical shell within whose radius some arbitrary
percentage of the electron density can be found.

When an atom is combined with other atoms in a solid element or


compound, an effective radius can be determined by observing the
distances between adjacent rows of atoms in these solids. This is most
commonly carried out by X-ray scattering experiments.

Distance on the atomic scale have traditionally been expressed in


Angstrom units (1 = 10-8 cm = 10-10 m ), but picometer (pm) is
preferred: 1pm = 10-12 m = 10-10 cm = 10-2 , or 1 = 100 pm. The
radii of atoms and ions are typically in the range of 70-400 pm.

Because of the different ways in which atoms can aggregate together,


several kinds of atomic radii can be defined. Although the radius of
an atom or ion cannot be measured directly, in most cases it can be
inferred from measurements of the distance between adjacent nuclei in
a crystalline solid. Many atoms have several different radii. For
example, sodium forms a metallic solid and thus has a metallic radius;
it forms a gaseous molecule Na2 in the vapour phase in which it has
the covalent radius, and, of course, it forms ionic solids such as NaCl
where its radius is still different.

Atomic radii
The atomic radius of an element is defined as half the distance
between the centres of neighbouring atoms. In a metallic element, it is
the distance between the centres of neighbouring atoms in a solid
sample, and is called the metallic diameter. For example, this distance
in solid copper is 256 pm, the atomic radius of copper is half of it i.e.
128 pm.

119
If the element is a non-metal, atomic radius is the appropriate distance
between the centres of atoms joined by a chemical bond. Half of this
distance is also called covalent radius of the element. For example,
since the distance between the nuclei in chlorine molecule is 198 pm,
the atomic (covalent) radius of chlorine is 99 pm.

Fig 7.3: Van der Waals, metallic and covalent diameter

Periodic trends in atomic radii


The size of an atom depends mainly on the principal quantum number
of the highest occupied orbital; in other words, on the "number of
occupied electron shells". Since each row in the Periodic Table
corresponds to an increment in principal quantum number (n), the
atomic radius increases as we move down a column.

The other important factor is the nuclear charge; the higher the atomic
number, the more strongly will the electrons be drawn toward the
nucleus, and the smaller the atom. This effect is responsible for the
contraction we observe as we move across the Periodic Table from
left to right.

Generally, atomic radii increases down a group and decreases from


left to right across a period. The figure below shows a Periodic Table
in which the sizes of the atoms are depicted. The apparent
discontinuities in this diagram reflect the difficulty of comparing the
radii of atoms of metallic and non-metallic bonding types. Radii of
120
the noble gas elements are estimates from those of nearby elements.

Fig. 7.4: Covalent radii of elements

Ionic radii
The ionic radius of an element is the sharing distance between
neighbouring ions in an ionic solid. The inter-nuclear distance
between the nuclei of a cation and anion is the sum of the two ionic
radii. The radius of the oxide ion is generally used to find the radius
of the other ion. For example, if the inter nuclear distance in
magnesium oxide is 205 pm, and the radius of the oxide ion is 140
pm, then the radius of the magnesium ion is (205-140) = 65 pm.

Periodic trend in ionic radii


A cation is always smaller than the neutral atom, since the neutral
atom loses one or more electrons from its largest orbital, to form the
cation. If a second electron is lost, the ion gets even smaller; for
example, the ionic radius of Fe2+ is 76 pm, while that of Fe3+ is 65 pm.
If formation of the ion involves complete emptying of the outer shell,
then the decrease in radius is especially great. Just as is the case for
atomic radii, cation radii too increase down the group because
electrons are being added to orbitals with successively higher
principal quantum numbers.

Anions are always larger than the parent atoms; the addition of one or
more electrons to an existing shell increases electron-electron
121
repulsion which results in a general expansion of the atom. The
variation of ionic radii follows the same trend as that for atoms and
cations, with the smallest at the upper right of the Periodic Table,
close to fluorine.

Fig. 7.5: Ionic radii (in pm) of selected elements

Isoelectronic species
An isoelectronic series is a sequence of species all having the same
number of electrons (and thus the same extent of electron-electron
repulsion) but differing in nuclear charge. Of course, only one
member of such a sequence can be a neutral atom (neon in the series
shown below.) The effect of increasing nuclear charge on the radius
is clearly seen.

Fig. 7.6: Example of an isoelectronic series

122
Example 7.1:
From each of the following pairs of species, choose the one with the
largest ionic radius: (i) Ar and Ne, (ii) Te and Te2, (iii) Ti and Ti4+.
Solution:
(i) Electron configurations of Ne and Ar are :
Ne (10): 1s2 2s2 2p6 (outer n = 2)
Ar (18): 1s2 2s2 2p6 3s2 3p6 (outer n = 3)
The size of the atom increases down the group, hence argon atom has a
larger radius than the neon atom.
(ii) Te and Te2- : Anions are larger than their parent atoms. Therefore,
Te2- ion has the larger radius.
(iii) Ti and Ti4+ : A cation is smaller than its parent atom. Hence, Ti atom
has the larger radius.

7.3. Ionization Energy


Ionization energy always refers to the formation of positive ions
(cations). In order to remove an electron from an atom, work must be
done to overcome the electrostatic attraction between the electron and
the nucleus. This work is called the ionization energy of the atom.
The ionization energy of an atom or ion is defined as the minimum
amount of energy required to remove an electron from the ground
state of the isolated gaseous atom or ion and corresponds to the
exothermic process:
M(g) M+(g) + e (I1)
in which M(g) stands for any isolated (gaseous) atom and is called
first ionization energy (I1).

An atom has as many ionization energies as it has electrons. Electrons


are always removed from the highest-energy occupied orbital.
Removal of the second electron is much more difficult from the
smaller, positively charged M+(g), so the second ionization energy, I2,
is larger than I1. Thus the successive ionization energies increase in
the order, I1< I2 < I3. An examination of the successive ionization
energies of the first ten elements (fig. 7.7) provides experimental
confirmation that the binding of the two innermost electrons (1s
orbital) is significantly different from that of the n = 2 electrons. Note
123
the very large jumps in the energies required to remove electrons from
the 1s orbitals of atoms of the second-row elements Li-Ne.

Fig. 7.7: Successive ionizations (MJ mol-1) of the first ten elements

Ionization energies increase with the nuclear charge, Z*, as we move


across the Periodic Table. They decrease as we move down the Table
because in each period the electron is being removed from a shell one
step farther from the nucleus than in the atom immediately above it.
This results in the familiar zig-zag lines when first ionization energies
are plotted as a function of Z* (fig. 7.8). Generally, ionization
energies increase across a period and decrease down a group.

Fig. 7.8: First ionization energies of the first 20 elements


124
The following more detailed plot of the first ionization energies of the
atoms of the first ten elements reveals some interesting irregularities
that can be related to the slightly lower energies (greater stabilities) of
electrons in half-filled subshells relative to completely-filled ones (fig
7.9).

Fig. 7.9: First ionization energies of the first ten elements

Each of the group 13 elements (e.g. boron B) has a lower first


ionization energy than that of the element preceding it. The reversal
of the ionization energy trend in this group is often attributed to the
more easy removal of the single outer-shell p electron compared to
that of electrons contained in filled (and thus spin-paired) s- and d-
orbitals in the preceding elements.
Ionization energies (as well as many other properties) tend not to vary
greatly amongst the d-block elements. This reflects the fact that as the
more-compact d orbitals are being filled, they exert a screening effect
that partly offsets that increasing nuclear charge on the outermost s
orbitals of higher principal quantum number.

Example 7.2:
Arrange the following elements in order of increasing first
ionization energy: Na, Si, Mg, Al
125
Solution:
(i) Write the electron configuration of elements:
Na (11): 1s2 2s2 2p6 3s1; Si (14): 1s2 2s2 2p6 3s2 3p2
Mg (12): 1s2 2s2 2p6 3s2; Al (13): 1s2 2s2 2p6 3s2 3p1
The first ionization energy of elements increases across a period in the
Periodic Table. Therefore, the arrangement is Na, Mg, Al, Si.

7.4 Electron Affinity


Electron affinity (EA) refers to the formation of a negative ion
(anion). It occurs when an electron from some external source enters
the atom or ion in gaseous state and becomes incorporated into the
lowest energy orbital that possesses a vacancy. Because the added
electron is attracted to the positive nucleus, the formation of negative
ions is usually exothermic. The energy given off is the electron
affinity of the atom. For some atoms, the electron affinity appears to
be slightly negative, suggesting that electron-electron repulsion is the
dominant factor in these instances.
Electron affinity (EA) corresponds to the following processes:
M(g) + e(g) M (g) first electron affinity (EA1)
M(g) + e(g) M2 (g) second electron affinity (EA2)

In general, electron affinities tend to be much smaller than ionization


energies, suggesting that they are controlled by opposing factors
having similar magnitudes. These two factors are, as before, the
nuclear charge and electron-electron repulsion. The latter, only a
minor factor in positive ion formation, is now much more significant.
One reason for this is that the electrons contained in the inner shells of
the atom exert a collective negative charge that partially cancels the
charge of the nucleus, thus exerting a so-called shielding effect which
diminishes the tendency for negative ions to form.
Because of these opposing effects, the periodic trends in electron
affinities are not as clear as are those of ionization energies. This is
particularly evident in the first few rows of the periodic table, in
which small effects tend to be magnified anyway because an added
electron produces a large percentage increase in the number of

126
electrons in the atom.

1 2
H He
73 -21
3 4 5 6 7 8 9 10
Li Be B C N O F Ne
60 -19 27 122 -7 141 328 -29
11 12 13 14 15 16 17 18
Na Mg Al Si P S Cl Ar
53 -19 43 134 72 200 349 -35
19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36
K Ca Sc Ti V Cr Mn Fe Co Ni Cu Zn Ga Ge As Se Br Kr
48 -10 18 8 51 64 16 64 112 118 -47 29 116 78 195 325 -39
37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54
Rb Sr Y Zr Nb Mo Tc Ru Rh Pd Ag Cd In Sn Sb Te I Xe
47 30 41 86 72 53 101 110 54 126 -32 29 116 103 190 295 -41
55 56 71 72 73 74 75 76 77 78 79 80 81 82 83 84 85 86
Cs Ba Lu Hf Ta W Re Os Ir Pt Au Hg Tl Pb Bi Po At Rn
45 31 79 14 106 101 205 223 -61 20 35 91 183 270 -41
87
Fr
44

Fig. 7.10: First electron affinities (lower entry, in kJ mol-1) of


negative ions formed

In general, we can say that electron affinities become greater as we


move from left to right across a period (owing to increased nuclear
charge and smaller atom size). There are some interesting
irregularities, however:
In the Group 2 elements, the filled 2s orbital apparently shields the
nucleus so effectively that the electron affinities are slightly
endothermic.
The Group 15 elements have rather low values, due possibly to the
need to place the added electron in a half-filled p orbital; why the
electron affinity of nitrogen should be so small is not clear. The
vertical trend is for electron affinity to become lower in successive
periods owing to better shielding of the nucleus by more inner shells
and the greater size of the atom, but here also there are some apparent
anomalies.
In general, electron affinities increase across a period and decrease
down a group in the periodic table.

127
7.5 Exercises
7.1. Arrange the following particles in order of increasing radius:
(a) Al, Al3+, Na, Rb (b) Cs, Cr, Ca, Ge2+

7.2. Arrange the following particles in order of increasing first


ionization energy:
(a) Ga, Ba, B, Ca (b) F, K, Br, Ge

7.3. Which one of following species most likely has the lowest second
ionization energy: Na, Mg, Cl, Si?

7.4. Arrange the following elements in the expected order of increasing


metallic character: Ca, Se, Ga, Rb.

7.5. (a) Write the reaction equations corresponding to the first and
second electron affinity of oxygen, O.
(b) Give the reaction equations corresponding to the first and
second ionization energy of copper, Cu.

128
CHAPTER 8
Basic Concepts of Chemical Bonding
A chemical bond is a strong force of attraction that holds atoms
together in a molecule and ions in an ionic compound. The formation
of a chemical bond usually involves the valence shell electrons of
atoms. Chemical bonds can be classified into two types: ionic and
covalent bonds.

8.1 Ionic Bonds


An ionic bond is formed between two atoms when one or more
electrons are transferred from the valence shell of one atom to the
valence shell of the other atom. The atom that loses electron(s)
becomes a cation and the atom that receives electron(s) becomes an
anion. The ionic bond results from the electrostatic force of attraction
between cations and anions. Ionic bonding occurs easily between an
element that has a low ionization energy (s-block, easy to remove
electron) and one with high electron affinity (non-metals, halogens
and oxygen). The compound thus formed is called ionic compound.

Consider the transfer of an electron in the formation of potassium


chloride. The electron configurations of potassium and chlorine are:
K (19) [Ar] 4s1
Cl(17) [Ne] 3s2 3p5
Electron transfer can be shown below:
K ([Ar])4s1 + Cl([Ne] 3s2 3p5 K+ ([Ar]) Cl - ([Ar])

As a result of the electron transfer K+ and Cl are formed, each of


which has a stable noble gas (ns2 np6 ) configuration. The potassium
atom has lost its electron from its valence shell (4s) while chlorine
atom has gained that electron. This tendency of atoms to achieve
noble gas configuration gives octet rule which states that an atom
tends to lose or gain electron(s) in its valence shell until there are eight
electrons in its valence shell (two electrons in case of He
configuration). There are eight electrons (ns2 np6) in an anion and a
129
cation except in the first shell (n=1) where there are two (1s2)
electrons. An anion acquires the closed shell configuration of the
following noble gas and a cation acquires a closed shell configuration
of the preceding noble gas or [ ] nd10 or in heavier elements [ ] (n-
1)d10 ns2.

Stability of ionic compounds


A well-known example of an ionic
solid compound is sodium chloride,
which consists of two interpenetrating
lattices of Na+ and Cl ions arranged
in such a way that every ion of one
type is surrounded (in three
dimensional space) by six ions of
opposite charge. The driving force (reason) for the formation of ionic
compounds is the lattice energy for ions in their crystal arrangement.
The more truly ionic compounds have most favorable lattice energies:
Lattice energy is defined as the energy required to fully separate a
mole of ionic compound into its gaseous ions. Since the energy is
required to break the compound, lattice energy is always a positive
value.
For a given arrangement of ions, the lattice energy increases as their
radii decrease and increases as the charge on the ion increases. Table
8.1 shows the lattice energies of some compounds.

Table 8.1: Lattice Energies of Some ionic compounds

Compound Lattice energy Compound Lattice Energy


(kJ/mol) (kJ/mol)
LiF 1046 CsCl 654
LiCl 861 CsI 602
LiBr 818 MgCl2 2325
LiI 759 SrCl2 2128
KF 809 MgO 3798
KCl 700 CaO 3410
KBr 672 SrO 3211

Highest (most favorable) lattice energies are between small highly


charged ions.
130
Lewis Dot Symbols
In chemical bonding only valence electrons are taken into
consideration. These electron (s) can be shown by a dot () or a cross
() in a Lewis symbol. A Lewis symbol consists of the symbol of an
element and a dot or cross for each valence electron in an atom of the
element. Some examples are as follows:
.
H . , He : , . N : , . O . , : Cl . , K .

: :
: :
, Mg :
.
The formation of potassium chloride using Lewis dot symbols can be
written as:
-
:

:
K . + . Cl : K+ : Cl :
:

:
In many cases charges on the cation and anions are not the same. The
formation of calcium chloride and aluminum oxide shown below
illustrates this.
- -
:
:

: Cl :
: :

. Cl : : 2+
Ca : + 2 Cl : Ca
:
:

2-
: :

2 Al3+ 3 :O : Al2O3

Example 8.1:
Describe the formation of Mg3N2 using Lewis dot symbols.
Solution:
The formation of Mg3N2 involves participation of three magnesium atoms
and two nitrogen atoms.
3 Mg + N2 Mg3N2
131
Electron configurations of Mg and N atoms are:
Mg(12) 1s2 2s2 2p6 3s2 or [Ne] 3s2
N(7) 1s2 2s2 2p3 or [He] 2s2 2p3
A magnesium atom has two valence electrons. A nitrogen atom has five
valence electrons and hence needs three electrons to acquire a stable noble
gas configuration. In the process three magnesium atoms will transfer a
total of six valence electrons and the two nitrogen atoms will gain a total
of six valence electrons to acquire the noble gas configuration. The
formation of Mg3N2 using Lewis dot symbols can be depicted as follows.

..
.. .. ..

Mg
.. ..
N
Mg . Mg3N2
.
Mg N
..
8.2 Covalent Bonds
A covalent bond is formed by the sharing of one or more pair of
electrons. The compounds thus formed are called covalent
compounds. When the electrons are shared equally by the bonded
atoms in a molecule, single, double or triple bonds are formed. In
covalent bonds atoms tend to share electrons until a closed shell
configuration is obtained for each atom (octet rule, based on ns2 np6
configuration). Octet rule is strictly applicable to the elements of
period two only. First period elements (1s2) need only two electrons,
3rd, 4th, 5th periods elements (for example P, S, Cl, Br and As) can
have expanded octet but may also follow octet rule in some cases.

Single, double and triple bonds


Single bond
A single bond is formed when two bonded atoms share one pair of
electrons.
(i) Formation of a hydrogen molecule: In a hydrogen molecule,
each atom has one valence electron (1s1).

132
. + . . .

H(1s1) H(1s1) H2
Both electrons are shared by/attracted to both nuclei which is a
stronger force than nuclei-nuclei repulsion but since both 1s electrons
have n = 1, = 0 and m = 0, the electrons must be spin-paired.

H . + .H H .. H or H H
The single line represents the shared pair of electrons, the single
covalent bond, between two atoms.

The valence of an element gives the number of covalent bonds an


atom of the element forms. Thus, hydrogen atom has only one valence
electron; it is expected to form one covalent bond.
Thus, a halogen atom (such as chlorine) has seven valence electrons
and needs only one electron to complete octet so it is expected to form
one single covalent bond. Similarly, each nitrogen atom needs three
electrons to complete octet, so it is expected to form three single
covalent bonds.
In noble gases, atoms already have octet of the valence electrons, so
the formation of any single covalent bond is not expected.

(ii) Formation of a fluorine molecule: The electron configuration of


a fluorine atom is 1s2 2s2 2p5 or [He] 2s2 2s5. The electrons in 1s sub
shell are in lower energy (ground) level so they do not take part in
covalent bonding. The valence electrons (normally spin-paired) not
used in covalent bonding are called lone pair of electrons or non-
bonding electrons. The shared pair of electrons between two bonded
atoms are called bonding electrons.

133
bonding
electrons

.. .. .. ..
..

..
..
F F

..
More simply
.. F

..
.. F
.. ..
non bonding
non bonding electrons
electrons
Lone pair repulsion on fluorine atoms weakens F-F bond, making the
F2 molecule more reactive than Cl2, Br2, or I2, which having larger
atoms and the lone pairs are farther apart leading to lower repulsion.

Double bond
Two shared electron pairs form a double bond. For example, a double
bond between a carbon atom and an oxygen atom, C : : O C O or
a double bond between two carbon atoms, C : : C C C.

Triple bond
Three shared pairs of electrons form a triple bond. For example, a
triple bond between two carbon atoms or a triple bond between two
nitrogen atoms.

..
C .. C
..
..
..
..
..

N N
..

.. C C or N .. N
Other examples are:

.... -
.... .... C H; C .. N
..

..

; H
..

C
.. O
..

..
..

..

Steps for writing Lewis structures for polyatomic species:


1. Write the skeletal structure for the polyatomic species.
134
(a) In general the element with the lowest ionization energy (lowest
electronegativity) occupies the central position. Hydrogen and
halogen atoms usually occupy terminal (end) position. For
example, in the methane molecule carbon is the central atom and
hydrogen atoms are at the terminal.
(b) Bonded atoms are usually symmetric around the central atom.
For example:
2-
O

O O S O

S O
2-
SO2 = OSO SO3 = O O SO4 =
(c) Central atom is usually written first: NH4+, PO43- etc.
(d) In oxyacids, acidic hydrogen is bonded to the O atom.
H2SO4 (HO)2 SO2 ; H3PO4 (HO)3PO
HClO2 (HO) ClO ; HClO4 (HO)ClO3
2. Calculate the total number of valence electrons for the molecule
by adding up the number of valence electrons for each atom.
For example: In a methane (CH4) molecule total number of
valence electrons is C (4) + H (1 4) = 8.
In a polyatomic anion the number of electrons for each negative
charge is added up to this total. Thus for the SO42- ion, total
number of electrons is S (6) + O (4 6) + 2 = 32 electrons.
In a polyatomic cation, the number of electrons for each positive
charge is subtracted from the total. Thus for the NH4+ ion, total
number of electrons is N (5) + H (1 4) - 1 = 8 electrons.
3. Draw a single bond between the central atom and each of the
surrounding atoms or place a pair of electrons for each bond.
4. Distribute electrons to the atoms surrounding the central atom to
complete the octet. For hydrogen atoms only two electrons are
required. If the central atom does not complete its octet, multiple

135
bonds are written between the surrounding atoms and the central
atom with non-bonding electrons on the surrounding atoms.
5. Place any additional electron (s) on the central atom if they are
not involved in bonding; even if doing so results in more than an
octet.

Example 8.2:
Write the Lewis structures for (i) CBr4 (ii) HNO3 (iii) CO32-
Solution:
(i) CBr4 (Carbon tetrabromide)
Step 1: The skeletal structure of CBr4 is
Br

Br C Br

Br
Step 2: The total number of valence electrons in CBr4 is C (4) + Br (4 x
7) = 32
Step 3: Place one single bond or a pair of electrons between carbon atom
and each of the bromine atom.
Br
Br
:

Br : C : Br or Br C Br
:

Br
Br

Step 4: Distribute the remaining (32 8 = 24) electrons to the


surrounding atoms to complete their octet.

136
:
: Br :

:
: Br :
:

: :
:
:

: :
: Br : C : Br : or : Br C Br :

:
: :
: Br :
: Br :

:
Since all atoms in CBr4 molecule have a complete octet, steps 4 and 5 are
not needed.

(ii) HNO3 (Nitric acid)


To write the Lewis structure for HNO3, follow the steps as described
above.

:
: :

:
O : : :O :
: :

H O N H O N
: :

:
O
:O :

(iii) CO32- (Carbonate ion)


2- 2- 2-

:
:

O :O : :O :
:
:

O C
:O : C :O : C
O
: O: :O :
:

8.3 Resonance and Resonance Structures


Sometimes two or more valid Lewis structures satisfying the octet rule
are possible for molecules or ions. The term resonance is used to
show such a situation. The actual structure is the average of the
contributing Lewis structures, called a resonance hybrid. A double

137
ended arrow ( ) is used between the possible Lewis structures to
show a resonance form of a molecule or ion.

Consider, for example, the nitrate ion NO3. The Lewis dot structure
satisfying the octet rule can be written as follows.
-
: :
O

:
: O N O
:

:
:

According to this structure, the ion contains two NO single bonds


(142 pm) and one (N O) double bond (122 pm) but the experimental
data show the presence of three N O bonds (126 pm). The double
bond can be written in any of the other two equivalent locations giving
the following three valid equivalent resonance structures for the
nitrate ion.
- - -
:

:
: :
O : : : :
O O
:

:
:

:
: : : :
O N O O N O O N O
:
:

:
:

Electrons and multiple bonds involved in resonance structures are said


to be delocalized. The equivalent structures have the same energy and
contribute equally to the resonance hybrid. Non-equivalent structures
have different energies; lower energy structure contributes more to the
resonance hybrid. Similar structures can be written for sulfur trioxide
SO3, and for the carbonate ion CO32.

Example 8.3:

Write three non-equivalent resonance structures for SCN (thiocynate
ion.)
Solution:

138
Three possible non-equivalent structures for the thiocynate ion are:
_ _ _

:
:

:
:
: S C N
: : S C N
: S C N

:
:
:
:

8.4 Formal Charge


The formal charge on an atom is the electric charge it would have if
all bonding electrons were shared equally with its bonded neighbors.
The formal charge on an atom in a molecule or an ion is calculated
using the following formula.
Formal Charge (FC)= [{Total no. of valence electrons in the free
atom} -{Total no. of non-bonding electrons}
{ Total no. of bonding electrons}]
The sum of all the formal charges must be equal to the overall charge
of the molecule or ion.
Formal charge(s) are used to predict most reasonable resonance
structure for a given compound. Best Lewis structure(s) have formal
charge(s) close to zero; good Lewis structure must have formal
charge(s) that are reasonable with regard to relative electronegativity
values, negative charge should be on the most electronegative atom.

Example 8.4:
Write a Lewis structure for the NO3 ion assuming that the ion obeys the
octet rule and then find the formal charge(s) on each atom.
Solution:
Lewis structure for the nitrate ion can be written as follows:
-
: :
O
:

: O N O
:
:
:

The formal charge on each individual atom can be calculated as follows:


1. Central nitrogen atom :
139
Total number of valence electrons in a free N atom = 5
Total number of non-bonding electrons = 0
Total number of bonding electrons = 8
Hence, the formal charge (FC) on N atom = 5 0 - (8) = +1
2. Oxygen atom bonded with a single bond ( O):
Total number of valence electron in a free O atom = 6
Total number of non-bonding electrons = 6
Total number of bonding electrons = 2
Hence the formal charge on the oxygen atom= 6-6-(2) = -1
3. Oxygen atom bonded with double bond (O):
Total number of valence electrons in a free O atom = 6
Total number of non-bonding electrons = 4
Total number of bonding electrons = 4
Hence the formal charge on the oxygen atom = 6-4-(4)=0
The formal charges on each atom in a Lewis structure can be shown as
follows.
(0) -
: :
O
(-1) (-1)
:

: O N O
:
:
:

(+1)

Example 8.5:
Find the formal charge on each atom in the SCN ion.
Solution:
The following are the three resonance structures for the SCN ion.
_ _ _
:
:

:
:

: S C N
: : S C N
: S C N
:
:
:
:

(a) (b) (c)

Charges on each Lewis structure can be calculated using the above


formula.

140
Atom Structure (a) Structure (b) Structure(c)
S FC = 6-6-(2) = -1 6-2-(6) = +1 6-4-2 = 0
C FC = 4-0 -(8) = 0 4-0-(8) = 0 4-0-4 = 0
N FC = 5-8-(6) = 0 5-6-(2) = -2 5-4-2 = -1

Three resonance structures with formal charges can be written as follows.


_ (0) _
(-1) (0) (0) (+1) (0) (-2) _ (0) (-1)

:
:

:
:
: S C N
: : S C N
: S C N

:
:
:
:

8.5 Electronegativity and Bond Polarity

Electronegativity (A) of an atom is the ability of an atom in a


molecule to attract (pull) electrons to itself in a covalent bond. Several
different scales have been used to measure electronegativity of atoms.
Millikan electronegativity scale theoretically suggested that
electronegativity of an atom is half the sum of its electron affinity
(EA) and ionization energy (IE) i.e (A) = ( EA + IE). However,
this scale has its limitations. A more widely used one is the Linus
Paulings electronegativity scale. Electronegativity decreases down
the group and increases from left to right across a period (except noble
gases) in the periodic table. Metals have low electronegativity values
(least electronegative) and are electropositive and non-metals are the
most electronegative elements.

The difference in electronegativity values of two bonded atoms on a


4.0 Pauling scale gives a rough measure of the expected polarity of a
bond. When this difference is small, the bond is non-polar. When it is
large, the bond is polar. The following procedure is used to predict the
type and the polarity of the bond.
If the electronegativity values of two atoms, A and B in a molecule (A
B) are represented by A and B respectively, then difference in their
electronegativity values () will be given by A - B

141
If, 0 , the bond is purely covalent ( equally shared electron
pair) and non-polar.
0 < < 2, increasingly polar covalent bond (unequal sharing of
electron pair)
2, the bond is ionic (electron transfer).
The following examples will illustrate the use of this scale.
NaCl = 3.2 0.93 = 2.3, ionic bond
CsF = 4.0 0.79 = 3.2, more ionic bond
Al-Cl = 3.2 1.60 = 1.6, highly polar but ionic bond
Al- F = 4.0 1.60 = 2.4, ionic bond (AlF3)
CH = 2.2 2.60 = 0.4, essentially non polar

8.6 Exceptions to the Octet Rule


When the octet rule is obeyed, eight electrons in a valence shell make
a noble gas (ns2 np6) configuration; this works best for the elements in
the second period (Li through F) of the Periodic Table. When the
central atom in a molecule is able to accommodate more than eight
electrons in empty d- orbitals, the electrons in such expanded valence
shell or expanded octet may be used by the central atom to form
additional bonds. Some of the period 3 (n = 3) and higher elements (P,
S, As, Se, Sb, Te, Cl, Br, I) have empty d-orbitals with larger atomic
size and can have more than eight electrons (expanded octet).
Elements with expanded octet also show variable co-valences and
have the ability to form different number of covalent bonds. For
example, phosphorous forms 3 bonds as in PCl 3, 4 (in PCl4+), 5 (in
PF5) or 6 (in PCl6).

The following examples will illustrate the use of expanded octet.


1. Molecules in which central atom has more than eight electrons,
e.g. PF5 , PCl5 PCl6 , SF6 and IF7.
Lewis structures of PF5 and SF6 are shown below in which the
central atoms P and S have 10 and 12 valence electrons
respectively.

142
.. ..
..

..
..
.. F F

..
.. ..
.. ..

..
.. F

..
F

.. ..
.. P
F
.. ..
F
.. S ..
.. ..
..F

..
F
..

..
..
F
..F ..

..
..
..F
2. Molecules in which additional lone pairs of electrons are placed
on the central atom.
Lewis structures of PCl3 and XeF4 shown below with one and two
lone pairs of electrons on the central atom P and Xe respectively.

.. ..
..

..
..

..
.. .. F

.. .. Xe
F
..

P Cl
.. .. ..F
..

Cl
.. ..

..
..
..
F
..
..
..

Cl
..
3. Sometimes expanded octet provides better resonance structures
than an octet. For example , SO42, SO2
(a) SO42
2- 2- 2-
(-1) (-1) (-1)
.. .. ..
..
..

..

O (0) O
..

(-1) (0) O
..
..

(-1)
.. .. .. ..
(-1)
.. ..(0)
..
..

..

S(+2) O O S O S O
(+1) ..
O O
.. .. .. .. (0) ..
..
..

..
..
..

..

O O O
.. .. ..
(-1) (-1) (-1)
(a) (b) (c)
Atom Structure (a) Structure (b) Structure (c)
S FC = 6-0-4 = +2 6-0-5 = +1 6-0-6 = 0
O FC = 6-6 -1= -1 6-6-1 = -1 6-6-1 = -1
O FC = N/A 6 -4-2 = 0 6-4-2 = 0

143
Structure (c) is the best structure because negative charges are on
the more electronegative oxygen atom.

(b) SO2

(0) (+1) (-1) (-1) (+1) (0) (0) (0) (0)


.. .. .. .. .. .. .. .. ..
..
O

..
S O O S O O S O
.. .. .. .. .. ..
(a) (b) (c)
Structures (a) and (b) are equivalent, (c) is a non-equivalent
structure.
Atom Structure (a) Structure (b) Structure (c)
S FC = 6-2-3 = +1 6-2-3 = +1 6-2-4 = 0
O FC = 6-6-1 = -1 6-6-1 = -1 6-4-2 = 0
O FC = 6-4-2 = 0 6-4-2 = 0 6-4-2 = 0
Structure (c) is the best Lewis structure because the sum of all the
charges is zero.

8.7 Exercises
8.1. (a) What is the Lewis symbol of an element. Give an example.
(b) What is meant by the octet rule. Explain with an example.
(c) What is a resonance hybrid? Show using an example.
(d) What is an expanded octet? Mention three elements which
could show this in their compounds.

8.2. Which compound most likely has the largest lattice energy:
(a) MgS or MgO (b) NaCl or NaBr (c) NaCl or MgS

8.3. Draw the Lewis structure for:


(a) CS2 (b) XeF4 (c) O3 (d) CO32 (e) CH3Cl

8.4. For the following species, draw the Lewis structure (if resonance is
possible, draw the non-equivalent resonance structures), calculate
the formal charges on all atoms, and indicate which structure is
most reasonable based on the calculated formal charges.
(a) OCN (b) PO43 (c) SO42

144
8.5. (a) Arrange the following elements in the expected order of
increasing electronegativity: In, Se, Sb, S.
(b) Which one is the most electronegative element: N, Sn, B, F, S?

145
CHAPTER 9
Molecular Geometry
Molecules are three-dimensional objects that occupy a three-
dimensional space. In general, only the smallest molecules can have a
fixed geometrical shape. In most molecules, those parts joined by
single bonds can rotate with respect to each other, giving rise to many
different geometric forms. However, the overall geometry surrounding
a given atom that is covalently bound to its neighbors is constant.
The Lewis electron-dot structures have no geometrical significance
other than depicting the order in which the various atoms are
connected to one another. Nevertheless, a slight extension of the
simple shared-electron pair concept is capable of rationalizing and
predicting the geometry of the bonds around a given atom in a wide
variety of situations.

9.1 Valence-Shell Electron Pair Repulsion Model


The valence shell electron pair repulsion (VSEPR) model focuses
on the bonding and nonbonding electron pairs present in the outermost
(valence) shell of an atom that connects with two or more other
atoms.
The covalent model of chemical bonding assumes that the electron
pairs responsible for bonding are concentrated into the region of space
between the bonded atoms. The fundamental idea of VSEPR theory is
that these regions of negative electric charge will repel each other,
causing them (and thus the chemical bonds that they form) to stay as
far apart as possible.

146
Thus the two electron clouds contained in a simple triatomic molecule
AX2 will extend out in opposite directions; an angular separation of
180 places the two bonding orbitals as far away from each other they
can get. Therefore the two chemical bonds are expected to extend in
opposite directions, producing a linear molecule.
If the central atom also contains one or more pairs of nonbonding
electrons, these additional regions of negative charge will behave very
much like those associated with the bonded atoms. The orbitals
containing the various bonding and nonbonding pairs in the valence
shell will extend out from the central atom in directions that minimize
their mutual repulsions.
If the central atom possesses partially occupied d-orbitals, it may be
able to accommodate five or six electron pairs, forming an expanded
octet.

9.2 Molecular Shapes


Consider the molecules that contain atoms of two elements A and X of
which A is the central atom surrounded by other atoms. Suppose that
the general formula of molecules AXn. In most cases, the value of n is
in between 2 to 6. The geometry of such molecules can be predicted
by considering the number of bonding and lone pairs of electrons
around the central atom.

AX2 molecule (linear): In such molecules, the central atom (A) is


bonded to two atoms (X) through two pairs of electrons.
As stated above, a simple triatomic molecule of the type AX2 has its
two bonding orbitals 180 apart, producing a molecule having linear
geometry as shown below:

Examples of triatomic molecules for which VSEPR theory predicts a


linear shape are BeCl2 and CO2. A double bond is present in Lewis
electron dot formula for carbon dioxide. This makes no difference to
VSEPR theory because all double and triple bonds present in any
147
molecule are treated as a single bond. The central carbon atom is still
joined to two other atoms, and the electron clouds that connect the two
oxygen atoms are 180 apart.

AX3 molecules (trigonal planar): The central atom (A) is surrounded


by three pairs of bonding electrons (X) and has no lone pair of
electrons.
In an AX3 molecule such as BF3, there are three regions of electron
density extending out from the central atom. The repulsion between
these will be at a minimum when the angle between any two is
(360 3) = 120. This requires that all four atoms be in the same
plane; the resulting shape is called trigonal planar, or simply trigonal.

Trigonal planar molecules with lone pairs


AX2E molecules: If only two of the three atoms are bonded leaving
one lone pair of electrons (E) around the central atom (A). The
geometry of such molecules is bent (angular) and the bond angle
between the atoms is 120o. Examples, SO2, O3.

..
A
X X
AX4 molecules (tetrahedral): In such molecules, the central atom (A)
has four pairs of bonding electrons and has no lone pair of electrons
around it. The molecular geometry is tetrahedral.
148
For example, methane, CH4, contains a carbon atom bonded to four
hydrogen atoms. What bond angle would lead to the greatest possible
separation between the electrons clouds associated with these bonds?
In analogy with the above cases, where the bond angles were 360/2 =
180 (AX2) and 360/3 = 120, you (AX3) might guess 360/4 = 90;
if so, you would be wrong. The latter calculation would be correct if
all the atoms were constrained to be in the same plane. Consequently,
the four equivalent bonds will point in four geometrically equivalent
directions in three dimensions corresponding to the four corners of a
tetrahedron centered on the carbon atom. The angle between any two
bonds will be 109.5.

Tetrahedral molecules with lone pairs


(1) AX3E molecules: Central atom (A) is bonded to three atoms
leaving one lone pair of electrons on the central atom. The molecular
geometry is trigonal pyramid.
For example, the electron-dot structure of NH3 places one pair of
nonbonding electrons in the valence shell of the nitrogen atom. This
means that there are three bonded atoms and one lone pair. Therefore
it can be predicted that the three hydrogen atom will lie at the corners
of a tetrahedron centered on the nitrogen atom. The lone pair orbital
will point toward the fourth corner of the tetrahedron, but since that
position will be vacant, the NH3 molecule itself cannot be tetrahedral.
Instead, it assumes a pyramidal shape. More precisely, the shape is
that of a trigonal pyramid (i.e., a pyramid having a triangular base).
The hydrogen atoms are all in the same plane, with the nitrogen above
(or below, or to the side; molecules of course dont know anything

149
about above or below!) The fatter orbital containing the non-
bonding electrons pushes the bonding orbitals together slightly,
making the HNH bond angles 109.5o.

(2) AX2E2 molecules: The central atom (A) has two pairs of
bonding electrons and two lone pairs of electrons. Such molecules
show a bent (angular) geometry.
For example, in the water molecule, the central atom is O, and the
Lewis electron dot formula predicts that there will be two pairs of
nonbonding electrons. The oxygen atom will therefore be tetrahedrally
bonded, meaning that it sits at the center of the tetrahedron as shown
below. Two of the bonded positions are occupied by the shared
electron-pairs that constitute the OH bonds, and the other two by the
non-bonding pairs. Thus although the oxygen atom is tetrahedrally
coordinated, the bonding geometry (shape) of the H2O molecule is
described as bent.

There is an important difference between bonding and non-bonding


electron orbitals. Because a nonbonding orbital has no atomic nucleus
at its far end to draw the electron cloud toward it, the charge in such
an orbital will be concentrated closer to the central atom. As a
consequence, nonbonding orbitals exert more repulsion on other
orbitals than do bonding orbitals. Thus in H2O, the two nonbonding

150
orbitals push the bonding orbitals closer together, making the HOH
angle 104.5 instead of the tetrahedral angle of 109.5.

AX5 molecules (trigonal bipyramid): The central atom (A) is bonded


to five other atoms (X). The geometry of such molecules is trigonal
bipyramidal. Compounds of the type AX5 are formed by some of the
elements in Group 15 of the periodic table; PCl5 and AsF5 are
examples.
The shape of PCl5 and similar molecules is a trigonal bipyramid. This
consists simply of two triangular-base pyramids joined base-to-base.
Three of the chlorine atoms are in the plane of the central phosphorus
atom (equatorial positions), while the other two atoms are above and
below this plane (axial positions). Equatorial and axial atoms have
different geometrical relationships to their neighbors, and thus differ
slightly in their chemical behavior. The angle between equatorial
bonds is 120o, whereas the bond between two axial positions is 180o.

Trigonal bipyramidal molecules with lone pairs


(1) AX4E molecules: In an AX4E molecule, the central atom A is
bonded to four other atoms (X) and to one nonbonding electron pair
(E). The molecule, such as SF4 will have a see-saw shape as shown
151
below. The bond angles between two bonded atoms are 120 o and 180o.
Other examples are: IF4+ , XeO2F2 .
Substitution of more nonbonding pairs for bonded atoms reduces the
triangular bipyramid shape to even simpler molecular shapes, as
shown in the following cases.

(2) AX3E2 molecules: The central atom is (A) has three bonding
pairs of electrons (X) and two lone pair of electrons (E). The bond
angle between two bonded atoms is 90o and the molecule (ICl3)
shows T-shaped geometry as given below. Other examples are ClF3.

(3) AX2E3 molecules: The central atom (A) has two pairs of
bonding electrons (X) and three lone pairs of electrons (E). The
geometry of such molecules or ions such as I3 (shown below) is linear
with 180o bond angle. Other examples are XeF2.

AX6 molecules (octahedral): In such molecules central atom (A) is


surrounded by six bonding pairs of electrons. The molecular geometry
is octahedral (shown below) with bond angle between bonded atoms is
90o (cis) and 180o (trans). Examples: SF6 and PF6- .
X

X A X

X
X

Just as four electron pairs experience the minimum repulsion when


they are directed toward the corners of a tetrahedron, six electron pairs
will try to point toward the corners of an octahedron which is simply
two square-based pyramids joined base to base.

152
Octahedral molecules with lone pairs
(1) AX5E molecules: The central atom (A) has five pairs of
bonding electrons and one lone pair of electrons. The molecule (ClF 5)
has square pyramid geometry as shown below. Other examples: BrF5 ,
XeOF4

(2) AX4E2 molecules: The central atom (A) has four pairs of
bonding electrons and two lone pairs of electrons. The geometry of
such molecules (e.g.XeF4) is square planar. Other examples : ICl4-.

Example 9.1:
Use VSPER theory to predict the geometry of: (i) AsH3; (ii) AlCl4-
Solution:
(i) AsH3 molecule
Total number of valence electrons in AsH3 molecule is 8 (4 pairs of
electrons). Therefore the Lewis structure of AsH3 molecule is

..
H As H

H
The central atom As has three pairs of bonding (X) and one pair of lone
pair (E) of electrons. This combination is similar to AX3E molecule which
corresponds to trigonal pyramid geometry. Therefore the geometry of
AsH3 molecule is trigonal pyramid.
153
(ii) AlCl4 ion:
Follow the above procedure. The central atom, aluminum, has four
bonding pairs of electrons having no lone pair of electrons. Hence the ion
is similar to AX4 molecule. Therefore the geometry of AlCl4- ion is
tetrahedral.

9.3 Molecular Shape and Polarity


Covalent bonds can be polar or non-polar. When non-identical atoms
are joined in a covalent bond, the electron pair will be attracted more
strongly to the atom that has the higher electronegativity. As a
consequence, the electrons will not be shared equally; the center of the
negative charges in the molecule will be displaced from the center of
positive charge. Such bonds are said to be polar and to possess partial
ionic character, and they may confer a polar nature on the molecule as
a whole and have a partial separation of charge. The overall molecule
can be polar or non-polar depending on its shape.
+ -
HF

When a solution of polar molecules is placed between two oppositely-


charged plates, they will tend to align themselves along the direction
of the electric field.

The polarity of a molecule is usually determined by its dipole moment


(). The dipole moment corresponding to an individual bond (or to a
diatomic molecule) is given by the product of the quantity of charge
displaced q and the bond length r:
=qx r

154
In SI units, q (charge) is expressed in coulombs and r in meters. The
unit of the dipole moment () is Debey (D) and is expressed in C.m.
If one entire electron charge is displaced by 100 pm (a typical bond
length), then
= (1.6022 1019 C) (1010 m) = 1.6 1029 C-m = 4.8 D

The dipole moment contribution from each bond (bond dipole) in a


molecule is shown by an arrow pointing towards the more
electronegative atom. Each bond dipole is a vector quantity which has
both magnitude and direction. The two bond dipoles of equal
magnitude and opposite directions cancel each other and the molecule
has no dipole moment ( = 0). Hence all homonuclear diatomic
molecules (O2, H2, F2 etc) does not possess dipole moment and are
non-polar and the bonds (O O ; H H; F F etc) are non- polar
bonds. A molecule that possesses a dipole moment ( 0) is a polar
molecule. Hence, heteronuclear diatomic molecules (HCl; CO ; NO
etc) are polar molecules and the bonds are polar covalent bonds.
Dipole moment in poly atomic molecules depends upon the geometry
and the polarity of the molecule, However, if polar bonds are present,
the molecules itself may not be a polar molecule.

Consider the following examples:


(1) Carbon dioxide (CO2) molecule
Carbon dioxide molecule has a linear geometry (AX2) with polar
bonds. The bond dipoles in the molecule point in opposite directions
(as shown below) and thus cancel each other giving zero dipole
moment for the molecule. As a result CO2 is a non-polar molecule.
The zero dipole moment of CO2 is one of the simplest experimental
methods of demonstrating the linear shape of this molecule.

- + + -
|
|
..
..

C
.O. .O.

155
(2) Water (H2O) molecule
Water molecule has a bent geometry and has a very large dipole
moment which results from the two polar HO bonds oriented at an
angle of 104.5. The non-bonding pairs of electrons on oxygen are a
contributing factor to the high polarity of the water molecule. The
bond dipoles point from the hydrogen atom to the more
electronegative oxygen atom. The individual bond dipoles do not
cancel completely and give a non-zero dipole moment. Thus water is a
polar molecule.
Resultant or
net dipole moment

O
..
..|

|

|
H H

The relationship between dipole moment and the geometry of the


molecules with some examples is given in the table below.

Table 9.1: Relationship between geometry and dipolemoment for


some molecules
Formula Dipole moment Molecular geometry Examples
AX2 zero Linear CO2
AX3 zero Trigonal planar BCl3, BF3, GaCl3
AX2E non zero Bent SO2
AX4 zero Tetrahedral CH4, NH4+, CCl4
AX3E non zero Trigonal pyramid NH3, PF3
AX2E2 non zero Bent H2O
AX5 zero Trigonal bipyramid PCl5.SbCl5
See-saw (distorted
AX4E non zero IF4+, SF4, TeCl4
tetrahedral)
AX3E2 non zero T-shape ClF3
AX2E3 zero Linear I3, XeF2
AX6 zero Octahedral SF6
AX5E non zero Square pyramid BrF5, XeOF4
AX4E2 zero Square planar ICl4 , XeF4

156
9.4 Exercises
9.1. What is the geometry of the following species (also indicate the
bond angles):
(a) PF6 (b) ClF3 (c) BF4 (d) H2S (e) NO3

9.2. (a) Which bond is the most polar: CF, CBr, CN, CO, CS?
(b) Which of the following molecules is expected to be polar: SF6,
PF5, CF4, NF3, OF2, HF?
(c) Which one of the following chlorides is expected to be non-
polar: HCl, ClO2, NCl3, SiCl4, ICl5?

9.3. Give the number of bonds, bonds and lone pairs of valence
electrons in a molecule of
(a) HCN (b) CO2 (c) N2 (d) CH2F2 (e) NH3

9.4. What is the expected hybridization on the indicated atom in the


following?
(a) P atom in PCl5,
(b) C atom in CO,
(c) S atom in SF4,
(d) S atom in SO2.

157
CHAPTER 10
Introduction to Organic Chemistry
10.1 Carbon-based Materials

The word organic is derived from the Greek word organ meaning life.
It is interesting to know how a word signifying life relates to carbon-
based materials.
In ancient times, chemists curiosity about living beings led them to
study the substances obtained from animals and plants. It soon
became clear that compounds obtained from living beings were
different from those obtained from non-living things such as minerals.
Most compounds from animals and plants were made up of carbon,
hydrogen, oxygen, nitrogen, and sometimes sulphur and phosphorus.
Carbon was present in all the compounds. The chemists thus believed
that these compounds are only formed inside the living beings and that
is why the term organic was used to describe such compounds. This
belief gave birth to the vital force theory according to which some
vital force was present in the living beings which was responsible
for the synthesis of these carbon-based materials. This belief was
disproved in 1828 by a German Chemist Friedrich Wohler, only 28
years old at that time, who synthesized an organic compound urea, a
constituent of urine, in the laboratory by heating ammonium cyanate,
a compound occurring in minerals.
The synthesis of urea led to the development of a new era in the study
of organic compounds. Chemists started investigating method to
synthesize other compounds isolated from plants and animals.
Synthesis of compounds for the study of their properties was
necessary because many time the amounts extracted from plants and
animals were small. Soon thousands of organic compounds were
synthesized. One might be led to think that the word organic is now
a misnomer because its origin is not limited to living beings. This is,
however, not true. The use of this word is justified in the sense that
carbon compounds form the basis of our lives.
The major constituents of living beings, amino acids, carbohydrates,
fat, nucleic acids and hormones, are organic compounds. A simple
example of each of a carbohydrate, an amino acid and a fat is shown
in Fig 10.1. Almost all the reactions in plants and animals involve

158
organic compounds.

CHO
H OH H
HO H O H OC(O)(CH2)14CH3
H OH H2N CHC OH H OC(O)(CH2)14CH3
H OH CH3 H OC(O)(CH2)14CH3
CH2OH H

D-glucose alanine Tripamitin


(a carbohydrate) (an amino acid) (a fat from palm oil)

Figure 10.1: Examples of a carbohydrate, an amino acid and a fat

Organic chemistry is related to our day-to-day life more than any


other branch of science. The papers of our books, the wood of
furniture, our clothing, most of the medicines, gasoline, plastic and
perfumes are all organic compounds.

Figure 10.2: Examples of organic compounds used in every-day life

Some common examples (Figure 10.2) include acetic acid, the


principal constituent of vinegar which is used as preservative in
pickles. You must be familiar with spirit which is used as an
antiseptic. Spirit is an alcohol, ethanol, which is also used in
preparation of many alcoholic beverages. Diethyl ether, commonly
known as ether, besides being used as solvent in chemical
laboratories, has been used as a general anaesthetic compound.
Nowadays its use is rather limited due to side-effects. Acetylsalicylic
acid tablets are sold in pharmacies under the name aspirin as pain-
killer. A halogen containing alkene such as vinyl chloride is used for
the preparation of polyvinyl chloride (PVC) which is used as floor
tiles or sheets. DDT, a halogen derivative of hydrocarbon is a well-
known insecticide, and urea is a well-known fertilizer.

159
The present day definition of organic chemistry is that it is the branch
of chemistry, which deals with the chemistry of carbon compounds.
Thus, all organic compounds have carbon as a central atom. This
definition has a broad scope. It takes into account both synthetic
compounds and the compounds obtained from nature. However, we
exclude carbon oxides, carbonate and bicarbonates because they
greatly resemble inorganic compounds.

10.2 Structural Features of Organic Compounds

Besides having carbon in common, organic compounds have other


characteristics, some of which are listed below:
1. Tetra-covalence of carbon: Each carbon has four covalent bonds.
2. Catenation: It is the ability of organic compounds to form long
chains.
3. Isomerism: Many organic compounds exhibit isomerism. There
are many compounds known with the same molecular formula
but different structures. Sometimes compounds with identical
structures show different properties due to different orientation of
groups attached to carbon in the space.
4. Homologous series: Compounds of a particular class, that show
similar chemical properties, can be arranged in order of
increasing molecular weight in a series. Each member of the
series differs in molecular formula from the member above or
below by CH2. The series of compounds so arranged is called
homologous series.
The first characteristic of organic compounds, the tetra-covalency of
carbon, is briefly explained in the sections below. The other three
aspects will be discussed under the various sub-topics.

Bonding in carbon-based materials: hybridization


Carbon is bonded to other atoms by four covalent bonds in its
compounds. The electronic configuration of carbon however shows
that it has only two unpaired electrons (in 2p) in its valence shell and
so it may form only two covalent bonds. For a carbon atom to form
four covalent bonds, an electron from 2s is promoted to 2p resulting

160
into four half-filled orbitals in the valence shell. Each of these half-
filled orbitals can now accommodate an additional electron to become
completely filled and form four covalent bonds.
The study of the simplest possible organic compound methane, CH4,
showed that all four covalent bonds were identical having the same
length, energy and bond angle even though the bond forming orbitals
2s and 2p had different energy levels. It led to the concept of
hybridization. It is the mixing of atomic orbitals of nearly the same
energy and redistribution of energy among themselves leading to the
formation of an equal number of hybrid orbitals having the same
energy. Three types of hybridization are shown by carbon.

sp3 hybridization
One 2s orbital and three 2p orbitals hybridize to form four hybrid
orbitals called sp3 hybrid orbitals, all having the same energy. They
arrange themselves equally in space leading to a tetrahedron shape.
These hybridized orbitals are used to form -bonds (sigma bonds)
with other atoms. All single bonds in organic compounds are thus
sigma bonds. The bond angle between any of the two bonds formed is
109.5o. See figure 10.3.
H

C
H H
H
H H
H
sp3
bonds

Figure 10.3: sp3 hybridization

sp2 hybridization
One 2s orbital and two 2p orbitals hybridize to form three equivalent
hybrid orbitals called sp2 hybrid orbitals. They arrange themselves in
a trigonal planar shape in space and form -bonds with other atoms.
The angle between any two bonds is 120o.

161
The remaining 2p orbital, that is not hybridized, exist perpendicular to
the plane of hybrid orbitals. It overlaps sideways (laterally) with
another non-hybridized 2p orbital parallel to it, leading to the
formation of a covalent bond that is called a -bond. Thus, in the
structure shown below, one of the two bonds between two carbon
atoms is a sigma bond while the other is a pi bond. The pi bond is
loosely bound due to sidewise overlapping and hence the electron pair
involved in a pi bond formation is readily available for an electron
deficient species. Therefore, hydrocarbons containing double bonds
or triple bonds are generally more reactive in comparison to saturated
hydrocarbons. See figure 10.4.

H H
C C
H H


sp2

Figure 10.4: sp2 hybridization

sp hybridization
One 2s orbital and only one of the 2p orbitals hybridize resulting into
the formation of two equivalent hybrid orbitals called sp hybrid
orbitals. These orbitals extend in opposite directions from the centre,
taking a linear geometry. They are able to form two bonds with
bond angle of 180o. The remaining two 2p orbitals, which are
perpendicular to each other and also perpendicular to the hybrid
orbitals, overlap sideways with properly aligned (parallel) orbitals on
adjacent atoms to form two pi bonds. See figure 10.5.
-bond, sideways overlap
sp hybrid carbon
p-orbital
H H H C C H
sp orbital

bond internuclear overlap

Figure 10.5: sp hybridisation


162
In summary, the geometry and hybridization of carbon atoms in
organic compounds can be described as follows:
Hybridization:
a. sp3 hybridised, four single bonds; bond angle of 109.5o.
b. sp2 hybridised, one double bond; bond angle of 120o.
c. sp hybridised, two double bonds; bond angle of 180o.
d. sp hybridised, a triple bond; bond angle of 180o.

Bond types:
a. All single bonds are sigma bonds.
b. A double bond is made up of a sigma and a pi bond.
c. A triple bond is made up of a sigma and two pi bonds.

Bond length and bond strength


The distance between the nuclei of two atoms in a bond is called bond
length. It depends on the atoms involved in the bonding and the
number of bonding electron pairs. More electron pairs involved in
bonding leads to a shorter bond length. This is because the electron
pairs pull the nuclei towards themselves, reducing the distance
between two nuclei. Thus, a carbon-carbon triple bond will be the
shortest, and a carbon-carbon single bond will be the longest:
C C > C C > C C

longest shortest

Bond strength is the energy required to break the bond. Thus, more
energy will be required to break a strong bond, and less energy will be
required to break a weak bond. The bond strength increases as the
bond length decreases. Thus, least energy is required to break a
carbon-carbon single bond and highest energy is required to break the
carbon-carbon triple bond:
C C < C C < C C

weakest strongest

163
10.3 A Simple Classification of Organic Compounds

Millions of organic compounds are known. Their study would be


impossibly complex if it was not for their classification based on the
types of atoms in them and their structural features. A simplistic
approach is to put all the organic compounds into two groups, one
with compounds containing only carbon and hydrogen atoms called
hydrocarbons, and the other comprising compounds containing other
atoms as well besides carbon and hydrogen. Both groups can be
further divided into several classes based on structural features such as
the presence of carbon-carbon double or triple bonds and the presence
of certain functional groups, which impart chemical properties to a
compound.

A functional group is an atom or a group of atoms containing at least


one heteroatom (an atom other than carbon) present in the compound,
and is responsible for the properties of the compound. Some chemists
treat double and triple bonds also as functional groups because their
presence in the compound gives specific properties to the compounds.
A brief classification of the main organic compounds is shown in
Chart 10.1
Organic Compounds

Hydrocarbons Compounds
(contain only C and H) containing other atoms
besides C and H

Open Chain Cyclic


(acyclic, aliphatic)

Saturated Unsaturated Saturated Unsaturaed


(alkanes) (alkenes, alkynes) (cycloalkanes)

Cycloalkenes Aromatic
Cycloalkynes (arenes)

Chart 10.1: A simple classification of organic compounds

164
10.4 Hydrocarbons

Hydrocarbons, the compounds containing only carbon and hydrogen,


are further divided into open chain (acyclic) and cyclic hydrocarbons
based on their molecular framework. Acyclic hydrocarbons are also
called aliphatic (Greek word aleipher meaning fat) hydrocarbons
because many compounds having long chains were isolated from
animal fat. These are either saturated (alkanes) or unsaturated
(alkenes and alkynes). The cyclic analogs of aliphatic hydrocarbons
are called alicyclic hydrocarbons. Another class of cyclic
hydrocarbons, aromatic hydrocarbons, differs from the rest
significantly in properties and will be studied separately.

A. Alkanes
Alkanes are saturated hydrocarbons, meaning they contain single bond
only. They are also called paraffins due to their little affinity towards
other reagents (Latin, Para = little; ffin = affinity). The general
molecular formula of this class of compounds is CnH2n+2. The
simplest compound with n = 1 has molecular formula CH4 and is
called methane.
The names of all alkanes end with ane. Before ane, the prefixes
meth, prop, but, pent, hex and so on are used for three,
four, five and six carbon atoms, respectively (see table 10.1).

Table 10.1: Naming of alkanes

n Compounds Molecular formula


1 Methane CH4
2 Ethane C2H6
3 Propane C3H8
4 Butane C4H10
5 Pentane C5H12
6 Hexane C6H14
7 Heptane C7H16
8 Octane C8H18
9 Nonane C9H20
10 Decane C10H22
165
Homologous series
A careful look at the above table showing the molecular formula of
alkanes reveals that a member in the series differs in molecular
formula from the member above or below it by CH2. The compounds
in Table 10.1 are said to constitute a homologous series. You will
observe a trend in the physical properties of the compounds of one
homologous series.

Structural formula: Straight-chain and branched-chain


hydrocarbons
In order to study the chemical properties of not only hydrocarbons but
other organic compounds as well, the knowledge of molecular formula
alone is not sufficient. One needs to know the structural formula or
simply the structure of the compounds.
There are various ways of representing a structure (see figure 10.6).
For example, a complete structural formula shows all the atoms and
bonds between them. It is, however, not always necessary to draw a
complete structural formula; instead, either a condensed or an
abbreviated structural formula is drawn. This shows the groups of
atoms and bonds between carbon atoms only, though the latter is not
essential. Alternatively, the line formula, which shows only the bond
between two carbon atoms or between the atoms heavier than
hydrogen is used. Even the carbon is not shown; it is understood to be
on every corner and at every end. In practice, it is not always
necessary to write a pure condensed or pure line formula. Often, only
that part of the structure is shown in detail where a reaction takes
place.

Figure 10.6: Different types of structural formula

The examples of the open chain hydrocarbons shown in figure 10.6


above have a straight chain because the carbon atoms are connected
continuously in one line. When there are branches or substituents on
atoms in the chain, the hydrocarbon is said to be a branched chain as
166
in the structures shown in below. These are examples of catenation.
H3C CH3
H2 CH
C CH CH3
H3C C C
H2 H2

Nomenclature of Alkanes
To be able to name so many compounds, it became necessary to
devise a systematic method of nomenclature. The International Union
of Pure and Applied Chemistry developed what are known as the
IUPAC rules of nomenclature, which are accepted worldwide. These
rules as applied to simple hydrocarbons are described below.
(a) Identify the longest carbon chain (treat it as the parent chain) in
the molecule.

(b) Identify the branches, also called substituents. They will be


named after the parent hydrocarbon by replacing suffix ane
with yl. For example, CH3 and C2H5 will be called methyl and
ethyl, respectively. If the substituent is a halogen (the compound
will not be a hydrocarbon), it will be named like chloro,
bromo, fluoro, etc.

(c) Number the parent chain from the side that gives the minimum
number to the substituent see example below.
6 4 2
5 3 1

The name and position of the substituent are shown in the name
of the compound by an Arabic numeral before the name of the
parent chain. Thus, the structure above is named as 3-
methylhexane.

167
Some other common branches you will encounter during your
studies are shown below.

(d) If there is more than one substituent, number the chain from the
side giving the minimum number to the first branch and arrange
them in an alphabetical order.
8 7 6 4 3 2 1 3 5 6 7
5 1 2 4 8

5-ethyl-3-methyloctane 4-ethyl-6-methyloctane
correct wrong

(e) If a particular substituent appears twice or three times, prefixes


such as di- for two, tri- for three, tetra- for four and so on are
used. However, these prefixes are not taken into consideration in
arranging the branches alphabetically.

Classification of carbon and hydrogen atoms in Organic


Compounds
In naming alkyl groups, we use prefixes pr for primary, sec for
secondary, tert for tertiary, etc. In a C4 alkyl group, four
possibilities arise depending on the number of other carbon atoms
attached to the branching atom (Figure 10.7). The hydrogen atom
attached to a 1o, 2o and 3o carbon is referred to as 1o, 2o and 3o
hydrogen, respectively. There is, thus, no quaternary or 4o hydrogen.
168
H R R R
R C H R C H R C H R C R
H H R R
Primary carbon (1o) Secondary carbon (2o) Tertiary carbon (3o) Quaternary carbon (4o)
bonded to one bonded to two bonded to three bonded to four
other carbon other carbons other carbons other carbons

Figure 10.7: Classification of carbon atoms


Isomerism in Alkanes
If you try to draw the structural formula of butane C4H10, you may do
it in two different ways - by drawing a straight chain or by drawing a
branched chain. In fact, there are many organic compounds, which
have the same molecular formula but different structural formula.
Such compounds are called isomers and the property is isomerism.
Isomerism is broadly divided into two types: structural isomerism and
stereoisomerism. Structural isomers differ in the bonding
arrangement of the atoms, whereas stereoisomers have the same
bonding arrangements of the atoms but differ in spatial orientation of
the groups. Isomers are in fact different compounds and so differ in
physical and chemical properties.
Structural isomerism is further divided into skeletal isomerism (see
figure 10.8), positional isomerism and functional group isomerism,
depending on whether the difference in the structure is in the carbon
skeleton, the position of the functional group or the type of the
functional group.
Stereoisomerism is of two types - geometrical isomerism and optical
isomerism. Structural isomerism is observed in open-chain alkanes
due to different carbon skeletons.
You will observe both structural and geometrical isomers in alkenes
and in cycloalkyl compounds.

butane isobutane pentane isopentane 2,2-dimethylpropane

skeletal isomers of butane skeletal isomers of pentane

Figure 10.8: Examples of skeletal isomerism


169
Occurrence
Petroleum and natural gases are the two main sources of
hydrocarbons. Petroleum is also called rock oil (Latin word, petra =
rock; oleum = oil). It is a complex mixture of hydrocarbons,
principally alkanes and cycloalkanes, and occurs deep down the earth
as black viscous crude oil. Crude oil usually occurs together with
volatile hydrocarbons called natural gases. These oils and gases are
formed over periods of time as a result of decay of buried living
beings. Petroleum and natural gases are brought to the surface by
drilling and pumping, and refined. The various fractions of
hydrocarbons (with different boiling point range) are separated by
fractional distillation.

Fractional distillation
It is a distillation process for separating the components of crude
oils. Distillation of a liquid depends on the boiling point. Figure 10.9
shows a fractional distillation apparatus. Crude oil is heated to vapour
state and introduced in the column where a gradation in temperature.
The various fractions condense and are collected. The gaseous
components escape on the top where they are collected.

Figure 10.9: Apparatus for fractional distillation

170
The boiling range and uses of the main fractions are given below.
C1C4: -164 to 20oC, natural gases; C3 and C4 are major
components of cooking gas
C5C11: 30 to 200oC, gasoline: C8 principle component of petrol
C11C18: 200-370oC, paraffin, kerosene and diesel
C17C22: > 370oC, lubricating oils
C23C34: solids, mostly wax such as Vaseline
C35: solids such as asphalt

Gasoline is the most valuable fraction as fuel and for petrochemical


industries that make fibres, plastics and other useful material.
Hydrocarbons, especially those with highly branched carbon chains,
burn smoothly in an automobile engine and drive the piston forward
smoothly. Straight-chain hydrocarbons tend to burn violently in the
engine and produce knocking as a result of uneven movement of the
piston. The best burning petrol in an engine is 2,2,4-trimethylpentane
(isooctane). Based on this, a system of rating the petrol is in use that
is called Octane Rating. 2,2,4-Trimethylpentane (isooctane) is
arbitrarily given a rating of 100. Heptane is rated 0 as its combustion
in an engine is not smooth. Thus a fuel rated 90 has a similar quality
as fuel as a mixture containing 90% isooctane and 10% heptane. The
octane rating of a fuel may be improved by adding tetraethyl lead,
methanol or ethanol. Lead is being phased out because of its toxicity.
Fuel stations in Botswana supply fuels with octane number 91 and 93.

Cracking
Cracking is the process by which high molecular weight hydrocarbons
(those containing larger number of C atoms) are converted to low
molecular weight hydrocarbons, which are more useful for syntheses
of other compounds. Cracking is carried out either thermally
(heating) or catalytically. The main catalysts for this purpose are
silica and alumina. Cracking results into formation of large amounts
of the hydrocarbons with lower molecular weights, for example:

171
C8H18 + C8H16
C10H22 C5H12 + C5H10
C4H10 + C4H8

Physical Properties
Solubility
Hydrocarbons are insoluble in water but soluble in solvents like ether
and benzene. The general rule for the solubility states that like
dissolves like. Water molecules are polar, whereas hydrocarbons are
non-polar. A large difference in the electronegativity of oxygen and
hydrogen makes the O-H bond in a water molecule polar. In
hydrocarbons, C-C and C-H bonds are nearly non-polar covalent
bonds. Ether and benzene molecules are also non-polar, so
hydrocarbons dissolve in them.

Density
Hydrocarbons are less dense than water and will float on its surface.
This is the reason why oil spilled on water remains on the surface.

Boiling points
Boiling point increases in a homologous series with the number of C
atoms for most classes of organic compounds. For a particular
molecular weight, however, it decreases with increased branching of
the chain. Branching decreases mutual forces of attraction among
molecules, since branches reduce the area of contact between
molecules. This reduces the attractive forces between molecules
leading to easier separation, hence lower boiling points compared to
straight-chain molecules. (See figure 10.10 below.)

Figure 10.10: Boiling points of pentane isomers (in oC)

172
Alkenes have lower boiling points than their saturated counterparts.
For example, pentene has a lower boiling point than pentane. This is
due to increased repulsion in alkenes because of the presence of the
double bond.

Chemical Properties
Chemical properties of organic compounds are described as their
reactions with particular reagents. In fact, all the reactions of organic
compounds can be divided into a few general types such as
substitutions, additions, eliminations, oxidations, reductions, and
molecular rearrangements.
Alkanes have a very limited reactivity because they are saturated
molecules. The two main reactions of alkanes are with oxygen
(combustion an oxidation reaction) and with halogens
(halogenation), which are substitution reactions.

Combustion
It is the most important reaction of alkanes as it provides both light
and heat energy. Heat energy is convertible to several other forms
such as moving of limbs, machines etc. In the presence of excess
oxygen, alkanes burn with a clear flame. The accounting of heat
energy produced will be studied in thermochemistry.
Examples: CH4 + 2O2 CO2 + 2H2O + heat
2C4H10 + 13O2 8CO2 + 10H2O + heat
We can see from the above reaction equation that the hydrogen atoms
in alkanes are substituted by oxygen atoms. Thus, combustion is an
oxidation reaction. If oxygen present is insufficient, carbon monoxide
can also be produced as a result of combustion.
Example: 2CH4 + 3O2 2CO + 4H2O

Halogenation
Halogenation reactions are examples of substitution reactions. One or
more hydrogen atoms of the hydrocarbon are replaced by halogen (Cl
or Br) atoms. The reaction is usually carried out by exposing a
mixture of alkanes and halogen to light. Remember that the organic

173
product(s) of a halogenation reaction is no more a hydrocarbon. It
contains halogen atoms besides carbon and hydrogen atoms. These
products are called haloalkanes. Example:

A haloalkane, 1-bromo-1-chloroethane, is used as general anaesthetic.


Halogen atom(s) will be treated as a branch on the main chain in
naming such compounds. Inorganic products are not treated as main
products of an organic reaction because the primary purpose of
studying chemical properties of organic compounds is to transform
them into other organic compounds.
The reaction is called chlorination if the halogen is chlorine and
bromination if the halogen is bromine. Monohalogenation
(replacement of only one hydrogen atom) in such reactions is rarely
possible. It is difficult to stop the reaction at the monohalogenation
step and a mixture of products is obtained. This is why the
halogenation of alkanes is not a very useful reaction for the
preparation of monohaloalkanes. If excess of halogen is present, the
reaction goes on until all hydrogen atoms are substituted. For
example, ethane reacts with excess of chlorine in the presence of light
to form hexachloroethane:
C2H6 + excess Cl2 C2Cl6 + 6HCl

Mechanism of halogenation of alkanes


A reaction equation tells us about the reactants taking part in the
chemical reaction and products formed in it. A chemical reaction
involves breaking of bonds in the reactants and formation of new
bonds. As a result of bond breaking, unstable species are formed in
the reaction. These unstable species are called reactive intermediates
or simply intermediates. These intermediates, because they are
unstable, quickly combine further in different fashions giving rise to
products. A step by step description of all the bond breaking and
bond forming processes in a chemical reaction is called reaction
mechanism. The reaction mechanism of a reaction is determined by
several studies.

Halogenation of alkanes is a three-step mechanism - initiation, chain


propagation and chain termination. The first two steps involve
174
formation of intermediates with a single electron in the valence shell.
These intermediates are called free-radicals or simply radicals. The
radicals react either with neutral molecules or with another radical.
This is why the halogenation reactions of alkanes are said to occur by
radical mechanism.

Initiation:
It involves homolytic cleavage of the Cl-Cl bond by light energy.
Homolytic cleavage is the cleavage of a covalent bond in which an
electron pair is shared equally between the two atoms joined by that
covalent bond. Thus, each of the chlorine atoms gets away with one
electron to form two chlorine radicals that are shown in the equation
below with the symbol of a chlorine atom and a dot over it, which
represents an unshared electron.
light
Cl Cl Cl + Cl

Propagation:
It involves the reaction of a radical with a neutral molecule forming a
new radical and a neutral product (the main reaction product).
H
H H + Cl CH3 + HCl
H

CH3 + Cl Cl Cl CH3 + Cl
Termination:
It involves a reaction between two similar or different radicals
forming molecules. Now you can see that the main product of the
reaction, chloromethane, is formed during propagation and
termination steps. The formation of a small amount of ethane in the
halogenation of alkanes supports the free radical mechanism.

CH3 + CH3 H3C CH3

CH3 + Cl Cl CH3

Cl + Cl Cl2

175
B. Cycloalkanes
The cyclic or ring analogs of the open chain alkanes shown above are
called cycloalkanes. The prefix cyclo is included in their names to
indicate the presence of the ring. Since the two ends of chain will be
joined in such compounds, they will have two hydrogen atoms less
than their open chain counterpart.
Thus, the general molecular formula of cycloalkanes is CnH2n.
Propane has the molecular formula C3H8 while cyclopropane has
molecular formula C3H6. The study of cycloalkanes is of immense
importance in organic chemistry because such rings are often present
in compounds found in nature. Examples are shown below.
H2 H2
C C
H2C CH2
- 2H
H H H2C CH2

Cyclobutane cyclopentane cyclohexane

If a cycle has a branch, always start numbering the cycle from the
carbon that bears the branch. If there is more than one branch, then
give alphabetical priority to the branches and number from the side
that gives the minimum number to other branches. Some examples
are:

methylcyclopentane 1-ethyl-2-methylcyclohexane
NOT 2-ethyl-1-methylcyclopentane
or
1-ethyl-6-methylcyclohexane

The physical properties of cycloalkanes will show a similar trend as


observed in alkanes. For example, the boiling point of cycloalkanes
increases with increasing ring size. Thus, cyclopropane has a lower
boiling point than cyclobutane, which has a lower boiling point than

176
cyclopentane. Chemical properties are also similar to those of
alkanes. Thus, cycloalkanes undergo combustion and halogenation
reactions.

Cycloalkanes show both structural and geometrical isomerism.


Structural isomerism of skeletal or positional types is possible due to
completely different skeletons with the same molecular formula or
due to different positions of branches. You must keep in mind that the
compounds belonging to different classes can also be isomers to each
other. To be of the same class is not a necessary criterion for
exhibiting isomerism. Some examples of isomerism in cycloalkanes
are shown below.
CH3 CH3 CH3

CH CH3 CH CH
H2C CH H2C CH2 H2C CH2

H2C CH2 H2C CH H2C CH2


C C CH3 CH
H2 H2
CH3
1,2-dimethylcyclohexane 1,3-dimethylcyclohexane 1,4-dimethylcyclohexane

structural isomers (positional)

As stated earlier, cycloalkanes differ from alkanes structurally in


terms of chain and cycle or ring. This difference gives a special
structural feature to cycloalkanes. The cycloalkanes are rigid
molecules in terms of rotation of single bonds. The carbon-carbon
single bonds in alkanes can rotate freely but not in cycloalkanes. If
there are two substituents on the ring, they may be on the same side
relative to the plane of the ring, or they may be on opposite sides.
This leads to geometrical isomerism. If the two substituents are on the
same side, the isomer is cis, if they are on opposite sides it is a
trans isomer:

177
cis-1,3-dimethylcyclobutane trans-1,2-dimethylcyclobutane

geometrical isomers

C. Alkenes
Alkenes are unsaturated hydrocarbons also known as olefins. They
contain at least one carbon-carbon double bond. The carbon atoms in
C=C are sp2 hybridized with bond angles of 120o. One of the double
bonds is a sigma bond and the other is a pi bond. The general
molecular formula of this class of compounds is CnH2n. The simplest
compound with n = 2 has molecular formula C2H4 and is called
ethene.
Table 10.2: Naming of alkenes

n Compounds Molecular n Compounds Molecular


formula formula
2 Ethene C2H4 7 Heptene C7H14
3 Propene C3H6 8 Octene C8H16
4 Butene C4H8 9 Nonene C9H18
5 Pentene C5H10 10 Decene C10H20
6 Hexene C6H12

The names of all the alkenes end with ene. Prefixes prop, but,
pent and so on are used for three, four and five carbons respectively,
as used for alkanes (Table 10.2). Analogous to cycloalkanes, cyclic
analogs of alkenes are called cycloalkenes. Their general molecular
formula has two hydrogen atoms less than the open-chain counterpart
(CnH2n-2).

178
Nomenclature
In alkenes containing four or more carbon atoms, the position of the
carbon-carbon double has to be indicated. The carbon chain is
numbered from the side that gives the minimum number to the first
carbon attached to the carbon-carbon double bond. Example:

In branched chain alkenes, the chain containing the double bond is


considered as main chain (even if it is not the longest one). In the
example shown below, the longest possible chain has nine carbon
atoms. But the chain with eight carbon atoms is taken as the main
chain because it contains the carbon-carbon double bond. The chain is
numbered from the side that gives the minimum number to the double
bond. If a chain has both a double bond and an alkyl substituent
(branch), give the minimum possible number to the double bond.

In naming a branched cycloalkene, the numbering of carbon always


starts from the carbon attached with the double bond and not from the
carbon bearing the branch as in cycloalkanes. The second carbon will
always be the other carbon atom attached to the double bond. The
numbering is started from that C of the two with the double bond
which gives the lower number to the substituent as shown in the
examples below.

179
Classification of carbon and hydrogen atoms
Sometimes common names are used to denote some particular types
of carbon and hydrogen atoms present in olefinic systems. For
example, the carbon and hydrogen atoms of a CH2 attached to C=C
are called allylic carbon and allylic hydrogen. However, if the C=C to
which the CH2 is attached is present in benzene ring, then the carbon
and hydrogen are referred to as benzylic carbon and hydrogen (see
figure below). A H2C=CH system is referred as vinyl group in alkene
chemistry.

Structural and geometrical isomers


As we saw above, a double bond in alkenes, having four or more
carbons can have different positions on the chain. As a result, such
alkenes can exist as structural isomers (positional isomers):

Alkenes also exhibit geometrical isomerism. It arises due to rigidity


of the carbon-carbon double bond. Alkenes having different atoms or
groups attached to the doubly bonded carbons will have two isomers;
one in which the identical groups are on same side of the double bond
called cis, and the other in which identical groups will be on the
opposite side of the double bond called trans. If either of the carbon
atoms attached to the double bond has identical atoms or groups the
compound cannot exhibit geometrical isomerism. You will study
other types of geometrical isomers such as conformers and optical
isomers at higher levels. Some examples of geometrical isomers are
shown below.

180
Reactions of Alkenes
Alkenes undergo many more reactions than alkanes. This is because
of the presence of pi-electrons in the double bond, which are loosely
bound.

Combustion
Alkenes, like alkanes, are easily oxidized to carbon dioxide and water
when burnt in excess oxygen. The flames are sooty, unlike alkanes,
because of a higher carbon to hydrogen ratio.

Addition reactions of Alkenes


The most common reactions of alkenes are addition reactions. In such
reactions, the atoms or groups of the reagent are added to carbon
atoms attached to the double bond yielding products in which these
carbon atoms will have a single bond. The reagents might be
symmetrical such as hydrogen (H2), chlorine (Cl2) or bromine (Br2), or
unsymmetrical such as hydrogen halides (H-X) and water (H-OH).

Addition of hydrogen (reduction)


Addition of hydrogen to alkenes, also known as hydrogenation, is an
example of reduction reaction of alkenes. The reaction occurs in the
presence of a catalyst. Some commonly used catalysts are metals like
Pt, Pd, and Ni. This reaction is of great value and is used in the
production of margarine.

181
Addition of halogens
These reactions take place at room temperature without any catalyst or
necessity of light. Note again that the product of halogen addition is
no more a hydrocarbon. Furthermore, two halogen atoms are added to
alkene molecules, hence, the products are dihaloalkanes. These
reactions do not need any control because the addition will stop if
there is no more unsaturation (double bond) in the alkene molecule.
Thus, only two halogen atoms can be added to one alkene molecule.
Bromine-water solution is used to test the presence for unsaturation in
alkenes and cycloalkenes. Such hydrocarbons decolorize the brown
bromine solution.

Addition of hydrogen halides


Common examples are reactions of H-Cl and H-Br with alkenes. The
reactions of alkenes with hydrogen halides lead to the formation of
monohaloalkanes. For example, the reaction of ethene with hydrogen
chloride forms chloroethane. Similarly, the reaction of 2-butene with
hydrogen bromide gives 2-bromobutane. The products of
halogenation of alkanes and of hydrogen halide additions to alkenes
belong to the same group: monohaloalkanes. However, they are
formed by substitution reaction in the case of alkanes, but in the case
of alkenes they are formed by addition reaction. The substitution
reactions are difficult to control at the monohaloalkane stage, but

182
addition of hydrogen halides to alkenes will always give only one
monohaloalkane as long as the carbon atoms joined by the double
bond are symmetrically substituted. Examples:

H H H H
+ H-Cl H Cl
H H H H

ethene chloroethane

H CH3
H CH3
+ H-Br H Br
H3C H
H3C H
trans-2-butene 2-bromobutane

In the above examples, the alkenes are symmetrical; the atom or group
present on both carbon atoms joined by the double bond is the same.
Each carbon has one hydrogen atom and one methyl group. If the
alkenes are unsymmetrical, two products are possible from addition
reactions.
For example, 1-butene is an unsymmetrical molecule because one of
the carbon atoms attached to the double bond has two hydrogen atoms
while the other carbon attached to the double bond has one hydrogen
atom and one ethyl group. The main addition product in reactions of
unsymmetrical alkenes with unsymmetrical reagents is governed by
Markovnikovs rule.

According to this rule, the hydrogen atom (or the proton) goes to that
carbon of the double bond which has the larger number of hydrogen
atoms. That means that the remaining part of the reagent attaches to
the carbon with the lower number of hydrogen atoms. Thus, in the
example below, the reaction of 1-butene with hydrogen bromide, the
proton goes to the 1st carbon which has two hydrogen atoms, and
bromine (bromide ion) goes to the 2nd carbon, which has only one
hydrogen atom.

183
H Br
H 1 2 + H-Br Br +
H 1-butene 1-bromobutane 2-bromobutane
(main product)

In the reaction shown below, the carbons attached by the double bond
have one hydrogen atom each, but the alkyl groups attached to them
are not same. One carbon has an ethyl group and the other carbon
atom has a propyl group. In this case, the halogen addition will give a
mixture of products.
H H Cl H H
+
+ H-Cl
H H H Cl
H
3-chloroheptane 4-chloroheptane

Mechanism of addition of hydrogen halides


The hydrogen halides are ionized to hydrogen and halide ions. The
pi-electrons, because they are loosely bound as mentioned earlier in
discussion of hybridization, capture the proton to form a carbon-
hydrogen single bond leading to an intermediate. Since the other
carbon attached to the double bond has lost an electron that it was
sharing, it now has a positive charge in the intermediate. Such
intermediates which bear a positively charged carbon are called
carbocations. The negative halide ion will then react with this
intermediate at the carbon atom with the positive charge to form a
carbon-halogen bond. The steps involved in the addition of hydrogen
bromide to ethene are shown below.
Such a reaction mechanism that involves ionic intermediates is called
ionic mechanism. The curved arrows shown in the scheme represent
movement of a pi-electron pair from the carbon-carbon double bond
and electron pairs from the bromide ion. Such arrows are used in
describing other ionic mechanisms as well. They are always shown to
originate from the electron source and end up at the electron-deficient
centre. Keep in mind that for the addition of unsymmetrical reagents
to unsymmetrical alkenes, the proton goes to the carbon that has the

184
higher number of hydrogen atoms. The explanation for this is beyond
the scope of this chapter and will be explained at higher levels.

Addition of water (hydration)


The addition of water to alkenes requires an acid as a catalyst. The
catalyst is necessary because water molecules do not ionize with the
same ease as do hydrogen halides. First, the alkene and the acid are
mixed, and then the mixture is poured slowly into water. Commonly
used acids are sulphuric acid and phosphoric acid.
Markovnikovs rule applies in the case of unsymmetrical alkenes.
The product in this case is not a hydrocarbon since it contains oxygen
in the form of OH (hydroxyl group). The product belongs to the class
of alcohols. The reaction of 1-butene with water is shown:

The catalytic reaction of ethene with water produces ethanol (ethyl


alcohol) see the reaction below. It is used as an antiseptic in
hospitals, and in the production of alcoholic beverages. Thus, the
reaction of alkenes with water is a method for the preparation of
alcohols. Alcohols are further used for synthesis of aldehydes,
ketones and acids by oxidation reactions.

185
D. Alkynes
Alkynes are also unsaturated hydrocarbons. They contain at least one
carbon-carbon triple bond. Remember that two carbon-carbon bonds
in the triple bond are pi bonds and the third one is the sigma bond.
Alkynes greatly resemble alkenes in their chemical properties. The
general molecular formula of this class of compounds is CnH2n-2. The
simplest compound with n = 2 has molecular formula C2H2 and is
called ethyne. Its common name is acetylene.
The names of all alkynes end with yne. The rules for naming
alkynes are the same as those for alkenes (Table 10.3). The smallest
cycloalkyne known is cyclooctyne (its structure is shown below). Due
to the linear geometry of the carbon atoms involved in triple bond
formation, formation of smaller cycloalkyne rings is not possible.

Table 10.3: Naming of alkynes.


n Compounds Molecular n Compounds Molecular
formula formula
2 Ethyne C2H2 7 Heptyne C7H12
3 Propyne C3H4 8 Octyne C8H14
4 Butyne C4H6 9 Nonyne C9H16
5 Pentyne C5H8 10 Decyne C10H18
6 Hexyne C6H10

Reactions of Alkynes
Alkynes do burn like other hydrocarbons. Their flames are smoky due
to an increased proportion of carbon atoms to hydrogen atoms in
186
comparison with alkanes and alkenes. Alkynes are highly reactive
compounds because of the presence of pi-electrons and undergo
addition reactions like alkenes. They incorporate two equivalents of
reagents, instead of one equivalent in the case of alkenes. Some
examples are given below.
Addition of hydrogen
Alkynes are converted into alkanes using two mole equivalents of
hydrogen in the presence of platinum, palladium or nickel as a
catalyst. For example:

It is possible to obtain an alkene from hydrogenation of an alkyne


using a specific catalyst called Lindlars catalyst. The reaction of an
alkyne in the presence of Lindlars catalyst gives the cis isomer of the
product alkene (see example below). If one wishes to synthesize a
trans-alkene from an alkyne by controlled hydrogenation, another set
of reagents needs to be used.

Addition of halogens
To each molecule of an alkyne, two halogen molecules or four
halogen atoms can be added to give tetrahaloalkanes:

187
Addition of hydrogen halides
This represents an addition of an unsymmetrical reagent. Since one
alkyne molecule can accommodate two hydrogen halide molecules,
the final product of the reaction will contain two halogen atoms and
two more hydrogen atoms in comparison to the reactant alkyne. The
final product is thus a dihaloalkane. Addition of the first molecule of
hydrogen halide to the carbon-carbon triple bond will form a
monohaloalkene. In this type of alkene, the carbon-carbon double
bond is unsymmetrically substituted. Hence, it would react with
another molecule of hydrogen halide in accordance with
Markovnikovs rule forming the dihaloalkane. However, if the alkyne
itself is unsymmetrical, it would react with the first molecule of
hydrogen halide also according to Markovnikovs rule. Example:

Addition of water
Addition of water to alkynes gives a ketone. This reaction requires a
mercury(II) salt such as Hg(COOCH3)2 as a catalyst besides an acid.
The immediate product of hydration is called an enol (because it
contains ene and ol). This product can exist in two isomeric forms.
The more stable form is a ketone (contains C=O group).

188
The two isomers are called tautomers. It is a special type of
isomerism in which two forms of the product are in dynamic
equilibrium, with the more stable product dominating in the
equilibrium mixture.

E. Arenes
Arenes are highly unsaturated hydrocarbons with a cyclic structure.
They are also called aromatic hydrocarbons due to their typical aroma
(flavour). You will see later that even though these compounds are
highly unsaturated, they do not show the typical chemical reactions of
alkenes, such as addition reactions, under normal conditions. This is
the reason why we study them as a separate class.
The simplest aromatic hydrocarbon, benzene, has a six-membered
carbon ring. Each carbon has one hydrogen giving it the molecular
formula C6H6. Thus, it has six hydrogen atoms less than its saturated
analog cyclohexane, indicating the presence of three double bonds in
it, see the following figure.

Structure of Benzene
As mentioned earlier, benzene does not undergo the typical addition
reactions of alkenes. All carbon atoms in benzene are sp2 hybridized.
Thus, each carbon atom has a non-hybridized p orbital containing one
electron. These orbitals overlap with each other, forming a ring of six
pi electrons. The pi electron pairs/bonds are not associated with a
particular carbon atom. These electrons are said to be delocalized. As
a result, there is neither any true carbon-carbon double bond nor any
true carbon-carbon single bond. It means that the structure of
benzene shown above is not a true representation. Benzene is
represented by a hexagon and a circle inside it which represents the
delocalized pi bonds (or a ring of pi electrons).
189
Delocalization of pi electrons provides additional strength to the ring.
It is thus very difficult to break an aromatic structure, and benzene
does not undergo an addition reaction. However, under the influence
of some catalysts it can undergo substitution reactions, in which the
aromatic ring remains intact.
Some common products of substitution of benzene are nitrobenzene,
alkyl benzenes, halobenzenes, benzenesulfonic acids and ketones.
These benzene derivatives can be further transformed to a number of
other compounds containing a benzene ring. For example, the methyl
group in methylbenzene is oxidized to get benzoic acid. The nitro
group (NO2) in nitrobenzene is reduced to an amino group (NH2),
forming benzenamine, commonly known as aniline. These derivatives
are summarised in Figure 10.11.

Fig 10.11: Structures of benzene and common benzene derivatives

10.5 Compounds containing Oxygen and Nitrogen Atoms besides


Carbon and Hydrogen: Functional Groups

Other compounds are grouped depending on the functional group


they contain. The functional group might be an atom (other than C
and H), or a group of atoms containing at least one heteroatom (atom
other than C and H). Since C=C is responsible for the specific
properties of alkenes, it is also considered as a functional group.
Functional groups are responsible for the physical and chemical
properties of organic compounds. Some important functional groups,
and naming of compounds containing these functional groups
according to the IUPAC rules, are shown in Table 10.4.

190
Table 10.4: Functional groups and nomenclature

Functional Class Nomenclature and Classification


group
OH Alcohols, Remove e from the end of alkyl group
(hydroxyl) Phenols and add ol.
CH3OH (methanol), CH3CH2OH (ethanol).
RCH2OH (primary or 1o alcohols)
R2CHOH (secondary or 2o alcohols)
R3COH (tertiary or 3o alcohols)
It is called phenol if the oxygen atom of
OH is attached to a benzene ring.
COC Ethers Name according to groups attached to
oxygen atom:
CH3OCH3 (dimethyl ether),
CH3OCH2CH3 (ethyl methyl ether).
RCH=O Aldehydes Remove e from the end of alkyl group
and add al.
HCH=O (methanal), CH3CH=O (ethanal),
CH3CH2CH=O (propanal).
R2C=O Ketones Name the carbon chain giving the minimum
number to ketone group. Remove e from
the alkyl group, locate the position of C=O,
and add one.

191
Carboxylic Remove e from the end of alkyl group
acids and add oic acid.

Esters Esters are formed by the reaction of a


carboxylic acid with an alcohol. Start the
name of the compound from the group
attached to oxygen. Number the chain
starting from C=O to the side opposite to
which the oxygen atom is attached. Delete
e from the end of the alkyl group and add
oate. If the chain has a multiple bond or
any other group on it, include its name and
position.

192
NH2 Amines Amines are organic derivatives of ammonia
formed by the replacement of one or more
hydrogen atoms. Remove e from the end
of the alkyl group and add amine.

Amines are classified as primary (1o),


secondary (2o) and tertiary (3o) amines.
The examples shown above are 1o amines.
The nitrogen in 2o amines will have one
hydrogen atom, but it will have no
hydrogen in a 3o amine.

Amides Amides are derivatives formed by the


reaction of carboxylic acids with amines.
To name, remove e from the end of the
alkyl group and add amide.

193
10.6 Exercises
10.1 Write the detailed structural formula and molecular formula of the
following compounds:
a. b. c. d.

10.2 Identify the hybridization of each carbon atom in the following


compound. What are the bond angles and what is the geometry
around each carbon?

10.3 Count the total number of sigma and pi bonds in each of the
molecules shown in Exercise 10.2.

10.4 Draw the structural formula of the following compounds:


(a) 3-methylhexane; (b) trans-4-methyl-2-pentene;
(c) 4-ethyl-2-methylnonane; (d) cyclopentanecarboxaldehyde;
(e) 5-isopropyl-1-octyne; (f) 4-tert-butyl-1-heptene.
(g) cis-4-methyl-2-hexene (h) 4-methylcyclohexene

10.5 Give the systematic names of the following compounds. If


geometrical isomers are possible, write the name of each one.
(a) CH2=CHCH2CH3; (b) CH3CH2CH=CH-CH2CH3;
(c) (CH3)2C=CHCH2CH2CH3.
(d) CH3C CCH(CH3)CH2CH2CH3
(e) CH3CH2CH2C CCH(CH3)CH3

10.6 Why do branched-chain alkanes have lower boiling points than


straight-chain alkanes with the same number of carbon atoms?

10.7 Write all possible open-chain structural isomers of the compound


with molecular formula C5H12.

10.8 Write the structural formulas and names of two cycloalkanes with
molecular formula C4H8.

194
10.9 Identify each of the following pairs as structural isomers, geometrical
isomers, or not isomers:
(a) pentane and cyclopentane (b) 1-butene and cyclobutane
(c) diethyl ether and 2-butanol (d) 1-butene and 2-butene
H H H3C H
H H H3C H and H2
e. and f. H3C C CH3
C CH3 H C C
H3C CH3 H CH3 H2 H2 H2
10.10 Write all the monochlorination product(s) possible from the
reaction of the compounds shown below with chlorine in the
presence of light. Classify the reaction as substitution or addition
reaction.

a. b. c.

10.11 Why do alkenes react more easily with chlorine and bromine in
comparison to alkanes?

10.12 Write the main products of the following chemical reactions.


a. C4H10 + O2

b. + HBr

c. + excess Cl2

Ni
d. + H2

e. Lindlar'catalyst
+ H2

H2SO4 (catalyst)
f. + H2O

H2SO4 (catalyst)
g. + H2O
Hg2+
h. HCl
+ HCl

195
10.13 Reactions of some alkenes were performed to get the products
shown in the following incomplete chemical equations. Complete
these equations by providing the structures of the appropriate
alkene and other reagents as shown in the equations of question 12.

a. CH3CHBrCHBrCH3

b. Br

CHBrCH3
c.

H3C
d. OH
H3C
CH3

10.14 Identify all the functional groups in the following compounds.


OH
O
H
N O
NH2
OHC OC2H5 OH O

H OH O
N OH
COOH
O H3CO

10.15 Write the structural formula of:


(a) methanal; (b) 2-pentanone; (c) propanone;
(d) benzoic acid; (e) phenol; (f) methylbenzene.

10.16 Write the structural formula of:


(a) 2-methylhexanal; (b) 2-methylbutanoic acid;
(c) dimethylamine; (d) ethyl 3-methylhexanoate;
(e) 2-isopropylheptanamine; (f) 3-methyl-2-butanone;
(g) isopropylhex-2-enoate (or isopropyl 2-hexenoate);
(h) 3-methylcyclopentanol.

196
CHAPTER 11
Thermochemistry
11.1 Introduction: the Basics of Energy, Heat and Work

Energy
All chemical changes are accompanied by the absorption or release of
heat. The branch of chemistry dealing with release or absorption of
heat during chemical processes is called Thermochemistry. The
intimate connection between matter and energy has been a source of
wonder and speculation from the most primitive times; it is no
accident that fire was considered one of the four basic elements (along
with earth, air, and water) as early as the fifth century BC. In this
chapter, we will review some of the fundamental concepts of energy
and heat and the relation between them.

Energy is one of the most fundamental and universal concepts of


physical science, but one that is remarkably difficult to define in way
that is meaningful to most people. This perhaps reflects the fact that
energy is not a thing that exists by itself, but is rather an attribute of
matter (and also of electromagnetic radiation) that can show itself in
different ways. It can be observed and measured only indirectly
through its effects on matter that acquires, loses, or possesses it.

Energy units

Energy is measured in terms of its ability to perform work or to


transfer heat. Work can take various forms: mechanical, electrical,
gravitational, etc. Mechanical work is done when a force f displaces
an object by a distance d: w = f d. The basic unit of energy is the
joule. One joule is the amount of work done when a force of 1 newton
acts over a distance of 1 m, thus 1 J = 1 N m. The newton is the
amount of force required to accelerate a 1 kg mass by 1 m/s2, so the
basic dimensions of the joule are kg m2 s2. Electrical work is done
when a body having a certain charge moves through a potential
difference.

Like work, heat is also measured in energy units. Heat (or thermal

197
energy) can be transferred from one body to another. We often refer
to this as a "flow" of heat, recalling the 18th century notion that heat
was an actual substance called caloric that could flow like a liquid.
We know that heat can only flow from a system at a higher
temperature to one at a lower temperature. This special characteristic
is often used to distinguish heat from other modes of transferring
energy across the boundaries of a system.

Heat and work are best thought of as processes by which energy is


exchanged, rather than as energy itself. That is, heat exists only
when it is flowing, work exists only when it is being done. When
two bodies are placed in thermal contact and energy flows from the
warmer body to the cooler one, we call the process heat. A transfer
of energy to or from a system by any means other than heat is called
work.

11.2 Endothermic and Exothermic Processes

As a result of chemical reactions heat is either adsorbed or


released. Endothermic reactions are those that are accompanied
by absorption of thermal energy, or heat. Exothermic reactions
are accompanied by release of heat. An example of an
exothermic reaction is the burning of sucrose in oxygen:
C12H22O11 + 12 O2(g) 12 CO2(g) + 11 H2O(l).
This reaction is accompanied by evolution of 5606 kJ heat per mole of
sucrose burnt.
An example of an endothermic process is the boiling of one mole
water, which takes place with the absorption of 40.7 kJ heat:
H2O(l) H2O(g)

Heat capacity
As a body loses or gains heat, its temperature changes in direct
proportion to the amount of thermal energy q transferred:
q = C T
The proportionality constant C is known as the heat capacity:
C = q / T
198
If T is expressed in kelvins and q in joules, the units of C are J K1.
In other words, heat capacity tells us how many joules of energy it
takes to change the temperature of a body by 1 K or 1C, since a
difference of one degree is the same on the two scales. The greater
the value of C, the smaller will be the effect of a given energy change
on the temperature.

It should be clear that C is an extensive property - that is, it depends on


the quantity of matter. Everyone knows that a much larger amount of
heat is required to bring about a 10C change in the temperature of 1 L
of water compared to 10 mL of water. For this reason, it is customary
to express C in terms of unit quantity, such as per gram, in which case
it becomes the specific heat capacity, commonly referred to as the
"specific heat" and has the units J K1g1.

Example 11.1:
How many joules of heat must flow into 150 mL of water at 0C to raise
its temperature to 25C?
Solution:
The mass of the water is (150 mL) (1.00 g mL1) = 150 g. The specific
heat of water is 4.18 J K1 g1.
From the definition of specific heat, the quantity of energy q is
(150 g) (25.0 K) (4.18 J K1 g1) = 15675 J = 1.57 x 104 J.

Example 11.2:
A piece of nickel weighing 2.40 g is heated to 200.0C, and is then
dropped into 10.0 mL of water at 15.0C. The temperature of the metal
falls and that of the water rises until thermal equilibrium is attained and
both are at 18.0C. What is the specific heat of the metal?
Solution:
The mass of the water is (10 mL) (1.00 g mL1) = 10 g.
The specific heat of water is 4.18 J K1 g1, and its temperature increased
by 3.0C (= 3 K), indicating that it absorbed:
(10 g) (3 K) (4.18 J K1 g1) = 125 J of energy.
The metal sample lost this same quantity of energy, undergoing a
temperature drop of 182C as the result. The specific heat capacity of the
metal is (125 J) / (2.40 g 182 K) = 0.287 J K1 g1.

199
Notice that no "formula" is required here, as long as you know the units of
specific heat; you simply place the relevant quantities in the numerator or
denominator to make the units come out correctly.

11.3 Calorimetry

How is the heat given out or absorbed in a chemical reaction


determined experimentally? The measurement of q is generally
known as calorimetry.

Coffee cup calorimeter


A very simple calorimeter can be made from a foam plastic coffee
cup, which is often used in student laboratories see figure 11.1.

Figure 11.1: Coffee-cup calorimeter

The reaction is carried out in the cup. Suppose that the reaction is
exothermic reaction. Since the foam is a good insulator, heat does not
escape but results in increasing the temperature of the contents of the
calorimeter. Heat given to the solution,
qsoln = (specific heat of solution) (grams of solution) T
Then, if the reaction gives heat, the solution takes it,
qsoln = qreaction
For dilute aqueous solutions, the specific heat of the solution will
approximately be the same as that of water, 4.18 J/g K.
If the specific heat of the solution is not available, the calorimeter can
be calibrated by doing another experiment, in which a known amount
200
of heat is supplied to the calorimeter, keeping the contents as nearly
the same as in the actual experiment. This heat can be supplied
electrically or by carrying out a reaction which releases a known
amount of heat.

A very simple calorimetric determination of the heat of the reaction


H+(aq) + OH(aq) H2O(l)
could be carried out by combining equal volumes of 0.1 M solutions
of HCl and NaOH, initially at 25C. Since this reaction is exothermic,
a quantity of heat q will be released into the solution. What we
actually measure is the resulting temperature rise; if we multiply T
by the specific heat capacity of the solution (which will be close to
that of pure water, 4.184 J/g K), we obtain the number of joules of
heat released into each gram of the solution, and q can then be
calculated from the mass of the solution.

For reactions that cannot be carried out in dilute aqueous solution, the
reaction vessel is commonly placed within a larger insulated container
of water. During the reaction, heat passes between the inner and outer
containers until their temperatures become identical. Again, the
temperature change of the water is observed, but in this case the value
of q cannot be found just from the mass and the specific heat capacity
of the water, for we now have to allow for the absorption of some of
the heat by the walls of the inner vessel. Instead, the calorimeter is
calibrated by measuring the temperature change that results from
the introduction of a known quantity of heat. The resulting
calorimeter constant, expressed in J K1, can be regarded as the heat
capacity of the calorimeter. The known source of heat is usually
produced by passing an electric current through a resistor within the
calorimeter.

Bomb calorimeter
For reactions involving gases, such as combustion reactions, a special
type of calorimeter is used. It is schematically shown in figure 11.2.
Since the process takes place at constant volume, the reaction vessel
must be constructed to withstand the high pressure resulting from the
combustion process, which amounts to a confined explosion. The
vessel is usually called a bomb, and the technique is known as bomb
201
calorimetry. In order to ensure complete combustion, the bomb is
initially charged with pure oxygen usually up to 25 atmosphere
pressure. The reaction is initiated by discharging a capacitor through
a thin wire which ignites the mixture. The bomb is immersed in water
contained in an outer jacket. The jacket has a stirrer and a
thermometer. The rise in temperature in the surrounding water is
noted.

If Ccalorimeter is known from a calibration experiment, the heat released


by the reaction (and absorbed by water) is calculated by using the
equation:
qcalorimeter = Ccalorimeter T = qreaction.
In the calibration experiment, a known amount of heat is supplied to
the calibration (usually by burning benzoic acid), and T is measured.

Thermometer and stirrer


Electrical wire Filled with O2 at high P

Water Sample

Figure 11.2: Schematic diagram of a bomb calorimeter

Example 11.3:
(a) 32.7 kJ of energy is supplied electrically to a calorimeter. The
temperature of the calorimeter rises from 24.70 to 26.83 oC. What is the
heat capacity of the calorimeter?
(b) A reaction was carried out in the above calorimeter. The temperature
rose by 2.00 oC. Find the heat released by the reaction.
Solution:
(a) T = 2.13 K (or oC); q = 32.7 kJ
Ccal = q / T = 32.7 kJ / 2.13 K = 15.4 kJ/K

202
(b) qcal = heat released by the reaction =
Ccal T = 15.4 kJ/K 2.0 K = 30.8 kJ

Example 11.4:
A sample of biphenyl (C6H5)2 weighing 0.526 g was ignited in a bomb
calorimeter initially at 25C, producing a temperature rise of 1.91 K. In a
separate calibration experiment, a sample of benzoic acid C6H5COOH
weighing 0.825 g was ignited under identical conditions and produced a
temperature rise of 1.94 K. For benzoic acid, the heat of combustion at
constant pressure is known to be 3226 kJ mol1. Use this information to
determine the standard enthalpy of combustion of biphenyl.
Solution:
The heat capacity of the calorimeter, Ccal, is given by
3226 kJ mol-1 0.825 g
11.2 kJ K -1
1.94 K 123 g mol-1
The heat released by the combustion of the biphenyl at constant pressure
is then:
- 11.2 kJ K -1 1.91 K 154 g mol-1
6260 kJ mol-1
0.526 g
The negative sign indicates that heat is released in this process, from the
reaction equation:
(C6H5)2(s) + 14 O2(g) 12 CO2(g) + 5 H2O(l)
This is the amount of heat that is lost by the system if reaction takes place
at constant pressure and the temperature is restored to its initial value.

Although calorimetry is simple in principle, its practice is a highly


exacting art, especially when applied to processes that take place
slowly or involve very small heat changes, such as the germination of
seeds.

11.4 Enthalpy

Enthalpy change
If an object absorbs a quantity of heat, say 160 kJ, at a constant
pressure, its enthalpy (H) is said to have increased by 160 kJ as a

203
result of the process. The object has then undergone a change in its
enthalpy during the process, H = 160 kJ in this example. Enthalpy is
not measured directly, only its change, or H, which equals q at
constant pressure, is measured. If, on the other hand, a reaction gives
out 160 kJ of heat, the energy of the reactants will decrease by 160 kJ,
and H = 160 kJ. This later case is depicted in figure 11.3 below.
The arrow indicates H = 160 kJ. H is negative for an exothermic
reaction and positive for an endothermic reaction.

H of reactants

H
H of products

Figure 11.3: Change in enthalpy for an exothermic reaction

State property
Enthalpy is a state property which means that it always depends on
the current state of the system and is independent of how the system
has been brought in that state. Lets consider two processes, (a) in
which 100 g of water at 25oC is heated to 35oC, the change in
enthalpy, H, would be 4.18 kJ; and (b) in which water is first cooled
from 25oC to 18oC and then heated to 35oC, in this case H would
also be 4.18 kJ. Process (a) is shown by a double headed arrow on the
left in figure 11.4 below. Process (b) is shown by the two single
headed arrows on the right. In both cases, the net H is the same.

(a) (b)
o
35 C
H
25oC

18oC

Figure 11.4: Heating water at 250C to 350C: (a) directly to 350C; and
(b) via 180C to 350C

204
We will apply these ideas to actual processes, first physical, and then
chemical processes.

Vaporization
Imagine one mole of liquid water at 100oC is changed to steam by
supplying it heat. The enthalpy of H2O(g) is higher than that of liquid
water, since heat has been given to the liquid to change it to vapour.
The process would be endothermic and enthalpy change would be
positive. The enthalpy change when one mole of a liquid vaporizes is
called the enthalpy of vaporization, Hvaporization. The enthalpy of
vaporization of water at 100oC is +40.7 kJ/mol:
H = Hvapour Hliquid = +40.7 kJ/mol at 100 oC
for the process: H2O(l) H2O(g)

Melting
Likewise, the enthalpy change of H2O when it melts is positive, since
heat is supplied to ice to melt it. The process of melting, or fusion, is
thus endothermic. The enthalpy change when one mol of a substance
melts is called enthalpy of fusion: Hfusion> 0. The enthalpy of fusion
of water at 0oC is 6.0 kJ/mol.

Freezing
The process of freezing is opposite that of melting. The change in
enthalpy per mole when a liquid turns into a solid is called the
enthalpy of freezing. Liquids are frozen by taking out heat from them
(cooling). So, the process would be exothermic and H would be
negative. The enthalpy of freezing of water at 0oC is 6.0 kJ/mol:
Hfreezing = 6.0 kJ/mol.

Sublimation

Sublimation is the direct conversion of a solid into its vapour. The


enthalpy of sublimation is the enthalpy change when one mole of a
solid sublimes. Since the change is endothermic, enthalpy of
sublimation would be positive. Furthermore, since enthalpy is a state

205
property, we simply add enthalpy of fusion and enthalpy of
vaporization to find enthalpy of sublimation.
Hvaporization = Hvapour Hliquid
Hfusion = Hliquid Hsolid
Hence, we get by adding the above two equations:
Hvaporization + Hfusion = Hvapour Hsolid = Hsublimation

Enthalpy of reaction
Consider, for example, the oxidation of a lump of sugar to carbon
dioxide and water:
C12H22O11(s) + 12 O2(g) 12 CO2(g) + 11 H2O(l)
This process can be carried out in many ways, for example by burning
the sugar in air, or by eating the sugar and letting your body carry out
the oxidation. Although the mechanisms of the transformation are
completely different for these two pathways, the overall change in the
enthalpy of the system will be identical if the reactants and products
are the same in each transformation. The enthalpy of the products
minus that of the reactants in the above example is found to be -5606
kJ.
The same quantity of heat is released whether the sugar is burnt in the
air or oxidized in a series of enzyme-catalyzed steps in your body.
When a chemical equation is written together with its H, it is called a
thermochemical equation. The thermochemical equation for the
process of combustion of sucrose would be:
C12H22O11(s) + 12 O2(g) 12 CO2(g) + 11 H2O(l) H = 5606 kJ
If we write the reverse equation, the sign of H would change. The
thermochemical equation then becomes:
12 CO2(g) + 11 H2O(l) C12H22O11 + 12 O2(g) H = +5606 kJ

The following points should be kept in mind when writing a


thermochemical equation:
Thermochemical equations for reactions taking place in solution
must also specify the concentrations of the dissolved species. For

206
example, the enthalpy of neutralization of a strong acid by a
strong base is
H+(aq, 1M) + OH(aq, 1M) H2O(l) H = 56.9 kJ mol1
The abbreviation aq refers to the hydrated ions as they exist in
aqueous solution.
Since most thermochemical equations are written for the standard
conditions of 298 K and 1 atm pressure, we can leave these
quantities out if these conditions apply both before and after the
reaction. If, under these same conditions, the substance is in its
preferred (most stable) physical state, then the substance is said to
be in its standard state. Thus the standard state of water at 1 atm
is the solid below 0C, and the gas above 100C. A
thermochemical quantity such as H that refers to reactants and
products in their standard states is denoted by H.
Any thermodynamic quantity such as H that is associated with a
thermochemical equation always refers to the number of moles of
substances explicitly shown in the equation. Thus, for the
synthesis of water we can write
2 H2(g) + O2(g) 2 H2O(l) H = 572 kJ
If now the chemical equation is multiplied or divided by a factor,
the value of H is also multiplied or divided by the same factor.
Thus, when the above equation is divided by 2, we get the
following thermochemical equation:
H2(g) + O2(g) H2O(l) H = 286 kJ

11.5 Hesss Law and its Applications

Hesss law
You probably know that two or more chemical equations can be
combined algebraically to give a new equation. Even before the
science of thermodynamics developed in the late nineteenth century, it
was observed that the heats associated with chemical reactions can be
combined in the same way to yield the heat of another reaction. For
example, the standard enthalpy changes for the oxidation of graphite
and diamond can be combined to obtain H for the transformation
between these two forms of solid carbon, a reaction that cannot be
studied experimentally.
207
C(graphite) + O2(g) CO2(g) H = 393.51 kJ mol1
C(diamond) + O2(g) CO2(g) H = 395.40 kJ mol1
Subtraction of the second reaction from the first (i.e., writing the
second equation in reverse and adding it to the first one) yields
C(graphite) C(diamond) H = +1.89 kJ mol1
This principle, known as Hesss law of independent heat summation is
a direct consequence of the enthalpy being a state function. Hesss
law is one of the most powerful tools of chemistry, for it allows the
change in the enthalpy of huge numbers of chemical reactions to be
predicted from a relatively small base of experimental data.

Enthalpy of formation
The standard enthalpy of formation of a compound is defined as the
heat associated with the formation of one mole of the compound from
its elements in their standard states.
The enthalpy change for a chemical reaction is the difference
H = Hproducts Hreactants
If the reaction in question represents the formation of one mole of the
compound from its elements in their standard states, as in
H2(g) + O2(g) H2O(l) H = 285.8 kJ
then we can arbitrarily set the enthalpy of the elements in their
standard states to zero and write
Hf = Hf products Hf reactants= 285.8 kJ 0 = 285.8 kJ mol1
which defines the standard enthalpy of formation of water at 298K.

Calculation of Horxn from Hof


In general, the standard enthalpy change for a reaction, Horxn , is
given by the expression:
Horxn = Hf oproducts Hf oreactants
in which the Hf terms indicate the sums of the standard enthalpies
of formations of all products and reactants. The above definition is
one of the most important in chemistry because it allows us to predict
the enthalpy change of any reaction without knowing any more than

208
the standard enthalpies of formation of the products and reactants,
which are widely available in tables. Table 11.1 gives the values of
the enthalpies of formation of some compounds.
Table 11.1:

Compound Hfo
C6H6(l) -49.04 kJ/mol
CO(g) -110.5 kJ/mol
CO2(g) -393.5 kJ/mol
HF(g) -271.1 kJ/mol
H2O(g) -241.8 kJ/mol
H2O (l) -285.8 kJ/mol
NH3(g) -46.1 kJ/mol
NO(g) 90.25 kJ/mol
NO2(g) 33.18 kJ/mol
N2O4(g) 9.16 kJ/mol
SO2(g) -296.8 kJ/mol
SO3(g) -395.7 kJ/mol
All values are in the units of kJ/mol and the physical conditions of
298.15 K and under 1 atm of pressure, which is referred to as the
"standard state", and are the conditions in which you will generally
find values of standard enthalpies of formation.
Note that while the majority of the values of standard enthalpies of
formation are exothermic, or negative, there are a few compounds
such as NO(g) and N2O4(g) that actually require energy from its
surroundings during its formation; these endothermic compounds are
generally unstable.
[Source: http://chemwiki.ucdavis.edu]

Example 11.5:
Calculate Horxn of the following reaction:

209
C6H6(l) + 7 O2(g) 6 CO2(g) + 3 H2O (l)
Given the following values of the standard enthalpy of formation:
Hf o [C6H6(l)] = 49.04 kJ/mol; Hf o [CO2(g)] = 393.5 kJ/mol;
Hf o [H2O(l)] = 285.8 kJ/mol
Solution:
The standard reaction enthalpy, Horxn
= 3Hf o[H2O(l)] + 6Hf o[CO2(g)] Hf o[C6H6(l)] 7Hf o[O2(g)]
= 3(-285.8 kJ) + 6(-393.5kJ) (-49.04 kJ) 7(0) = -3169 kJ

The following examples illustrate some important aspects of the


standard enthalpy of formation of substances.
The thermochemical equation defining Hf is always written in
terms of one mole of the substance in question:
N2(g) + 1H2(g) NH3(g) H = 46.1 kJ (for 1 mol NH3)
Hence, Hf (NH3) = 46.1 kJ/mol
A number of elements, of which sulfur and carbon are common
examples, can exist in more than one solid crystalline form. The
standard heat of formation of a compound is always taken in
reference to the forms of the elements that are most stable at 25C
and 1 atm pressure. In the case of carbon, this is the graphite, rather
than the diamond form:
C(graphite) + O2(g) CO2(g) H = 393.5 kJ mol1
C(diamond) + O2(g) CO2(g) H = 395.8 kJ mol1
The physical state of the product of the formation reaction must
be indicated explicitly if it is not the most stable one at 25C and 1
atm pressure:
H2(g) + O2(g) H2O(l) H = 285.8 kJ mol1
H2(g) + O2(g) H2O(g) H = 241.8 kJ mol1
Notice that the difference between these two H values is just
the heat of vaporization of water.

Although the formation of most molecules from their elements is an


exothermic process, the formation of some compounds is mildly
endothermic:
210
N2(g) + O2(g) NO2(g) H = +33.2 kJ mol1
A positive heat of formation is frequently associated with instability -
the tendency of a molecule to decompose into its elements, although it
is not in itself a sufficient cause. In many cases, however, the rate of
this decomposition is essentially zero, so it is still possible for the
substance to exist. In this connection, it is worth noting that all
molecules will become unstable at higher temperatures.

The thermochemical reactions that define the heats of formation of


most compounds cannot actually take place; for example, the
synthesis of methane from its elements:
C(graphite) + 2 H2(g) CH4(g)
cannot be observed directly owing to the large number of other
possible reactions between these two elements. However, the
standard enthalpy change for such a reaction can be found indirectly
from other data, as explained in the next section.

The standard enthalpy of formation of gaseous atoms from the


element is known as the heat of atomization. Heats of atomization are
always positive, and are important in the calculation of bond energies.
Fe(s) Fe(g) H = +417 kJ mol1

Because most substances cannot be prepared directly from their


elements, heats of formation of compounds are seldom determined by
direct measurement. Instead, Hesss law is employed to calculate
enthalpies of formation from more accessible data. The most
important of these are the standard enthalpies of combustion. Most
elements and compounds combine with oxygen, and many of these
oxidations are highly exothermic, making the measurement of their
heats relatively easy. For example, by combining the heats of
combustion of carbon, hydrogen, and methane, we obtain the standard
enthalpy of formation of methane, which as we noted above, cannot
be determined directly.

211
Enthalpy diagrams and their uses
Comparison and interpretation of enthalpy changes is materially aided
by a graphical construction in which the relative enthalpies of various
substances are represented by horizontal lines on a vertical energy
scale. The zero of the scale can be placed anywhere, since energies
are always arbitrary; it is generally most useful to locate the elements
at zero energy, which reflects the convention that their standard
enthalpies of formation are zero.
The very simple enthalpy diagram for carbon and oxygen and its two
stable oxides shows the changes in enthalpy associated with the
various reactions this system can undergo (see figure 11.5). Notice
how Hesss law is implicit in this diagram; we can calculate the
enthalpy change for the combustion of carbon monoxide to carbon
dioxide, for example, by subtraction of the appropriate arrow lengths
without writing out the thermochemical equations in a formal way.
The zero-enthalpy reference states refer to graphite, the most stable
form of carbon, and gaseous oxygen. All temperatures are 298 K.

Figure 11.5: Enthalpy diagram for carbon, oxygen and the


two stable carbon oxides
Figure 11.6, the enthalpy diagram for the hydrogen-oxygen system,
shows the known stable configurations of these two elements.
Reaction of gaseous H2 and O2 to yield one mole of liquid water
releases 285 kJ of heat . If the H2O is formed in the gaseous state,
the energy release will be smaller.
212
Fig. 11.6: Enthalpy diagram for hydrogen, oxygen and the
two stable hydrogen oxides

11.6 Bond Enthalpies

Bond enthalpy
Bond enthalpy is the enthalpy change for the breaking of a particular
bond in one mole of a gaseous substance. Thus the enthalpy change
associated with the reaction:
HI(g) H(g) + I(g) Ho = +299 kJ/(mol of HI)
is the enthalpy of dissociation of the HI molecule; it is also the bond
enthalpy of the hydrogeniodine bond in this molecule. Bond
enthalpy is expressed in kJ/mol. Since energy needs to be supplied to
break bonds, bond enthalpy is always positive. Stronger bonds have
higher bond enthalpies.

As the bond order increases, bonds become stronger and bond


enthalpy increases. The triple bond between C and C (C C) has a
bond enthalpy of 839 kJ/mol compared to 614 kJ/mol for a double
bond, and 348 kJ/mol for a single bond. The strength of these bonds
decreases as:
CC > CC > CC

213
Average bond enthalpy

The total bond enthalpy of a molecule can be thought of as the sum of


the enthalpies of the individual bonds. This principle, known as
Paulings Rule, is only an approximation, because the energy of a
given type of bond is not really a constant, but depends somewhat on
the particular chemical environment of the two atoms. In other words,
all we can really talk about is the average energy of a particular kind
of bond. For example, for CO, the average is taken over a
representative sample of compounds containing this type of bond,
such as CO, CO2, COCl2, (CH3)2CO, CH3COOH, etc.
Estimation of average bond enthalpy
How can the values of bond enthalpies be determined? Let us
consider again the bond enthalpy in HI.
Since the reaction HI(g) H(g) + I(g) cannot be studied directly, its
enthalpy change, or in other words the HI bond enthalpy, is
calculated from the appropriate standard enthalpies of formation:
H2(g) H(g) H = +218 kJ
I2(g) I(g) H = +107 kJ
H2(g) + I2(g) HI(g) H = +26.5 kJ
Using the above three equations, we obtain:
HI(g) H(g) + I(g) H = +299 kJ

Bond enthalpies are important properties of chemical bonds, and it is


very important to be able to estimate their values from other
thermochemical data. The total bond enthalpy of a more complex
molecule such as ethane can be found from a combination of the
following reactions:

(1) C2H6(g) 2 C(graphite) + 3 H2(g) H = +84.7 kJ


(2) 3 H2(g) 6 H(g) H = +1308 kJ
(3) 2 C(graphite) 2 C(g) H = +1430 kJ
Adding equations (1), (2) and (3) together will give equation (4):

214
C2H6(g) 2 C(g) + 6H(g) H = +2823 kJ
which gives the total bond enthalpy for ethane, C2H6(g): +2823 kJ.

Despite the lack of strict additivity of bond energies, Paulings Rule is


extremely useful because it allows one to estimate the heats of
formation of compounds that have not been studied, or have not even
been prepared. Thus in the foregoing example, if we know the
enthalpies of the CC and CH bonds from other data, we could
estimate the total bond enthalpy of ethane, and then work back to get
some other quantity of interest, such as ethanes enthalpy of
formation. By assembling a large amount of experimental information
of this kind, a consistent set of average bond energies can be obtained.

Using bond enthalpies to estimate reaction enthalpies


Enthalpies of formation of compounds as well as reaction enthalpies
can be estimated by using the values of average bond enthalpies see
Table 11.1.
Table 11.2: Selected average bond enthalpies

Average enthalpies of some single bonds (kJ/mol)


H C N O F Cl Br I Si
H 436 415 390 464 569 432 370 295 395
C 345 290 350 439 330 275 240 360
N 160 200 270 200 270
O 140 185 205 185 200 370
F 160 255 160 280 540
Cl 243 220 210 359
Br 190 180 290
I 150 210
Si 230
All the bond enthalpies of the bonds that are broken are added as well
as the bond enthalpies of the bonds that are formed. These aggregate
values are then used to calculate the net enthalpy required by the
reaction or released by it. This is illustrated in example 11.6 below.

215
Example 11.6:
Estimate Hrxn for the reaction:
H CH3(g) + Br Br(g) Br CH3(g) + H Br(g)
Solution:
Bonds broken: 1 mol H C, 1 mol Br Br
Bonds made: 1 mol Br C, 1 mol H Br
From Table 11.1: HB values are
415 kJ/mol for H C, 190 kJ/mol for Br Br,
275 kJ/mol for Br C, and 370 kJ/mol for H Br.
Energy needed to break the bonds = 415 kJ + 190 kJ = 605 kJ
Energy given out in making the bonds = 275 + 370 = 645 kJ
Net energy change = 645 605 = 40 kJ. This energy is given out, so:
Hrxn = 40 kJ.
The reaction is exothermic because bonds being formed are stronger than
those being broken.

11.7 Exercises
11.1 Given the thermochemical equation
MgCl2(s) + 2Na (l) 2NaCl (s) + Mg(s) Ho = 180.2 kJ
Calculate the heat evolved when 7.34 g Mg is produced by this
reaction.

11.2 How much heat is given out when 18.7 g copper metal is cooled
from 32.0 oC to 27.4 oC? The specific heat capacity of Cu(s) is 0.387
J g1 K1.

11.3 1.700 g benzoic acid (C6H5COOH) is burned in a bomb calorimeter


whose heat capacity is 8320 J/K. Calculate the increase in the
temperature of the calorimeter. The enthalpy of combustion of
benzoic acid is 3226.7 kJ/mol.

11.4 Calculate the heat needed to change one mole ice at 0.0 oC to liquid
water at 22.8 oC. The enthalpy of fusion of water is 6.01 kJ at 0 oC;
specific heat capacity of water is 4.18 J g1 K1. Hfusion =6.01
kJ/mol.

216
11.5 250.0 mL of 0.940 M HCl is mixed with 250.0 mL of 0.470 M
Ba(OH)2 in a coffee cup calorimeter. The initial temperature of both
solutions is 23.80 oC. The thermochemical equation is:
H+(aq) + OH(aq) H2O(l) Ho = 56.2 kJ/mol
Calculate the final temperature of the mixture.

11.6 A 1.430 g sample of naphthalene (C10H8) was burned in a bomb


calorimeter. The temperature increased by 2.32 oC. What was the
heat capacity was of the calorimeter? The enthalpy of combustion of
naphthalene is 5151 kJ/mol.

11.7 Calculate Ho for the reaction:


Fe2O3(s) + 3C(graphite) 2Fe(s) + 3CO(g)
using the thermochemical equation:
2CO(g) + O2(g) 2CO2(g) Ho = 583 kJ
and the values Hof (Fe2O3(s)) = 822.3 kJ/mol and Hof (CO2(g)) =
393.5kJ/mol.

11.8 Given the thermochemical equations:


N2O4(g) 2NO2(g) Ho = 59.73 kJ
2NO(g) + O2(g) N2O4(g) Ho = 171.07 kJ
and the value Hof (NO2(g)) = 34.0 kJ/mol.
Find Hof for NO(g).

11.9 Use the bond enthalpies given in Table 11.1 to estimate the enthalpy
changes for the following reactions:
(a) H2(g) + Cl2(g) 2HCl(g)
(b) Cl2(g) + C2H4(g) C2H4Cl2(g)

217
CHAPTER 12
Solutions
12.1 Introduction to Solutions

When you stop to think about it, we live in a world of solutions! The
air we breathe is a huge gaseous solution, the oceans are solutions of
about fifty different salts in water, and many of the rocks and minerals
of the earth are solid solutions. We ourselves are largely aqueous
solutions, most of it within our cells (whose water content contributes
to about half our body weight) and in our blood plasma and the
interstitial fluid, that bathes our cells (about 5 L in the adult). So in
order to understand the world in which we live and the organisms that
inhabit it, we need to know something about solutions.

Solutions
Solutions are homogeneous (single phase) mixtures of two or more
components. For convenience, we often refer to the majority
component as the solvent; minority components are solutes. However,
there is really no fundamental distinction between them.
Solutions play a very important role in Chemistry because they allow
intimate and varied encounters between molecules of different kinds, a
condition that is essential for rapid chemical reactions to occur.
Several, more explicit reasons can be cited for devoting a significant
amount of effort to the subject of solutions:
For the reason stated above, most chemical reactions that are
carried out in the laboratory and in industry, and that occur in
living organisms, take place in solution.
Solutions are so common; very few pure substances are found in
nature.
Solutions provide a convenient and accurate means of introducing
known small amounts of a substance to a reaction system.
Advantage is taken of this in the process of titration, for example.
The physical properties of solutions are sensitively influenced by
the balance between the intermolecular forces of like and unlike
(solvent and solute) molecules. The physical properties of

218
solutions thus serve as useful experimental probes of these
intermolecular forces.
We usually think of a solution as a liquid made by adding a gas, a
solid or another liquid solute in a liquid solvent. Actually, solutions
can exist as gases and solids as well. Gaseous mixtures don't require
any special consideration beyond what you learned about Daltons
Law earlier. Solid solutions are very common; most natural minerals
and many metallic alloys are solid solutions.
Still, it is liquid solutions that we most frequently encounter and must
deal with. Experience has taught us that sugar and salt dissolve
readily in water, but that oil and water dont mix. Actually, this is
not strictly correct, since all substances have at least a slight tendency
to dissolve in each other. This raises two important and related
questions: why do solutions tend to form in the first place, and what
factors limit mutual solubility of substances?

Solubility and saturation

Saturated solution is the solution in which no more solute can be


dissolved. It means that the solvent has dissolved solute to its
maximum ability.
The solubility of a solute is defined as the concentration of the
saturated solution. Similarly, molar solubility is molar
concentration of the saturated solution. Thus, if one litre of a solvent
dissolves a maximum of 5 moles of sodium chloride, the molar
solubility of sodium chloride in that solvent would be 5 mol/L.

12.2 Concentration; Concentration Units

Concentration is a general term that expresses the quantity of solute


contained in a given amount of solution. Various ways of expressing
concentration are in use; the choice is usually a matter of convenience
in a particular application. You should become familiar with all of
them.

219
Mass percent, volume percent, mass by volume percent
In the consumer and industrial world, the most common method of
expressing the concentration is based on the quantity of solute in a
fixed quantity of solution. The quantities referred to here can be
expressed in mass, in volume, or both (i.e., the mass of solute in a
given volume of solution.) In order to distinguish among these
possibilities, the abbreviations (w/w), (v/v) and (w/v) are used.
In most applied fields of chemistry, (w/w) measure is often used, and
is commonly expressed as weight percent concentration, or simply
"percent concentration". For example, a solution made by dissolving
10 g of salt with 200 g of water contains "1 part of salt per 20 g of
water". It is usually more convenient to express such concentrations
as "parts per 100", which we all know as "percent". The solution
described is then a "5% (w/w) solution" of NaCl in water.
In clinical chemistry, (w/v) is commonly used, with weight expressed
in grams and volume in mL.

Example 12.1:
The Normal Saline solution used in medicine for nasal irrigation, wound
cleaning and intravenous drips is a 0.91% (w/v) solution of sodium
chloride in water. How would you prepare 1.5 L of this solution?
Solution:
The solution will contain 0.91 g of NaCl in 100 mL of water, or 9.1 g in 1
L. Thus you will add (1.5 9.1g) = 13.6 g of NaCl to 1.5 L of water.

Example 12.2:
Describe how you would prepare 30 g of a 20 percent (w/w) solution of
KCl in water.
Solution:
The weight of potassium chloride required is 20% of the total weight of
the solution, or 0.2 (30 g) = 6.0 g of KCl. The remainder of the solution
(30 6 = 24) g consists of water. Thus you would dissolve 6.0 g of KCl
in 24 g of water.

220
Percent means parts per 100; we can also use parts per thousand (ppt)
for expressing concentrations in grams of solute per kilogram of
solution. For more dilute solutions, parts per million (106; ppm) and
parts per billion (109; ppb) are used. These terms are widely
employed to express the amounts of trace pollutants in the
environment.
It is sometimes convenient to base concentration on a fixed volume,
either of the solution itself, or of the solvent alone. In most instances,
a 5% by volume solution of a solid will mean 5 g of the solute
dissolved in 100 ml of the solvent.

Example 12.3:
Fish, like all animals, need a supply of oxygen, which they obtain from
oxygen dissolved in the water. The minimum oxygen concentration
needed to support most fish is around 5 ppm (w/v). How many moles of
O2 per litre of water does this correspond to?
Solution:
5 ppm (w/v) means 5 grams of oxygen in one million mL (1000 L) of
water, or 5 mg per litre. This is equivalent to
(0.005 g)/(32.0 g mol1) = 1.6 104 mol.

If the solute is itself a liquid, volume/volume measure usually refers to


the volume of solute contained in a fixed volume of solution (not
solvent). The latter distinction is important because volumes of mixed
substances are not strictly additive. These kinds of concentration
measure are mostly used in commercial and industrial applications.
The "proof" of an alcoholic beverage is the (v/v) percent, multiplied
by two; thus a 100-proof vodka has the same alcohol concentration as
a solution made by adding sufficient water to 50 mL of alcohol to give
100 mL of solution.

Molarity: mole/volume
Molarity is the method most used by chemists to express
concentration, and it is the one most important for you to master.
Molar concentration (molarity) is the number of moles of solute per
litre of solution.

221
The important point to remember is that the volume of the solution is
different from the volume of the solvent; the latter quantity can be
found from the molarity only if the densities of both the solution and
of the pure solvent are known. Similarly, calculation of the weight
percentage concentration from the molarity requires information of
density.

Example 12.4:
How would you make 120 mL of a 0.10 M solution of potassium
hydroxide in water?
Solution:
The amount of KOH required is
(0.120 L) (0.10 mol L1) = 0.012 mol.
The molar mass of KOH is 56.1 g, so the mass of KOH required is
(0.012 mol) (56.1 g mol1) = 0.67 g.
We would dissolve this mass of KOH in a volume of water that is less
than 120 mL, and then add sufficient water to bring the volume of the
solution up to 120 mL.
Comment: if we had simply added the KOH to 120 mL of water, the
molarity of the resulting solution would not be the same. This is because
volumes of different substances are not strictly additive when they are
mixed. Without actually measuring the volume of the resulting solution,
its molarity would not be known.

Mole fraction: mole/mole


This is the most fundamental of all methods of concentration measure,
since it makes no assumptions at all about volumes. The mole
fraction of substance i in a mixture is defined as
ni
i
n j
j

in which nj is the number of moles of substance j, and the summation


is over all substances in the solution. Mole fractions run from zero
(substance not present) to unity (the pure substance).
The sum of all mole fractions in a solution is, by definition, unity:

222
i 1
i
In the case of ionic solutions, each kind of ion acts as a separate
component.

Example 12.5:
Find the mole fraction of water in a solution prepared by dissolving 4.5 g
of CaBr2 in 84.0 mL of water. Assume the species in the solution to be
Ca2+ and Br-, in addition to water.
Solution:
The molar mass of CaBr2 is 200 g/mol. Thus the number of moles of
CaBr2 in the solution is (4.50 g) / (200 g mol1) = 0.0225 mol.
Because this salt is completely dissociated in solution, the solution will
contain 0.0225 mol of Ca2+ and (2 0.0225) = 0.0450mol of Br ions.
84.0 mL of H2O has a mass very close to 84.0 g at its assumed density of
1.00 g mL1. The number of moles of water is
(84.0 g) / (18.0 g mol1) = 4.67 mol.
The mole fraction of water is then
(4.67 mol) / (0.0225 + 0.0450 + 4.67) mol = 4.67 / 5.47 = 0.986.
Thus, H2O constitutes 986 out of every 1000 molecules + ions in the
solution.

Molality: mole/mass
A 1 molal solution contains one mole of solute per 1 kg of solvent.
Molality is a hybrid concentration unit, retaining the convenience of
mole measure for the solute, but expressing it in relation to a
temperature independent mass rather than a volume. Molality, like
mole fraction, is used in applications dealing with certain physical
properties of solutions.

Conversion between concentration units


Anyone doing practical chemistry must be able to convert one kind of
concentration measure into another. The important point to remember
is that any conversion involving molarity requires knowledge of the
density of the solution.

223
Example 12.6:
A solution prepared by dissolving 66.0 g of urea (NH2)2CO in 950 g of
water had a density of 1.018 g mL1. Express the concentration of urea in
a) weight-percent; b) mole fraction; c) molarity; d) molality.
Solution:
a) The mass percent of solute is
(100%) (66.0 g) / (950 + 66.0 g) = 6.50%
b) The molar mass of urea is 60.06, so its number of moles is
(66.0 g) /(60.06 g mol1) = 1.10 mol.
The number of moles of H2O is
(950 g) / (18.02 g mol1) = 52.8 mol.
Mole fraction of urea: (1.10 mol) / (1.10 + 52.8 mol) = 0.0204
c) Volume of solution is
(66.0 + 950) g / (1018 g L1) = 0.998 L.
The number of moles of urea (from a) is 1.10 mol.
Its molarity is then (1.10 mol) / (0.998 L) = 1.10 mol L1.
d) The molality of urea is
(1.10 mol) / (0.950) kg = 1.16 mol kg1.

Example 12.7:
Ordinary dry air contains 21% (v/v) oxygen. How many moles of O2 can
be inhaled into the lungs of a typical adult woman with a lung capacity of
4.0 L?
Solution:
The number of molecules (and thus the number of moles) in a gas is
directly proportional to its volume (Avogadros law), so the mole fraction
of O2 is 0.21.
The number of moles of a gas occupying 4.0 L volume at 25 C and 1.0
atm pressure are given by the Ideal Gas Law:
n = PV/RT = (1.0) (4.0) /(0.08206 298) = 0.164 mol
The number of moles of O2 in 4.0 L of air will be
mole fraction n = 0.21 0.164 = 0.034 mol O2.

224
Dilution calculations

These kinds of calculations arise frequently in both laboratory and


practical applications. If you have a thorough understanding of
concentration definitions, they are easily tackled. The most important
things to bear in mind are
Concentration is inversely proportional to volume.
Molarity is expressed in mol L1, so it is usually more convenient
to express volumes in litres rather than in millilitres.
Use the principles of unit cancellations to determine what to
divide by what.

Example 12.8:
Commercial hydrochloric acid is available as a 10.17 molar solution.
How would you use this to prepare 500 mL of a 4.00 molar solution?
Solution:
The desired solution requires (0.50 L) (4.00 mol L1) = 2.0 mol of HCl.
This quantity of HCl is contained in (2.0 mol) / (10.17 mol L1) = 0.197 L
of the concentrated acid.
So one would add 197 mL of the concentrated acid to some water, and
then add more water to make the total volume 500 mL.

Example 12.9:
Calculate the molarity of the solution produced by adding 120 mL of 6.0
M HCl to 150 mL of 0.15 M HCl. What important assumption must be
made here?
Solution:
The assumption, of course, is that the density of HCl within this
concentration range is constant, meaning that their volumes will be
additive.
Moles of HCl in 1st solution: (0.120 L) (6.0 mol L1) = 0.72 mol
Moles of HCl in 2nd solution: (0.150 L) (0.15 mol L1) = 0.0225 mol
Molarity of mixture is
(0.72 + 0.0225) mol / (0.120 + 0.150) L = 2.7 mol L1.

225
12.3 Solution Types; Why Solutions Form (or Dont)

Why do solutions form in the first place except when they don't?
In this section, we look at some of the underlying physical phenomena
that govern the transformation of mixtures into solutions.
To see what type of substances are likely to be mutually soluble, let us
consider the individual steps that must be carried out when a solute is
dissolved in a solvent:
If the solute is a solid or liquid, it must first be dispersed that
is, its molecular units must be pulled apart. This requires energy,
and so this step always works against solution formation.
Similarly, solvent molecules have to separate. Solute particles
are then incorporated among solvent molecules. This step
releases energy. If the second step releases more energy than is
consumed in the first step, this will favour solution formation,
and we can generally expect the solute to be soluble in the
solvent.

Figure 12.1: Dark spheres represent solvent molecules and the


light sphere represents a solute

The solution process is exothermic if AB attractions stronger than


AA + BB. It is endothermic if attractions between like molecules
are stronger than those between unlike molecules.

226
Even if the dissolution process is slightly endothermic, there is a third
important factor, the entropy increase that will very often favour the
dissolved state.

"Like dissolves like"

Whereas all gases will mix to form solutions regardless of the


proportions, liquids are much more fussy. Some liquids, such as
ethanol and water, are miscible in all proportions. Others, like the
proverbial oil and water, are not; each liquid has only a limited
solubility in the other, and once either of these limits is exceeded, the
mixture separates into two phases.
A useful general rule is that liquids are completely miscible when
their intermolecular forces are very similar in nature: like dissolves
like. Thus, water is miscible with other liquids that can engage in
hydrogen bonding, whereas a hydrocarbon liquid in which London or
dispersion forces are the only significant intermolecular effect will
only be completely miscible with similar kinds of liquids.
Substances such as the alcohols, CH3(CH2)nOH, which have hydrogen
bonding (and thus are hydrophilic) at one end and hydrophobic at the
other, tend to be at least partially miscible with both kinds of solvents.
If n is large, the hydrocarbon properties dominate and the alcohol has
only a limited solubility in water. Very small values of n will allow
the OH group to dominate, so miscibility in water increases and
becomes unlimited in ethanol (n = 1) and methanol (n = 0), but
miscibility with hydrocarbons decreases owing to the energy required
to break alcohol-alcohol hydrogen bonds when the non-polar liquid is
added.
Polar and non-polar molecules
Polar molecules are those in which electric charge is distributed
asymmetrically. The most familiar example is ordinary water, in
which the highly electronegative oxygen atom pulls part of the electric
charge cloud associated with each OH
bond closer to itself. Although the
H2O molecule is electrically neutral
overall, this charge imbalance gives
rise to a permanent electric dipole
moment.
227
"Associated" liquids
Chemists use this term to refer to liquids in which the effects of
hydrogen bonding dominate the local structure.
Water is the most important of these, but
ammonia NH3 and hydrogen cyanide, HCN,
are other common examples.
Thus, liquid water consists of an extended network of H2O molecules
linked together by dipole-dipole
attractions that we call hydrogen bonds.
Because these are much weaker than
ordinary chemical bonds, they are
continually being disrupted by thermal
forces. As a result, the extended
structure is highly disordered (in contrast
to that of solid ice) and continually changing.

When a solute molecule is introduced into an associated liquid, a


certain amount of energy must be expended in order to break the local
hydrogen bond structure and make space for the new molecule. If the
solute is itself an ion or a polar molecule, new ion-dipole or dipole-
dipole attractions come into play. In favourable cases, these may
release sufficient potential energy to largely compensate for the
energy required to incorporate the solute into the structure.
An extreme example of this occurs when ammonia dissolves in water.
Each NH3 molecule can form three hydrogen
bonds, so the resulting solution is even more
hydrogen bonded than is pure water,
accounting for the considerable amount of heat
released in the process and the extraordinarily
large solubility of ammonia in water.

Non-polar solutes are sparingly soluble in water


When a non-polar solute such as oxygen or hexane is introduced into
an associated liquid, we might expect that the energy required to break
the hydrogen bonds to make space for the new molecule is not
compensated by the formation of new attractive interactions,
suggesting that the process will be energetically unfavourable. We

228
can therefore predict that solutes of these kinds will be only sparingly
soluble in water.

12.4 Solubility of Gases

Have you ever noticed the tiny bubbles that form near the bottom of a
container of water when it is placed on a hot stove? These bubbles
contain air that was previously dissolved in the water, but reaches its
solubility limit as the water is warmed. You can completely rid a
liquid of any dissolved gases (including unwanted ones such as Cl 2 or
H2S) by boiling it in an open container.

Solubility of oxygen in water

Fresh water at sea level dissolves 14.6 mg of oxygen per litre at 0C


and 8.2 mg/L at 25C. These saturation levels ensure that fish and
other gilled aquatic animals are able to extract sufficient oxygen to
meet their respiratory needs. In actual aquatic environments however,
the presence of decaying organic matter or nitrogenous runoff can
reduce these levels far below saturation. The health and survival of
these organisms is severely curtailed when oxygen concentrations fall
to around 5 mg/L.
The temperature dependence of the solubility of oxygen in water is an
important consideration for the well-being of aquatic life; thermal
pollution of natural waters (due to the influx of cooling water from
power plants) has been known to reduce the dissolved oxygen
concentration to levels low enough to kill fish. The advent of summer
temperatures in a river can have the same effect if the oxygen
concentration has already been partially depleted by reaction with
organic pollutants.

Solubility of gases increases with pressure: Henry's Law

Solubility of a gas in a liquid increases as the partial pressure of the


gas over the liquid increases. The direct proportionality of gas
solubility to pressure was discovered by William Henry (1775-1836)
and is known as Henry's Law. It is usually written as
S = kH P

229
where, S = solubility, P = partial pressure of the gas above the
solution, and kH = Henrys constant, which depends on the gas, the
solvent and the temperature.
This direct proportionality is shown in figure 12.2 below.
O2
N2
Molar solubility

1.0
He
mol/L

0.5

0.5 1.0
Partial pressure (atm)

Figure 12.2: Henrys law graphs for some gases


Henrys law constant can be expressed in various units, and in some
instances is defined in different ways, so be very careful to note these
units when using published values.
In Table 12.1 below, kH is given in units of mol/ (L atm) as
(concentration in liquid, mol L1) / (partial pressure of gas in atm).
Table 12.1: Henrys law constants

Gas: He N2 O2 CO2 NH3


kH :
3.70104 6.10104 1.30103 3.40102 1.75102
(mol/ L atm)

Example 12.10:
If the partial pressure of oxygen gas over a sample of water is 2.0 atm,
what would be the solubility of the gas in water?
Solution:
Solving Henry's law for the concentration, we get
S= kH P = 1.30 103 mol/(L atm) 2.0 atm = 0.0026 mol L1

230
12.5 Solubility of Solids

Solutions of molecular solids in liquids


The stronger intermolecular forces in solids require more input of
energy in order to disperse the molecular units into a liquid solution.
However, there is also a considerable increase in entropy that can
more than compensate if the intermolecular forces, that must be
broken in order to introduce the solute into the liquid, are not too
strong. For example, at 25 C and 1 atm pressure, 20 g of iodine
crystals will dissolve in 100 ml of ethyl alcohol, but the same quantity
of water will dissolve only 0.30 g of iodine, showing the stronger
intermolecular forces in water as compared to ethanol.
As the molecular weight of the solid increases, the intermolecular
forces holding the solid together also increase, and solubility tends to
fall off. Thus, the solid linear hydrocarbons CH3(CH2)nCH3 (n > 20)
show diminishing solubility in hydrocarbon liquids.

Solutions of ionic solids in liquids

Since the coulombic forces that bind ions and highly polar molecules
into solids are quite strong, we might expect these solids to be
insoluble in just about any solvent. Ionic solids are insoluble in most
non-aqueous solvents, but the high solubility of some (including
NaCl) in water suggests the need for some further explanation.
The key factor here turns out to be the interaction of the ions with the
solvent. The electrically-charged ions exert a strong coulombic
attraction on the end of the water molecule that has the opposite
partial charge. As a consequence, ions in solution are always
hydrated; that is, they are quite tightly bound to water molecules
through ion-dipole interaction. The number of water molecules
contained in the primary hydration shell varies with the radius and
charge of the ion.

231
Figure 12.3 shows the hydration shells around some ions in a sodium
chloride solution. The average time an ion spends in a shell is about
2-4 nanoseconds; this is about two orders of magnitude longer than
the lifetime of an individual H2OH2O hydrogen bond.

Figure 12.3: Dissolution of NaCl in water; hydrated Na+ and Cl ions

Lattice and hydration energies


Lattice energy is defined as the energy required to completely
separate ions of one mole of a solid ionic compound and changes them
to gaseous ions. Since energy is required to separate and vaporize
ions, lattice energy is always positive.
Hydration energy is the energy released as a result of hydration of
one mole of ions. This quantity is negative, since water molecules are
attached to the charged ions.
The dissolution of an ionic solid MX in water can be thought of as a
sequence of two steps:

232
Gaseous ions

Lattice Energy

Hydration Energy
Ionic solid

Heat of solution

Hydrated ions

1. MX(s) M+(g) + X(g) H > 0 (lattice energy)


2. M+(g) + X(g) + H2O(l) M+(aq) + X(aq)
H < 0 (hydration energy)
The heat (enthalpy) of solution is the sum of the lattice and hydration
energies, and can have either sign, depending on the relative
magnitudes of lattice and hydration enthalpies.
Table. 12.2: Hydration and Lattice dissociation energies

Hydration energy (kJ mol1) Lattice dissociation energy (kJ mol1)


H+(g) 1075 F(g) 503 F Cl Br I
Li+(g) 515 Cl(g) 369 Li+ +1031 +848 +803 +759
Na+(g) 405 Br(g) 336 Na+ +918 +780 +742 +705
K+(g) 321 I(g) 398 K+ +817 +711 +679 +651
Mg2+(g) 1922 OH(g) 460 Mg2+ +2957 +2526 +2440 +2327
Ca2+(g) 1592 NO3 328 Ca2+ +2630 +2258 +2176 +2074
Sr2+(g) 1445 SO42 1145 Sr2+ +2492 +2156 +2075 +1963
Single-ion hydration energies (as shown in Table 12.2) cannot be
observed directly, but are obtained from the differences in hydration
energies of salts (which are measurable; see below) having the given
ion in common.
When you encounter tables such as above in which numeric values are
related to different elements, you should always stop and see if you
can make sense of any obvious trends. In this case, the things to look
for are the size and charge of the ions as they would affect the

233
electrostatic interaction between two ions, or between an ion and a
(polar) water molecule.

Example 12.11:
When calcium chloride, CaCl2, is dissolved in water, will the temperature
immediately after mixing rise or fall?
Solution:
Estimate the heat of solution of CaCl2.
Lattice energy of solid CaCl2: +2258 kJ mol1
Hydration energy of the three gaseous ions is
(1562 381 381) = 2324 kJ mol1
Heat of solution: (2258 2324) kJ mol1 = 66 kJ mol1
Since the process is exothermic, this amount of heat will be released,
warming the solution.

Lattice energies are not measured directly, but are estimates based on
electrostatic calculations which are reliable only for simple salts.
Enthalpies of solution are observable either directly or, for sparingly
soluble salts, indirectly. Hydration energies are not measurable; they
are estimated as the sum the other two quantities.
It follows that any uncertainty in the lattice energies is reflected in
those of the hydration energies. For this reason, tabulated values of
the latter will vary depending on the source.
As often happens for a quantity that is the sum of two large terms
having opposite signs, the overall dissolution process can come out as
either endothermic or exothermic, and examples of both kinds are
common see Table 12.3.
Table. 12.3: Energy terms of some salts (kJ mol1)

Substance LiF NaI KBr CsI LiCl NaCl KCl AgCl


Lattice energy 1021 682 669 586 846 778 707 910
Hydration
1017 686 649 552 884 774 690 844
energy
Enthalpy of
+3 4 +20 +34 38 +4 +17 +66
solution

234
12.6 Colligative Properties

We are accustomed to describing a solution in terms of the


concentration of one or more solutes. However, many of the
important physical properties of a solution depend more directly on
the concentration of the solvent. These properties include the vapour
pressure, the freezing point, the boiling point, and the osmotic
pressure. Because they are "tied together" (Latin, co ligare) in this
way, they are referred to as the colligative properties of solutions.
The colligative properties that we will consider apply to solutions in
which the solute is non-volatile; that is, it does not make a significant
contribution to the overall vapour pressure of the solution. Solutions
of salt or sugar in water fulfil this condition exactly. Other solutes
that have very small vapour pressures, such as iodine or ethylene
glycol antifreeze, can often be considered non-volatile in comparison
to the solvent at the same temperature.

Vapour pressure lowering

The number 55.5 mol L1 (= 1000 g L1 18 g mol1) is a useful one


to remember if you are dealing a lot with aqueous solutions; this
represents the concentration of water in pure water. (Strictly
speaking, this is the molal concentration of H2O; it is only the molar
concentration at temperatures around 4 C, where the density of water
is closest to 1.000 g cm1.)
The container on the left in figure 12.4 represents pure water whose
concentration in the liquid is 55.5 M. A tiny fraction of the H 2O
molecules will escape into the vapour space, and if the top of the
container is closed, the pressure of water vapour builds up until
equilibrium is achieved. Once this happens, water molecules continue
to pass between the liquid and vapour in both directions, but at equal
rates, so the partial pressure of H2O in the vapour remains constant at
a value known as the vapour pressure of water at the particular
temperature.

235
Figure 12.4: Vapour pressure of solutions: Raoult's law
.
In the container on the right, we have replaced a fraction of the water
molecules with a substance that has zero or negligible vapour pressure
a non-volatile solute such as salt or sugar. This has the effect of
diluting the water, reducing its tendency to escape into the vapour
phase and thus its vapour pressure.
What is important to remember is that the reduction in the vapour
pressure of a solution of this kind is directly proportional to the mole
fraction of the solvent. The reduced vapour pressure is given by
Raoult's law (1886): P = X Po, where P is the vapour pressure of the
solution, X is the mole fraction of the solvent and P o is the vapour
pressure of the pure solvent.

Figure 12.5:
Raoults Law

Example 12.12:
Estimate the vapour pressure of a 40 percent (w/w) solution of ordinary
cane sugar (C22O11H22, 342 g mol1) in water. The vapour pressure of
pure water at this particular temperature is 26.0 torr.

236
Solution:
100 g of solution contains (40 g) (342 g mol 1) = 0.12 mol of sugar and
(60 g) (18 g mol1) = 3.3 mol of water.
The mole fraction of water in the solution is 3.3/(3.3 + 0.12) = 0.96, and
its vapour pressure will be 0.96 26.0 torr = 25 torr.

Since the sum of all mole fractions in a mixture must be unity, it


follows that the more moles of solute, the smaller will be the mole
fraction of the solvent. If the solute is a salt that dissociates into ions,
then the proportion of solvent molecules will be even smaller.

Example 12.13:
The vapour pressure of water at 10C is 9.2 torr. Estimate the vapour
pressure at this temperature of a solution prepared by dissolving 0.200
mole of CaCl2 in 1.00 L of water.
Solution:
Each mole of CaCl2 dissociates into one mole of Ca2+ and two moles of
Cl1, giving a total of 0.600 of solute particles. The mole fraction of water
in the solution will be
55.5/(0.600 + 55.5) = 0.989
The vapour pressure will be 0.989 9.2 torr = 9.1 torr.
This is an approximate value, since the effective concentration of the ions
will be less than 0.600 because of interaction between the ions in the
solution.

Boiling point elevation


When a liquid is heated its vapour pressure rises. When the vapour
pressure equals the outside (atmospheric) pressure, the liquid starts to
boil. If addition of a non-volatile solute lowers the vapour pressure of
the solution, then it follows that the temperature must be raised to
restore the vapour pressure to the value corresponding to the pure
solvent. In particular, the temperature at which the vapour pressure is
1 atm will be higher than the normal boiling point by an amount
known as the boiling point elevation.
The exact relation between the boiling point of the solution and the
mole fraction of the solvent is rather complicated, but for dilute

237
solutions the elevation of the boiling point is directly proportional to
the molal concentration of the solute:
Tb = Kb m
where Tb is the boiling point elevation, m is the molality (mol solute
/ mass solvent in kg), and Kb is a constant which depends on the
nature of the solvent - it has the units K kg mol-1.
Bear in mind that the proportionality constant Kb is a property of the
solvent, because this is the only component that contributes to the
vapour pressure in the model we are considering in this section see
Table 12.4.
Table.12.4: Boiling point elevation constants

Solvent Normal bp (C) Kb (K kg mol1)


water 100 0.514
ethanol 79 1.19
acetic acid 118 2.93
carbon tetrachloride 76.5 5.03

Example 12.14:
Sucrose (C22O11H22, 342 g mol1), like many sugars, is highly soluble in
water; almost 2000 g will dissolve in 1 kg of water, giving rise to what
amounts to pancake syrup. Estimate the boiling point of such a sugar
solution.
Solution:
No. of moles of sucrose: (2000 g) / (342 g mol1) = 5.85 mol
Mass of water: assume 1000 g
The molality of the solution is (5.85 mol) (1.00 kg) = 5.85 m
Using the value of Kb from the table, the boiling point will be raised by
(0.514 K kg mol1) (5.85 mol kg1) = 3.01 K,
so the boiling point will be 103 C

Freezing point depression


The freezing point of a substance is the temperature at which the solid
and liquid forms can coexist indefinitely that is, they are in

238
equilibrium. Under these conditions, molecules pass between the two
phases at equal rates because their escaping tendencies from the two
phases are identical.
Suppose that a liquid solvent and its solid (water and ice, for example)
are in equilibrium, and we add a non-volatile solute (such as salt,
sugar, or automotive antifreeze liquid) to the water. This will have the
effect of reducing the mole fraction of H2O molecules in the liquid
phase, and thus reduce the tendency of these molecules to escape from
it, not only into the vapour phase (as we saw above), but also into the
solid (ice) phase. This will have no effect on the rate at which H2O
molecules escape from the ice into the water phase, so the system will
no longer be in equilibrium and the ice will begin to melt.
If we wish to keep the solid from melting, the escaping tendency of
molecules from the solid must be reduced. This can be accomplished
by reducing the temperature; this lowers the escaping tendency of
molecules from both phases, but it affects those in the solid more than
those in the liquid, so we eventually reach the new, lower freezing
point where the two quantities are again in exact balance and both
phases can coexist.

If you prefer to think in terms of vapour pressures, you can use the
same argument if you bear in mind that the vapour pressures of the
solid and liquid must be the same at the freezing point. Dilution of the
liquid (the solvent) by the non-volatile solute reduces the vapour
pressure of the solvent according to Raoults law, thus reducing the
temperature at which the vapour pressures of the liquid and frozen
forms of the solution will be equal.

As with boiling point elevation, in dilute solutions there is a simple


linear relation between the freezing point depression and the molality
of the solute:
Tf = Kf m
Table 12.5 shows values for Kf for some solvents.

239
Table. 12. 5: Freezing point depression constants

Solvent Normal fp (C) Kf (K kg mol1)


water 0.0 1.86
acetic acid 16.7 3.90
benzene 5.5 5.10
camphor 180 40.0
cyclohexane 6.5 20.2
phenol 40 7.3

The use of salt to de-ice roads is a common application of this


principle. The solution formed when some of the salt dissolves in the
moist ice reduces the freezing point of the ice. If the freezing point
falls below the ambient temperature, the ice melts. In very cold
weather, the ambient temperature may be below that of the salt
solution, and the salt will have no effect.
The effectiveness of a de-icing salt depends on the number of particles
it releases on dissociation and on its solubility in water:

Name Formula Lowest practical T (C)


Ammonium sulphate (NH4)2SO4 7
Calcium chloride CaCl2 29
Potassium chloride KCl 15
Sodium chloride NaCl 9
Urea (NH2)2CO 7

Automotive radiator antifreezes are mostly based on ethylene glycol,


(CH2OH)2. Owing to the strong hydrogen bonding properties of this
double alcohol, the substance is miscible with water in all proportions,
and contributes only a very small vapour pressure of its own. Besides
lowering the freezing point, antifreeze also raises the boiling point,
increasing the operating range of the cooling system. The pure glycol
freezes at 12.9C and boils at 197C, allowing water-glycol mixtures
to be tailored to a wide range of conditions.
Example 12.15:
Estimate the freezing point of an antifreeze mixture that is made up by

240
combining one volume of ethylene glycol (MW = 62, density 1.11 g cm3)
with two volumes of water.
Solution:
Assume that we use 1 L of glycol and 2 L of water (the actual volumes do
not matter as long as their ratios are as given.) The mass of the glycol will
be 1.10 kg and that of the water will be 2.0 kg, so the total mass of the
solution is 3.11 kg. We then have:
Number of moles of glycol = (1110 g) (62 g mol1) = 17.9 mol
Molality of glycol = (17.9 mol) (2.00 kg) = 8.95 mol kg1
Freezing point depression is
Tf = (1.86 K kg mol1) (8.95 mol kg1) = 16.6 K
so the solution will freeze at 16.6 C.

Any ionic species formed by dissociation will also contribute to the


freezing point depression. This can serve as a useful means of
determining the fraction of a solute that is dissociated.
Example 12.16:
An aqueous solution of nitrous acid (HNO2, MW = 47) freezes at
0.1980C. If the solution was prepared by adding 0.100 mol of the acid to
1000 g of water, what percentage of the HNO2 is dissociated in the
solution?
Solution:
The nominal molality of the solution is
(0.100 mol) (1.00 kg) = 0.100 mol kg1.
But the effective molality according to the observed Tf value is given by
m = Tf Kf = (0.198 K) (1.86 K kg mol1) = 0.106 mol kg1.
This is the total number of moles of species present after the dissociation
reaction HNO2 H+ + NO2 has occurred.
If we let x = [H+] = [NO2], then by stoichiometry, [HNO2] = (0.100 x).
Then, (0.100 x) + 2 x = 0.106 x = 0.006
The fraction of HNO2 that is dissociated is 0.006 0.100 = 0.06,
corresponding to 6% dissociation of the acid.

The molar mass of a substance can be determined from freezing point


depression. The technique is called cryoscopy.
241
Example 12.17:
When 350 mg of an organic compound. was added to 100.0 g of camphor,
it lowered the freezing point of camphor by 0.72 oC. Calculate the molar
mass of the compound. (Kf =40.0 K.kg.mol-1)
Solution:
First, arrange the equation for freezing point depression to find solute
molality. Then find the number of moles of solute in the sample (multiply
molality by mass of the solvent). Finally, find the molar mass (mass
/mol).
Molality m = Tf / Kf = 0.72 K / 40.0 K kg mol-1
Moles of solute is
= (0.100 kg) (0.72 K / 40.0 K kg mol1) = (0.100) (0.72 / 40.0) mol
Since number of moles = mass / MM; MM = mass / number of moles
MM of the compound
= 0.350 g 40.0K kg mol-1 / (0.100 kg 0.72 mol) = 1.9 102 g / mol

Another view of f.p. depression and b.p. elevation


A simple phase diagram (see figure 12.6) can provide more insight
into these phenomena.

Figure 12.6:
Example of a phase
diagram

Figure 12.7 below expands on this by plotting lines for both pure
water and for its "diluted" state produced by the introduction of a non-
volatile solute.

242
Figure 12.7: Phase diagram of a pure solvent and its solution

The normal boiling point of the pure solvent is indicated by point


where the vapour pressure curve intersects the 1 atm line that is,
where the escaping tendency of solvent molecules from the liquid is
equivalent to 1 atmosphere pressure. Addition of a non-volatile solute
reduces the vapour pressures, shifting the boiling point to the right ,
corresponding to the increase in temperature Tb required to raise the
escaping tendency of the H2O molecules back up to 1 atm.

To understand freezing point depression, notice that the vapour


pressure line intersects the curved vapour pressure line of the solid
(ice) at , which corresponds to a new triple point at which all
three phases (ice, water vapour, and liquid water) are in equilibrium
and thus exhibit equal escaping tendencies. This point is by definition
the origin of the freezing (solid-liquid) line, which intersects the 1 atm
line at a reduced freezing point Tf, indicated by .
Note that the above analysis assumes that the solute is soluble only in
the liquid solvent, but not in its solid form. This is generally more or
less true. For example, when arctic ice forms from seawater, the salts
mostly get "squeezed" out. This has the interesting effect of making
the water that remains more saline, and hence denser, causing it to

243
sink to the bottom part of the ocean where it gets taken up by the
south-flowing deep current.

12.7 Osmotic Pressure and Osmosis

Osmotic pressure is the fourth member of the quartet of colligative


properties that arise from the dilution of a solvent by non-volatile
solutes. Because of its great importance, we are devoting a separate
section to this topic with special emphasis on some of its many
practical applications.

Semi permeable membranes and osmotic flow


Osmosis is the process in which a liquid passes
through a membrane whose pores permit the
passage of solvent molecules but are too small for
the larger solute molecules to pass through see
figure 12.8 on the right.

Figure 12.8: Osmosis

In an osmotic cell (figure 12.9), both compartments contain water, but


the one on the left also contains a solute whose molecules (represented
by black circles) are too large to pass
through the membrane. Many
artificial and natural substances are
capable of acting as semi-permeable
membranes. The walls of most plant
and animal cells fall into this category.
Figure 12.9: Osmotic cell

If the cell is set up so that the liquid level is initially the same in both
compartments, you will soon notice that the liquid rises in the left
compartment and falls in the right side, indicating that water
molecules from the right compartment are migrating through the semi-

244
permeable membrane and into the left compartment. This migration
of the solvent is known as osmotic flow, or simply osmosis.
What is the force that drives the molecules through the membrane?
This is a misleading question, because there is no real force in the
physical sense other than the thermal energies all molecules possess.
Osmosis is a consequence of simple statistics: the randomly directed
motions of a collection of molecules will cause more to leave a region
of high concentration than return to it. The escaping tendency of a
substance from a phase increases with its concentration in the phase.

Osmotic equilibrium and osmotic pressure

One way to stop osmosis is to raise the hydrostatic pressure on the


solution side of the membrane. This pressure squeezes the solvent
molecules closer together, raising their escaping tendency from the
phase. If we apply enough pressure (or let the pressure build up by
osmotic flow of liquid into an enclosed region), the escaping tendency
of solvent molecules from the solution will eventually rise to that of
the molecules in the pure solvent, and osmotic flow will cease. The
pressure required to achieve osmotic equilibrium is known as the
osmotic pressure. Note that the osmotic pressure is the pressure
required to stop osmosis, not to sustain it.
Caution! It is common usage to say that a solution has an osmotic
pressure of "x atmospheres". It is important to understand that this
means nothing more than that a pressure of this value must be applied
to the solution in order to prevent flow of pure solvent into this
solution through a semi permeable membrane separating the two
liquids.

Osmotic pressure and solute.concentration


The Dutch scientist Jacobus vant Hoff (1852-1911) was one of the
giants of physical chemistry. He discovered this equation after a
chance encounter with a botanist friend during a walk in a park in
Amsterdam. The botanist had learned that the osmotic pressure
increases by about 1/273 for each degree of temperature increase.
Vant Hoff immediately grasped the analogy to the ideal gas law.

245
The osmotic pressure of a solution containing n moles of solute
particles in a solution of volume V is given by the van't Hoff equation:
= nRT/V
in which R is the gas constant (0.0821 L atm mol1 K1) and T is the
absolute temperature. The fraction n/V corresponds to the molarity of
a solution of a non-dissociating solute, or to twice the molarity of a
totally-dissociated solute such as NaCl. In this context, molarity
refers to the summed total of the concentrations of all solute species.
Note therefore, that in contrast to the need to employ solute molality
to calculate the effects of a non-volatile solute on changes in the
freezing and boiling points of a solution, we can use solute molarity to
calculate osmotic pressures.

Recalling that is the Greek equivalent of P, the re-arranged form


V = nRT of the above equation should look familiar. Much effort
was expended around the end of the 19th century to explain the
similarity between this relation and the ideal gas law, but in fact, the
vant Hoff equation turns out to be only a very rough approximation
of the real osmotic pressure law, which is considerably more
complicated and was derived after van't Hoff's formulation. As such,
this equation gives valid results only for extremely dilute ("ideal")
solutions. According to the van't Hoff equation, an ideal solution
containing 1 mole of dissolved particles per litre of solvent at 0C will
have an osmotic pressure of 22.4 atm.
Example 12.18:
Sea water contains dissolved salts at a total ionic concentration of about
1.13 mol L1. What pressure must be applied to prevent osmotic flow of
pure water into sea water through a membrane permeable only to water
molecules?
Solution:
= MRT = 1.13 mol L1 0.0821 L atm mol1 K1 298 K = 27.6 atm

246
Some practical applications of osmosis
Molecular weight determination by osmotic pressure

Since all of the colligative properties of solutions depend on the


concentration of the solvent, their measurement can serve as a
convenient experimental tool for determining the concentration, and
thus the molecular weight, of a solute.
Osmotic pressure is especially useful in this regard, because a small
amount of solute will produce a much larger change in this quantity
than in the boiling point, freezing point, or vapour pressure. Even a
106 molar solution would have a measurable osmotic pressure.
Molecular weight determinations are very frequently made on proteins
or other high molecular weight polymers. These substances, owing to
their large molecular size, tend to be only sparingly soluble in most
solvents, so measurement of osmotic pressure is often the only
practical way of determining their molecular weights.
Example 12.19:
The osmotic pressure of a benzene solution containing 5.0 g of
polystyrene per litre was found to be 7.6 torr at 25C. Estimate the
average molecular weight of the polystyrene in this sample.
Solution:
Osmotic pressure: = (7.6 torr) / (760 torr atm1) = 0.0100 atm
Using the form of the van't Hoff equation V = nRT, the number of
moles of polystyrene, n, is
(0.0100 atm 1 L) / (0.0821 L atm mol1 K1 298 K)
= 4.09 104 mol
Molar mass of the polystyrene is
(5.0 g) (4.09 x 104 mol) = 1.2 x 104 g mol1.

The experiment is quite simple: pure solvent is introduced into one


side of a cell that is separated into two parts by a semi permeable
membrane. The polymer solution is placed in the other side, which is
enclosed and connected to a manometer or some other kind of
pressure gauge. As solvent molecules diffuse into the solution cell the
pressure builds up; eventually this pressure matches the osmotic
pressure of the solution and the system is in osmotic equilibrium. The

247
osmotic pressure is read from the measuring device and substituted
into the vant Hoff equation to find the number of moles of solute.
Reverse osmosis

If it takes a pressure of atm to bring about osmotic equilibrium, then


it follows that applying a hydrostatic pressure greater than this to the
high-solute side of an osmotic cell will force water to flow back into
the fresh-water side. (See figure 12.10.)

Fig. 12.10: Reverse osmosis

This process, known as reverse osmosis, is now the major technology


employed to desalinate ocean water and to reclaim "used" water from
power plants, runoff, and even from sewage. It is also widely used to
deionize ordinary water and to purify it for industrial uses (especially
beverage and food manufacture) and drinking purposes.

Figure 12.11: Reverse osmosis

Pre-treatment commonly employs activated-carbon filtration to


remove organics and chlorine (which tends to damage RO
membranes). Although bacteria are unable to pass through semi
permeable membranes, the latter can develop pinhole leaks, so some
form of disinfection is often advised.

248
Osmotic generation of electric power
The osmotic pressure of seawater is almost 26 atm. Since a pressure
of 1 atm will support a column of water 10.6 m high, this means that
osmotic flow of fresh water through a semi permeable membrane into
seawater could in principle support a column of the latter by
26 x 10.3 = 276 m!
So imagine an osmotic cell in which one side is supplied with fresh
water from a river, and the other side with seawater. Osmotic flow of
fresh water into the seawater side forces the latter up through a riser
containing a turbine connected to a generator, thus providing a
constant and fuel-less source of electricity.

Figure 12.12: Osmotic


generation of electric power

The key component of such a scheme, first proposed by an Israeli


scientist in 1973 and known as pressure-retarded osmosis (PRO) is of
course a semi permeable membrane capable of passing water at a
sufficiently high rate.
A plant based on this principle was opened in 2009 in Norway. Its
capacity is only 4 kW, but it serves as proof-in-principle of a scheme
that is estimated capable of supplying up to 2000 terawatt-hours of
energy worldwide. The semi permeable membrane operates at a
pressure of about 10 atm and passes 10 L of water per second,
generating about 1 watt per m2 of membrane
PRO is but one form of salinity gradient power that depends on the
difference between the salt concentrations in different bodies of water.

249
Osmosis in biology and physiology
Because many plant and animal cell membranes and tissues tend to be
permeable to water and other small molecules, osmotic flow plays an
essential role in many physiological processes.
(i) Normal saline solution
The interiors of cells contain salts and other solutes that dilute the
intracellular water. If the cell membrane is permeable to water,
placing the cell in contact with pure water will draw water into the
cell, tending to rupture it. This is easily and dramatically seen if red
blood cells are placed in a drop of water and observed through a
microscope as they burst. This is the reason that "normal saline
solution", rather than pure water, is administered in order to maintain
blood volume or to infuse therapeutic agents during medical
procedures. In order to prevent irritation of sensitive membranes, one
should always add some salt to water used to irrigate the eyes, nose,
throat or bowel. Normal saline contains 0.91% w/v of sodium
chloride, corresponding to 0.154 M, making its osmotic pressure close
to that of blood.

(ii) Food preservation

The drying of fruit, the use of sugar to preserve jams and jellies, and
the use of salt to preserve certain meats, are age-old methods of
preserving food. The idea is to reduce the water concentration to a
level below that in living organisms. Any bacterial cell that wanders
into such a medium will have water osmotically drawn out of it, and
will die of dehydration. A similar effect is noticed by anyone who
holds a hard sugar candy against the inner wall of the mouth for an
extended time; the affected surface becomes dehydrated and
noticeably rough when touched by the tongue.

(iii) Water transport in plants: osmosis pushes, hydrogen-


bonding pulls

Osmotic flow plays an important role in the transport of water from its
source in the soil to its release by transpiration from the leaves; it is
helped along by hydrogen-bonding forces between the water
molecules. Capillary rise is not believed to be a significant factor.
250
Colligative properties of electrolyte solutions
Colligative properties of solutions are determined by the number of
solute particles in the solution. Since electrolytes dissociate in
solutions, the actual number of particles in solutions is not the same as
the number of units dissolving. Thus, an amount of solvent in which
0.010 mol NaCl is dissolved will have 0.010 mol Na+ and 0.010 mol
Cl. Assuming complete dissociation, there would be 0.020 moles
ions in the solution. A 0.010 molal BaCl2 solution would have an
effective molality of 3 x 0.010 = 0.030 molal. The equations for the
colligative properties have to be modified to take this into account.
The modified equation for the depression of freezing point becomes:
Tf = i Kf m.
The factor i is called the vant Hoff factor. It is the ratio of the actual
number of particles in solution to the number of units initially
dissolved. Since even in dilute solutions, ions of opposite charges do
pair up to certain extent, the vant Hoff factor is usually less than what
the formula of the solute indicates. Thus in a 0.050 molal NaCl
solution, experiment shows that i = 1.9 and not 2.
12.8 Exercises
12.1 Calculate the mass percent, molality and mole fraction of the solute
in an aqueous solution containing 0.468 g Ca(NO3)2 in 270 g water.

12.2 What is the molarity of an aqueous solution containing 7.804 g


Na2SO4.10H2O in 385.0 mL of the solution?

12.3 What mass of KMnO4 is required to prepare a 350 mL solution


whose concentration is 1.37 M?

12.4 An aqueous solution is 0.340 molal NaCl. The density of the


solution is 1.015 g cm-3. What is the molarity of the solution?

12.5 Calculate the vapour pressure of water above a solution which has
17.4 g of urea, CO(NH2)2 in 250.0 g of water. The vapour pressure
of pure water is 23.76 torr.

12.6 Calculate the freezing point and boiling point of a solution that has
3.14 g NaCl in 240 g of water.

251
12.7 To 200 g of water 3.18 g fructose and 2.13 g KI are added. What is
the freezing point and boiling point of the resulting solution?
12.8 A solution of an unknown carbohydrate containing 5.40 g in 100.0
mL of solution has an osmotic pressure of 11.0 atm at 25 oC. What
is the molar mass of the carbohydrate?

12.9 The osmotic pressure of a 0.020 M solution of Na2SO4 is 1.08 atm


at 25oC. What is the vant Hoff factor, i, for the solution? What
happens to the value of i as the solution becomes more concentrated
and why?

252
CHAPTER 13
The Rates of Chemical Reactions
13.1 Reaction Rate

Chemical reactions vary greatly in the speed at which they occur.


Some are essentially instantaneous, while others may take years to
reach equilibrium. Before dealing with rates of reactions
quantitatively, let us consider factors that speed up reactions.

Factors affecting rates of reactions


Rates of reaction are influenced by the following factors:
1. Nature of reactants: Some reactants inherently react among
themselves faster than others.
2. Concentration of reactants: Generally, reactions are faster at
higher concentrations.
3. Temperature: Most reactions speed up as the temperature is
increased.
4. Large surface area or small particle size in the case of solid
reactants: If the area of contact between reactants is large,
reactions go faster other things being equal.
5. Action of catalysts: Catalysts speed up chemical reactions
without themselves being used up.

Definition of reaction rate

The speed or rate of a chemical reaction may be defined as the change


in concentration of a substance divided by the time interval during
which this change is observed:
(concentration )
rate
( time )

For a reaction of the form A + B C, the rate can be expressed in


terms of the change in concentration of any of its components:

253
[A] [B] [C]
rate rate rate
t t t

in which [A] is the difference between the concentration of A over


the time interval t2 t1: [A] = [A]2 [A]1
Notice the minus signs in the first two examples above. The
concentration of a reactant always decreases with time, so [A] and
[B] are both negative. Since negative rates don't make much sense,
rates expressed in terms of a reactant concentration are always
preceded by a minus sign in order to make the rate come out positive.

Consider now a reaction in which the coefficients are different:


A + 3B 2D
It is clear that [B] decreases three times as rapidly as [A], so in order
to avoid ambiguity when expressing the rate in terms of different
components, it is customary to divide each change in concentration by
the appropriate coefficient:
[A] [B] [D]
rate
t 3t 2t

Example 13.1:
For the oxidation of ammonia 4NH3 + 3O2 2N2 + 6H2O it was found
that the rate of formation of N2 was 0.27 mol L1 s1.
(a) At what rate was water being formed?
(b) At what rate was ammonia being consumed?
Solution:
6
(a) From the equation stoichiometry, [H2O] = [N2], so the rate of
2
formation of H2O is 3 (0.27 mol L1 s1) = 0.81 mol L1 s1.
(b) 4 moles of NH3 are consumed for every 2 moles of N2 formed, so the
rate of disappearance of ammonia is
2 (0.27 mol L1 s1) = 0.54 mol L1 s1.
Because of the way this question is formulated, it would be acceptable
to express this last value as a negative number.

254
Instantaneous rate
Most reactions slow down as the reactants are consumed.
Consequently, the rates given by the expressions shown above tend to
lose their meaning when measured over longer time intervals t.
Thus for the reaction whose progress is plotted here, the actual rate (as
measured by the increasing concentration of product) varies
continuously, being greatest at time zero. The instantaneous rate of a
reaction is given by the slope of a tangent to the concentration -vs.-
time curve. Three such rates have been identified in Figure 13.1.
Note: Instantaneous rates are also known as differential rates.

Figure 13.1: Plot


showing differential
rates

An instantaneous rate taken near the beginning of the reaction (t = 0)


is known as an initial rate (label (1) here). As we shall soon see,
initial rates play an important role in the study of reaction kinetics.
These tangent slopes are derivatives whose values can very at each
point on the curve, so that these instantaneous rates are really limiting
rates defined as:
d[A]
rate
dt

255
13.2 Rate Laws and Reaction Order

The rate law


The relation between the rate of a reaction and the concentrations of
reactants is expressed by its rate law. For example, the rate of the gas
phase decomposition of dinitrogen pentoxide
2N2O5 4NO2 + O2
has been found to be directly proportional to the concentration of
N2O5:
rate = k [N2O5]

The expression for the rate law generally bears no necessary relation
to the reaction equation, and must be determined experimentally.
More generally, for a reaction of the form
nA A + nB B + ... products
the rate law will be
rate = k [A]a [B]b
in which the exponents a and b are usually (but not always) integers
and, we must emphasize once again, bear no relation to the
coefficients nA, nB.

Units of the rate constant


Since the rate of a reaction has the dimensions of (concentration/time),
the dimensions of the rate constant k will depend on the exponents of
the concentration terms in the rate law. To make this work out
properly, if we let p be the sum of the exponents of the concentration
terms in the rate law
p = a + b + ...
then k will have the dimensions (concentration1P/time). If the order
of reaction is n, then, since k = rate/ (concentration)n, units of k = units
of rate / (units of concentration)n.
For a reaction of order 1: units of k = M s-1 / M = s-1 or 1/s.
For a second order reaction: units of k = M s-1 / M2 = M-1 s-1.
Similarly, units of k for other orders can be worked out.
256
Reaction order

The order of a rate law is the sum of the exponents in its concentration
terms. For the N2O5 decomposition with the rate law k[N2O5], this
exponent is 1 (and thus is not explicitly shown); this reaction is
therefore a first order reaction. We can also say that the reaction is
"first order in N2O5".

For more complicated rate laws, we can speak of the overall reaction
order and also the orders with respect to each component. As an
example, consider a reaction
A + 3B + 2C products
with the experimental rate law
rate = k [A] [B]2
We would describe this reaction as third order overall, first order in A,
second order in B, and zero order in C.
Zero order means that the rate is independent of the concentration of
a particular reactant. However, of course enough C must be present to
allow the equilibrium mixture to form.
Example 13.2:
The rate of oxidation of bromide ions by bromate in an acidic aqueous
solution: 6H+ + BrO3 + 5Br 3 Br2 + 3 H2O
is found to follow the rate law: rate = k [Br][BrO3][H+]2
What happens to the rate if, in separate experiments, (a) [BrO3] is
doubled; (b) [H+] is decreased by a factor of 10; (c) the solution is diluted
to twice its volume, but the [H+] is kept constant by use of a buffer?
[Buffers maintain [H+] constant; well study them later].
Solution:
(a) Since the rate is first-order in bromate, doubling its concentration will
double the reaction rate.
(b) Since the reaction is second order in [H+], decreasing [H+] by a factor
of 10 will decrease the rate by a factor of 102 or by 100.
(c) Dilution reduces the concentrations of both Br2 and BrO3 to half their
original values. Doing this to each concentration alone would reduce
the rate by a factor of 2, so reducing both concentration will reduce
the rate by a factor of 4, to () () = of its initial value.

257
13.3 Determining Reaction Orders

In order to determine the value of the exponent in a rate equation term,


we need to see how the rate varies with the concentration of the
substance.
For a single-reactant decomposition reaction of the form
A products
in which the rate is d[A]/dt, we simply plot [A] as a function of time,
draw tangents at various intervals, and see how the slopes of these
tangents (the instantaneous rates) depend on [A]:
If doubling the concentration of A doubles the rate, then the
reaction is first order in A.
If doubling the concentration results in a fourfold rate increase,
the reaction is second order in A.

Initial rate method


In the initial rate method, we measure only the rate near the beginning
of the reaction, before the concentrations have had time to change
significantly. The experiment is then repeated with a different starting
concentration of the reactant in question, but keeping the
concentrations of any others the same. After the order with respect to
one component is found, another series of trials is conducted in which
the order of another component is found.

Figure 13.2.(a): Molarity of


N2O5 vs. time for the reaction
2N2O5 4 NO2 + O2

258
Figure 13.2.(b): Initial
rate vs. concentration of
N2O5 for the reaction
2N2O5 4 NO2 + O2

The graphs in figure 13.2.(a) and (b) show how the initial rates are
determined using the example of the reaction 2N2O5 4 NO2 + O2.
In figure 13.2.(a), concentrations are plotted against time and the
slopes at time = 0 are obtained. In this example, a series of runs using
five different initial concentrations of N2O5 has been made. The
slopes at t = 0 have been measured. These are the initial rates. The
initial rates can then be plotted against initial concentrations to
determine the order (figure 13.2.(b)).
If suitable initial concentrations of reactants are chosen, the order can
be determined without drawing the graph as in the following example.

Example 13.3:
A study of the gas-phase reduction of nitric oxide by hydrogen
2 NO + 2 H2 N2 + 2 H2O
yielded the following initial-rate data (all pressures in torr):
Experiment P(NO) P(H2) Initial rate (torr s1)
1 300 300 1.03
2 150 300 0.25
3 300 600 2.00
Find the order of the reaction with respect to each component.

259
Solution:
In looking over this data, take note of the following:
All the data are expressed in pressures, rather than in concentrations. We
can do this because the reactants are gases, whose concentrations are
directly proportional to their partial pressures when T and V are held
constant. Since we are only interested in comparing the ratios of
pressures and rates, the units cancel out and don't matter. It is far easier
experimentally to adjust and measure pressures than concentrations.
Experiments 1 and 2: Reduction of the initial partial pressure of NO by a
factor of (300/150) = 2, results in a reduction of the initial rate by a factor
of (1.03/0.25) = about 4, so the reaction is second-order in nitric oxide.
Experiments 1 and 3: Increasing the initial partial pressure of hydrogen by
a factor of 2 (600/300), causes a similar increase in the initial rate
(2.00/1.03), so the reaction is first-order in hydrogen.
The rate law is thus rate = k[NO]2[H2].

Dealing with multiple reactants: the isolation method


It is not always practical to determine orders of two or more reactants
by the method illustrated in the preceding example. Fortunately, there
is another way to accomplish the same task: we can use excess
concentrations of all the reactants except the one we wish to
investigate. For example, suppose the reaction is
A + B + C products
and we need to find the order with respect to [B] in the rate law. If we
set [B]o to 0.020 M and let [A]o = [C]o = 2.00M, then if the reaction
goes to completion, the change in [A] and [C] will also be 0.020 M
which is only 1 percent of their original values. This will often be
smaller than the experimental error in determining the rates, so it can
be neglected. By "flooding" the reaction mixture with one or more
reactants, we are effectively isolating the one in which we are
interested.

13.4 Differential and Integral Rate Laws

Now we extend the concept of differential rate laws introduced


previously to integral rate laws and reaction half-life that are of great
importance in most practical applications of kinetics.
260
Measuring instantaneous rates as has been described, is the most
direct way of determining the rate law of a reaction, but is not always
convenient, and it may not even be possible to do so with any
precision.
If the reaction is very fast, its rate may change more rapidly than
the time required to measure it; the reaction may be finished
before even an initial rate can be observed.
In the case of very slow reactions, observable changes in
concentrations occur so slowly that the observation of a truly
"instantaneous" rate becomes impractical.
The ordinary rate law (more precisely known as the instantaneous or
differential rate law) tells us how the rate of a reaction depends on the
concentrations of the reactants. But for many practical purposes, it is
more important to know how the concentrations of reactants (and of
products) change with time.
For example, if you are carrying out a reaction on an industrial scale,
you would want to know how long it will take for, say, 95% of the
reactants to be converted into products.
This is the purpose of an integrated rate law.

Integrated rate laws


How long does it take for a chemical reaction to occur under a given
set of conditions? As with many "simple" questions, no meaningful
answer can be given without being more precise. In this case,
How do we define the point at which the reaction is "completed"?
A reaction is "completed" when it has reached equilibrium that is,
when concentrations of the reactants and products are no longer
changing.
If the equilibrium constant is quite large, then the answer reduces to a
simpler form: the reaction is completed when the concentration of a
reactant falls to zero. In the interest of simplicity, we will assume that
this is the case in the remainder of this discussion.
"How long?" may be too long.
If the reaction takes place very slowly, the time it takes for every last
reactant molecule to disappear may be too long for the answer to be

261
practical. In this case, it might make more sense to define
"completed" when a reactant concentration has fallen to some
arbitrary fraction of its initial value 90%, 70%, or even only 20%.
The particular fraction one selects depends on the cost of the reactants
in relation to the value of the products, balanced against the cost of
operating the process for a longer time or the inconvenience of
waiting for more product. This kind of consideration is especially
important in industrial processes in which the balances of these costs
affect the profitability of the operation.
The half-life of a reaction
Instead of trying to identify the time required for the reaction to
become completed, it is far more practical to specify the time required
for the concentration of a reactant to fall to half of its initial value.
This is known as the half-life (or half-time) of the reaction.
First-order reactions: the law of exponential change
The rate at which a reactant is consumed in a first-order process is
proportional to its concentration at that time. This general
relationship, in which a quantity changes at a rate that depends on its
instantaneous value, is said to follow an exponential law.
Exponential relations are widespread in science and in many other
fields. Consumption of a chemical reactant or the decay of a
radioactive isotope follows the exponential decay law. Its inverse, the
law of exponential growth, describes the manner in which the money
in a continuously-compounding bank account grows with time, or the
population growth of a colony of reproducing organisms.
The reason that the exponential function y = ex so efficiently describes
such changes stems from the remarkable property that dy/dx = ex; that
is, ex is its own derivative, so the rate of change of y is identical to its
value at any point.

Integrating the rate law: first-order reaction


Integrated expressions for rate laws are useful if the concentration
time data are given. The first order differential rate equation for the
reaction A Products, can be integrated to:
ln [A]t ln [A]o = kt
or ln [A]t = kt + ln [A]o

262
Here, [A]o is the starting concentration of A, and [A]t is the
concentration at time t.
This integrated equation is now of the form y = mx + c, representing a
straight line. A plot of ln[A]t against t will be a straight line for a first
order reaction, with slope = k. Thus to test if a reaction is first order,
ln[concentration] values are plotted against time. If the plot is a
straight line, it shows that the reaction is first order.

The integrated equation can be written in its alternative forms:


ln {[A]t / [A]o]} = kt
[A]t = [A]o exp(kt)
The above equations can be used to get any one of the quantities,
concentration at any time, rate constant and the time taken by the
reaction to reach any given concentration provided the others are
given.

Figure 13.3: The integrated rate law for a first-order reaction

Half-life for a first order reaction


After a period of one half-life, t = t and we can write

e kt
[A] 1
[A]o 2
Taking logarithms of both sides (remember that ln ex = x) yields
ln 0.5 = k t1/2
Solving for the half-life, we obtain the simple relation t1/2 = 0.693/k.

263
This tells us that the half-life of a first-order reaction is a constant.
The concentration of a reactant will decrease from 0.80 M to 0.40 M
in the same time as it takes to decrease from 0.10 M to 0.05 M. The
half-life of a first order reaction does not depend on the starting
concentration. Figures 13.4.(a) and (b) show plots of the first order
integrated rate law; figure 13.5 shows the effect of k on t.

(a) (b)
Figure 13.4: Plots of the first order integrated rate law

Figure 13.5: Effect of k on t

It should be clear from figure 13.5


that the rate constant and the half-life
of a first-order process are inversely
related.
The decay of radioactive nuclei
(radioactive decay) is always a first-
order process.

Example 13.4:
The half-life of a first-order reaction was found to be 10 min at a certain
temperature. What is its rate constant in reciprocal seconds?
Solution:
From the equation t1/2 = 0.693/k, k = 0.693/(600 s) = 0.00115 s1

264
Second-order reactions
Integration of the second-order rate law (rate = k[A]2) yields
1/[A]t 1/[A]o = kt
or 1/[A]t = 1/[A]o + kt
A plot of 1/[A]t against time gives a straight line with slope = k.

12
9
1/[A]t 6
(102 M-1) 3
0
0 20 40 60 80
time (s)

Figure 13.6: The integrated rate law for a second-order reaction

Half-life of a second order reaction:


The integrated equation for a second order reaction can be used to get
the expression for the half life. Thus,
1/[A]t 1/[A]o = kt
[A]t = [A]o/2, when t = t1/2
Substitution in the first equation gives
1/[A]o =kt1/2
or t1/2 = 1/k[A]o
Notice that the half-life of a second-order reaction depends on the
initial concentration, in contrast to its constancy for a first-order
reaction. For this reason, the concept of half-life for a second-order
reaction is far less useful.

Zero-order processes
In some reactions, the rate is apparently independent of the reactant
concentration, in which case integration of the differential rate law
(rate = k[A]0 = k) yields [A]t = [A]0 kt.

265
Note the word "apparently" in the
preceding sentence; zero-order kinetics
is always an artifact of the conditions
under which the reaction is carried out.
For this reason, reactions that follow
zero-order kinetics are often referred
to as pseudo-zero-order reactions.
Clearly, a zero-order process cannot
continue after a reactant has been
exhausted. Just before this point is
reached, the reaction will revert to
another rate law instead of falling
directly to zero as depicted at the
upper left of figure 13.7.

Fig. 13.7: The integrated rate law for a zero-order reaction

There are two general conditions that can give rise to zero-order rates:
1. Only a small fraction of the reactant molecules are in a location
or state in which they are able to react, and this fraction is
continually replenished from the larger pool.
This situation commonly occurs when a reaction is catalysed by
attachment to a solid surface (heterogeneous catalysis) or to an
enzyme.
For example, the decomposition of nitrous oxide
N2O(g) N2(g) + O2(g)
in the presence of a hot platinum wire (which acts as a catalyst) is
zero-order, but it follows more conventional kinetics when
carried out entirely in the gas phase.
In this case, the N2O molecules that react are limited to those that
have attached themselves to the surface of the solid catalyst.
Once all of the sites on the limited surface of the catalyst have
been occupied, additional gas-phase molecules must wait until
the decomposition of one of the adsorbed molecules frees up a
surface site.

266
Enzyme-catalysed reactions in organisms begin with the
attachment of the substrate to the active site on the enzyme,
leading to the formation of an enzyme-substrate complex. If the
number of enzyme molecules is limited in relation to substrate
molecules, then the reaction may appear to be zero-order.
2. When two or more reactants are involved, the concentrations of
some are much greater than those of others.
Thus if the reaction
A + B products
is first-order in both reactants so that
rate = k[A][B]
then if B is present in great excess, the reaction will appear to be
zero order in B (and first order overall). This commonly happens
when B is H2O and the reaction is carried out in aqueous solution.

Summary
The following table compares the rate parameters of zero-, first-, and
second-order reactions of the form A products.
Table.13.2: Rate parameters of the three orders of reaction

zero order first order second order


Differential law rate = k rate = k[A] rate = k[A]2

Rate law [A] = [A] kt ln[A]o/[A] = kt 1/[A]- 1/[A]o = kt

Half-life [A]o/2k ln2/k 1/k[A]o

Straight line [A] vs. t ln [A] vs. t 1/[A] vs. t

13.5 Temperature and Rate

The rates of chemical reactions increase as temperature rises. The


increase in rate is due to the increase in rate constant as T is increased.
For many reactions, the rate constant doubles for every 10 degrees rise
in temperature.

267
The Arrhenius equation
Experimentally the rate constant is seen to be dependent on T
according to the equation:
k = A exp(Ea/RT)
From a set of k values at various values of T, A and E a can be
determined. The equation above can be written as:
ln k = ln A Ea/RT
A graph of ln k vs. 1/T gives a straight line. The equation is called the
Arrhenius equation and the plot is an Arrhenius plot. Ea is called the
Activation Energy.

Example 13.5:
For the isomerization of cyclopropane to propene: C3H6 (g) C3H6 (g),
the following data were obtained:
T (0C): 477 523 577 623
1/T (K1 103): 1.33 1.25 1.18 1.11
k (s1): 0.00018 0.0027 0.030 0.26
ln k: 8.62 5.92 3.51 1.35
Plot a graph of ln k vs. 1/T:

0
1 1.1 1.2 1.3 1.4
-4
ln k
-8 y = -33.1x + 35.6

-12
1/T

From the calculated slope, we have: (Ea/R) = 3.27 104 K


Ea = (8.314 J mol1 K1) (3.27 104 K) = 273 kJ mol1
Comment:
This activation energy is rather high, which is not surprising because a
carbon-carbon bond must be broken in order to open the cyclopropane
ring. (CC bond energies are typically around 350 kJ/mol.) This is why
the reaction must be carried out at high temperature.
268
You don't always need a plot. Since the ln k-vs.-1/T plot yields a
straight line, it is often convenient to estimate the activation energy
from experiments at only two temperatures. To see how this is done,
consider that

in which we have made the ln A term disappear by subtracting the


expressions for the two ln k terms. Solving the expression on the right
for the activation energy yields

13.6 Theories of Reaction Rates

Why are some reactions so much faster than others and what is the
explanation for the dependence of rate on temperature described by
Arrhenius equation? To answer these questions, chemists have
proposed theories as to what happens at the molecular level when a
reaction takes place.

There are the central questions discussed here. In doing so, we open
the door to the important topic of reaction mechanisms: what happens
at the microscopic level when chemical reactions take place? We can
thank Svante Arrhenius for unlocking this door!
To keep things as simple as possible, we will restrict ourselves to
reactions that take place in the gas phase. The same principles will
apply to reactions in liquids and solids.
Collision theory of chemical change
In order for molecules to react, they must collide. Consider a collision
between two molecules, A and BC to produce AB and C molecules
A + BC AB + C

269
Clearly, if two molecules A and BC are to react, they must approach
closely enough to disrupt some of their existing bonds and to permit
the creation of any new ones that are needed in the products. Such an
encounter is called a collision.
Not every collision, however, gives rise to products. According to
kinetic theory of gases, at room temperature and normal atmospheric
pressure, there will be about 1033 collisions in each cm3 every second.
If every collision between two reactant molecules yielded products, all
reactions would be complete in a fraction of a second.
During most collisions, molecules just bounce off without undergoing
disruption or rearrangement of the bonds between their atoms. There
are reasons for not every collision producing a reaction. These are
orientation factor and activation energy as explained below.

Orientation Factor: On of the requirements for the collision to be


effective is that the molecules have to be suitably oriented with
respect to one another during a collision. In the above example, the
molecule A must hit the B end of the molecule BC for the collision to
have a chance of being effective. The collision during which A hits
the C end of BC cannot produce the reaction.

A B C A B + C

A C B No reaction. Orientation is not correct

Figure 13.8: Collision reaction of A with BC

Owing to the extensive randomization of molecular motions in a gas


or liquid, there are always enough correctly-oriented molecules for
some of the molecules to react. But of course, the more critical this
requirement is of orientation is, the fewer collisions will be effective.

Activation Energy: A more important factor that determines which


collision will be effective is how much kinetic energy the colliding

270
molecules have. Only those collisions will be effective that are
between molecules possessing the needed kinetic energy for breaking
bonds. If the total energy of colliding molecules is not large enough,
they will simply bounce off each other and no reaction will take place.
The minimum energy that the molecules must have in order to react is
called activation energy. It is a positive quantity.
Activation Energy Diagrams:
Figure 13.9 is a typical activation energy diagram. It shows the
change in the total energy as the reaction goes from reactants to
products. Along the x-axis is the reaction coordinate or the reaction
path as the mixture progresses from reactants to products. The energy
of the products is shown at the right. The example in the diagram is
an exothermic reaction, where the energy of the products is lower than
that of the reactants. In going from the reactants to the products the
mixture has to cross an energy barrier represented by the peak in the
diagram.
For the case of the hypothetical example of molecule A reacting with
C, the ascending part of the curve represents stages in which a new
bond is forming between A and B at the same time as the bond
between B and C is being broken. At the peak, which represents the
activated complex, or the transition state, the structure might be
A---B---C, where the dotted lines represent partial bonds. Along the
descending part of the curve, the partial bond between B and C
changes to full bond resulting in formation of the molecule AB. At
the same time the partial bond between B and C breaks.
The energy difference between the reactants and the activated
complex is the activation energy of the forward reaction. The
activation energy of the reverse reaction is the difference between the
energy of the activated complex and the products.

The diagram shows that the enthalpy change, H, of the reaction, is


given by the difference in the energy of the product and the reactants.
This difference is independent of the path of the reaction.

271
Figure 13.9: Example of an activation energy diagram

Figure 13.10 shows the activation energy diagrams for both


exothermic and endothermic reactions.

Figure 13.10: Example of an activation energy diagram for an


exothermic (left) and endothermic reaction (right).

The collision theory discussed above gives the theoretical explanation


of the Arrhenius equation described earlier.

272
According to kinetic molecular theory, a population of molecules at a
given temperature is distributed over a variety of kinetic energies that
is described by the Maxwell-Boltzmann distribution law (figure
13.12).
The area under each curve represents the total number of molecules
whose energies fall within particular range. The shaded regions
indicate the number of molecules which are sufficiently energetic to
meet the requirements dictated by the two values of Ea that are shown.
It is clear from these plots that the fraction of molecules whose kinetic
energy exceeds the activation energy increases quite rapidly as the
temperature is raised. This is the reason that virtually all chemical
reactions are more rapid at higher temperatures.

Figure 13.12: Distribution plots of lower temperature T1 and a higher


temperature T2.
It can be shown that the fraction of molecules of a gas with energy
equal to or greater than Ea in given by exp(Ea/RT). This fraction
when multiplied with the total number of collisions taking place with
the correct orientation gives the rate constant. Thus,
k = A exp(Ea/RT),
where A represents the number of collisions that have the correct
orientation. A is called frequency factor, or the pre-exponential
factor.
273
13.7 Catalysts
Catalysts are substances that speed up a reaction but which are not
consumed by it and do not appear in the net reaction equation. Also
and this is very important catalysts affect the forward and
reverse rates equally; this means that catalysts have no effect on the
equilibrium constant and thus on the composition of the equilibrium
state.

A catalyst provides an alternative, lower activation energy pathway


between reactants and products. As such, they are vitally important to
chemical technology; approximately 95% of industrial chemical
processes involve catalysts of various kind. In addition, most
biochemical processes that occur in living organisms are mediated by
enzymes, which are catalysts made of proteins.

It is important to understand
that a catalyst affects only the
kinetics of a reaction; it does
not alter the thermodynamic
tendency for the reaction to
occur. Thus there is a single
value of H for the two
pathways depicted in the plot
in figure 13.11 on the right.

Fig. 13.11: Activation energy with and without a catalyst


Since the rate constant of a reaction is an exponential function of the
activation energy, even a moderate reduction of Ea can yield an
impressive increase in the rate.

274
13.8 Enzymes
Catalysts are conventionally divided into two categories:
homogeneous and heterogeneous. Enzymes, natural biological
catalysts, are often included in the former group, but because they
share some properties of both but exhibit some very special properties
of their own, we will treat them here as a third category.
A mechanism that has been proposed for the action of enzyme is the
lock and key mechanism. Enzymes are highly specific. An enzyme is
usually a large protein molecule that has active sites where the
reactants (called substrates) fit like a key fits a lock. The reaction can
take place between the substrate molecules before the products
molecules get desorbed. The following diagram shows the steps of
the process.

Reactants (substrate) Products

Enzyme Complex

13.9 Exercises

13.1 For the reaction:


2NO(g) + O2(g) 2NO2(g)
how does the rate of disappearance of NO compare to the rate
of disappearance of O2, and the rate of appearance of NO2?

13.2 The decomposition of C2H5Cl takes place as follows:


C2H5Cl(g) C2H4(g) + HCl(g).
In an experiment at 600 oC, the concentration of the reactant
decreased from 0.0250 M to 0.0142 M in 160 s. What is the
average rate of the reaction?

275
13.3 The initial rate method was followed to study the reaction:
2NO(g) + Cl2(g) 2NOCl(g).
The following data were obtained for rates at the given initial
concentrations:
[NO]o [Cl2]o Initial Rate
Exp. 1: 0.0240 M 0.0180 M 0.00341 M/s
Exp. 2: 0.0480 M 0.0180 M 0.0136 M/s
Exp. 3: 0.0240 M 0.0360 M 0.00682 M/s
Obtain the rate law. What is the value of the rate constant?

13.4 For the first order isomerization reaction:


CH3NC CH3CN
the rate constant is 1.2 10-3 s-1. In an experiment, the initial
concentration of CH3CN was 0.462 M. What is the
concentration after 40 minutes?

13.5 What is the half-life of the first order reaction


SO2Cl2 (g) SO2(g) + Cl2(g)?
The rate constant at a particular temperature is 3.6 10-4 s-1.
How long will it take for the concentration to fall to 15 % of
the initial value?

13.6 The reaction A Z is second order with a rate constant 0.648


L/(mol.s). If the initial concentration of A is 0.0463 M, what is
the concentration after 26 s? What is the half-life of the
reaction?

13.7 The decomposition of ammonia gas on gold surface is zero


order. If the rate constant for this reaction is 5.2 10-2
mol/(L.s), how long will it take for the concentration of
ammonia to change from an initial concentration of 0.17 M to
0.0084 M?

13.8 The rate constant for a first order reaction is 6.40 10-3 s-1 at
280 oC. If the activation energy is 86.0 kJ/mol, find the
temperature at which the rate constant is 1.24 10-2 s-1.

13.9 On raising the temperature of a reaction from 700 K to 760 K,


the rate increases 10-fold. What is the activation energy of the
reaction?
276
13.10 For the decomposition of N2O into nitrogen and oxygen, the
following data were obtained:

Rate constant (L.mol-1.s-1) Temperature (K)


0.0124 925
0.0574 980
0.250 1030
Determine the activation energy graphically.

277
CHAPTER 14
Introduction to chemical equilibrium
14.1 Chemical Equilibrium

When a chemical reaction takes place in a container that prevents the


entry or escape of any of the substances involved in the reaction, the
quantities of these components change as some are consumed and
others are formed. Eventually this change will come to an end, after
which the composition will remain unchanged as long as the system
remains undisturbed. The system is then said to be in its equilibrium
state, or more simply, "at equilibrium".

Let us consider the reaction between hydrogen and iodine in the


gaseous state in a closed vessel to produce HI:
H2(g) + I2(g) 2HI(g)
As the reaction proceeds, the concentration of each of the reactants
decreases and that of the product HI increases. A stage is reached
when there is no further change in the concentration of any of the
species. The system is said to have reached equilibrium.
The reaction continues to go on in both the directions forward and
reverse. At equilibrium, the rate of the forward reaction equals the
rate of the backward reaction; hence there is no net change in the
concentrations. Chemical equilibrium is an example of dynamic
equilibrium. The reaction at equilibrium is written as:
H2(g) + I2(g) 2HI(g)

Figure 14.1 shows how the concentrations of the three components of


this chemical reaction change with time. Examine the two sets of
plots carefully, noting which substances have zero initial
concentrations, and are thus "products" of the reaction equations
shown. Satisfy yourself that these two sets represent the same
chemical reaction system, but with the reactions occurring in opposite
directions. Most importantly, note how the final (equilibrium)
concentrations of the components are the same in the two cases.
Whether we start with an equimolar mixture of H2 and I2 (left) or a
pure sample of hydrogen.iodide (shown on the right in figure 14.1,

278
using twice the initial concentration of HI to keep the number of
atoms the same), the composition after equilibrium is attained will be
the same.
The equilibrium composition is independent of the direction from
which it is approached.

Figure 14.1: Chemical reaction system for H2(g) + I2(g) 2HI(g)

14.2 The Equilibrium Constant

For the general reaction:


aA(g) + bB(g) cC(g) + dD(g)
the ratio [C]c [D]d/ [A]a [B]b is always constant at equilibrium, and is
denoted by the symbol Kc. This ratio has a specific value for a
reaction at a fixed temperature no matter what the starting
concentrations are.
For the hydrogen and iodine reaction, Kc = [HI]2 / [H2] [I2]
Kc is written without units, even though the expression may yield units
when concentrations are substituted in the expression.

Equilibrium: kinetic point of view


Consider a single-step reaction:
AB
Forward rate = kf [A]
Reverse rate = kr [B]; where kf and kc are the rate constants for the
forward and reverse reactions respectively.
279
At equilibrium, the two rates are equal.
Hence, kf [A] = kr [B]
and kf / kr = [B] / [A]
The left-hand side of this equation is the ratio of rate constants, which
is a constant at a particular temperature. This ratio can be identified
with Kc. Thus Kc= kf / kr.

Some examples of expressions of Kc


To write the expression for an equilibrium constant, we have to know
the balanced chemical equation. If the equation is written in reverse
or with different coefficients, the expression for Kc will be different.
For the equation:
N2(g) + 3H2(g) 2NH3(g),
Kc = [NH3]2 / [N2] [H2]3.
If the equation is rewritten as:
1/2 N2(g) + 3/2 H2(g) NH3(g),
then the equilibrium constant would be given by:
Kc = [NH3] / [N2]1/2 [H2]3/2
showing that Kc = Kc.
It is apparent that if the equation is multiplied by n, the equilibrium
constant is raised to the power n.
If the equation is written as:
2NH3(g) N2(g) + 3H2(g),
then its equilibrium constant, Kc, in relation to Kc is:
Kc= 1 / Kc.

Adding reactions
When reactions are added to get an overall reaction, the equilibrium
constant of the overall reaction is obtained by multiplying individual
Kc values. An example is the reaction between P(g) and Cl2(g):
(1) P(g) + 3/2 Cl2(g) PCl3(g); Kc1
(2) PCl3(g) + Cl2(g) PCl5(g); Kc2
280
(3) P(g) + 5/2 Cl2(g) PCl5(g); Kc3
Here reaction (3) is obtained by adding reactions (1) and (2). By
writing out the expressions for all the three equilibrium constants, it
can be verified that
Kc3 = Kc1 Kc2
So, if reactions are added, Ks are multiplied.

Example 14.1:
Given the following equilibrium constants:
(1) CaCO3(s) Ca2+(aq) + CO32(aq) K1 = 106.3
(2) HCO3(aq) H+(aq) + CO32(aq) K2 = 1010.3
Calculate K for the reaction CaCO3(s) + H+(aq) Ca2+(aq) + HCO3(aq)
Solution:
The net reaction is the sum of reaction (1) and the reverse of reaction (2).
So the net reaction can be obtained by (1) (2). Hence, for the net
reaction, K = K1/K2 = 106.3/1010.3 = 10+1.9
Comment:
This net reaction describes the dissolution of limestone by acid; it is
responsible for the eroding effect of acid rain on buildings and statues.
This is an example of a reaction that has practically no tendency to take
place by itself (small K1), being "driven" by a second reaction having a
large equilibrium constant (K2). From the standpoint of the Le Chtelier
principle (to be discussed shortly), the first reaction is "pulled to the right"
by the removal of carbonate by hydrogen ion. Coupled reactions of this
type are widely encountered in all areas of chemistry, and especially in
biochemistry, in which a dozen or so reactions may be linked.

Equilibrium constant in terms of pressures, Kp


For ideal gases, partial pressures are proportional to concentrations.
Hence, equilibrium constant expression can also be written in terms of
partial pressures.
Consider the reaction:
2NO2(g) N2O4(g)
Kp for this reaction is defined as:

281
Kp = P(N2O4) / [P(NO2)]2
where P(N2O4) and P(NO2) are the partial pressures of the gases in the
mixture.
Kp like Kc is written without units.

Relationship between Kp and Kc


Let us consider the general reaction:
aA(g) + bB(g) cC(g) + dD(g)
If we assume that the gases are ideal, then for each of the gases in the
mixture expressions like the following hold:
PA V = nA RT
PA = (nA/V) RT = [A] RT
Substituting these into Kp, we get:
Kp = PCc .PDd / PAa. PBb = ([C] RT)c.([D] RT)d / ([A] RT)a.([B] RT)b
= {[C]c [D]d / [A]a [B]b}(RT)c+d-a-b
= Kc (RT)c+d-a-b = Kc (RT)n,
where n = (c + d a b) = moles of gaseous products moles of
gaseous reactants. R should be in L.atm.K-1.mol-1.
Thus for the reaction, 2NO2(g) N2O4(g)
KP = Kc (RT)1-2 = KC / RT
Similarly for the reaction, H2(g) + I2(g) 2HI(g),
Kp = Kc, since n = 0.

What can we learn from the value of the equilibrium constant?


A large value of the equilibrium constant tells us that the equilibrium
mixture will have products mostly. A small value of the equilibrium
constant indicates that the reactants will predominate. An
intermediate value shows that products and reactants are both present
in comparable concentrations.
As an example, the reaction 2NO2 (g) 2NO(g) + O2(g) has Kc =
2.5 10-14 at 25 oC. We know from looking at the value of the

282
equilibrium constant that the equilibrium mixture will contain mostly
NO2.
For the reaction: 2NO(g) N2(g) + O2(g), Kc = 2.2 x 1030. The
equilibrium mixture will have mostly products and negligible amounts
of NO will be present.
If the value of Kc is between 0.01 and 100, both reactants and products
are present in appreciable quantities.

Example 14.2:
For the reaction
CO(g) + 3H2(g) CH4(g) + H2O(g),
Kc = 1.38 103 at 800K, and Kc = 3.92 at 1200 K.
Is the formation of methane favoured at higher or lower temperature?
Solution:
The equilibrium constant is larger at the lower temperature, so the forward
reaction is favoured at 800K as compared to 1200 K.

14.3 Heterogeneous Equilibrium

So far, we have considered homogenous equilibria, where the reaction


mixture comprises one phase only. How do we treat reactions that
have different phases, such as solids and liquids or gases; in other
words heterogeneous equilibrium?
An example is the reaction
CaCO3(s) CaO(s) + CO2 (g)
While writing the equilibrium constant, we leave out solids. Thus,
Kc = [CO2].
Pure liquids are also excluded. The justification is that the
concentration of a solid is constant. As the amount of solid increases,
its volume increases as well, leaving the concentration unchanged.
The same applies to pure liquids.
Another example is
PbI2(s) Pb2+(aq) + 2I-(aq) with Kc = [Pb2+] [I-]2.

283
Solvent too is not included in the expression for Kc, since the
concentration of the solvent is nearly constant if the solution is dilute.
Thus for the equilibrium:
NH3(aq) + H2O(l) NH4+ (aq) + OH-(aq)
Kc = [NH4+] [OH-] / [NH3].
Thus, the substances whose concentrations undergo no significant
change in a chemical reaction do not appear in equilibrium constant
expressions.

How can the concentration of a reactant or product not change when a


reaction involving that substance takes place? There are two general
cases to consider.
The substance is also the solvent
This happens all the time in acid-base chemistry. Thus, for the
hydrolysis of the cyanide ion
CN(aq) + H2O(l) HCN(aq) + OH(aq)
we write
Kc = [HCN][OH-] / [CN-]
in which no [H2O] term appears. The justification for this omission is
that water is both the solvent and reactant, but only the tiny portion
that acts as a reactant would ordinarily go in the equilibrium
expression. The amount of water consumed in the reaction is so
minute (because Kc is very small) that any change in the concentration
of H2O from that of pure water (55.5 mol L1) will be negligible.
Similarly, for the "dissociation" of water:
H2O(l) H+(aq) + OH(aq)
the equilibrium constant is expressed as the "ion product":
Kw = [H+][OH].

But be careful about throwing away H2O whenever you see it. In the
esterification reaction
CH3COOH(l) + C2H5OH(l) CH3COOC2H5(l) + H2O(l)

284
a [H2O] term must be present in the equilibrium expression if the
reaction is assumed to be between the two liquids acetic acid and
ethanol. If, on the other hand, the reaction takes place between a
dilute aqueous solution of the acid and the alcohol, then the [H2O]
term would not be included. The reaction would then be written as:
CH3COOH(aq) + C2H5OH(aq) CH3COOC2H5(aq) + H2O(l)

The substance is a solid or a pure liquid phase.


This is most frequently seen in solubility equilibria, but there are
many other reactions in which solids are directly involved:
CaF2(s) Ca2+(aq) + 2F(aq)
Fe3O4(s) + 4H2(g) 4H2O(g) + 3Fe(s)
These are heterogeneous reactions (meaning reactions in which some
components are in different phases), and the argument here is that
concentration is only meaningful when applied to a substance within a
single phase.
Thus the term [CaF2] would refer to the concentration of calcium
fluoride within the solid CaF2", which is a constant depending on the
molar mass of CaF2 and the density of that solid. The concentrations
of the two ions will be independent of the quantity of solid CaF 2 in
contact with the water; in other words, the system can be in
equilibrium as long as any CaF2 at all is present.

Throwing out the constant-concentration terms can lead to some rather


sparse-looking equilibrium expressions. For example, the equilibrium
expression for the following process consists solely of a single term
involving the partial pressure of a gas:
CaCO3(s) CaO(s) + CO2(g) Kp = PCO2

14.4 Calculations with Kc and Kp

If concentrations of all the reactants and products are known at


equilibrium, Kc can be calculated by substituting them in the
expression directly. Similarly, if partial pressures are known, Kp can
be calculated.
Example 14.3:

285
A 2.00 L vessel had 0.0150 mol PCl3, 0.0118 mol PCl5 and 0.0760 mol
Cl2 at a certain temperature. What is the value of Kc for the reaction:
PCl5(g) PCl3 (g) + Cl2 (g)
Solution:
Concentrations at equilibrium are: [PCl5] = 0.0118/2.0 mol/L;
[PCl3]= 0.0150/2.00; [Cl2] = 0.0760/2.0 mol/L.
Kc = [PCl3] [Cl2] / [PCl5] = (0.0150)(0.076)/(2 0.0118) = 0.0483

If all the initial concentrations are known and at least one of the
equilibrium concentrations is known, Kc can be calculated. To get the
rest of the concentrations at equilibrium, the use is made of
stoichiometry, as illustrated in the example below:
Example 14.4:
In the contact process for the manufacture of sulphuric acid, sulphur
dioxide is first oxidized to sulphur trioxide which is then reacted with
water. The equilibrium is
2SO2(g) + O2(g) 2SO3(g)
The initial partial pressures of SO2 and O2 were 0.600 atm and 0.300 atm
at 1000 K; at equilibrium the partial pressure of SO3 is 0.380 atm. Find
Kp.
Solution:
2SO2(g) + O2(g) 2SO3(g)
Initial 0.600 0.300 0
Change -2x -x +2x
Equilibrium 0.600-2x 0.300-x 2x
From the data given, 2x = 0.380 atm; so, x = 0.380/2 = 0.190 atm
Hence, the equilibrium partial pressure of SO2
= 0.600 2 x 0.190 = 0.220 atm.
For O2 the equilibrium partial pressure
= 0.300 0.190 = 0.110 atm.
All the equilibrium partial pressures are now known, so
Kp = (0.380)2 / (0.220)2 (0.110) = 27.1

Predicting the direction of reactions from Kc and Kp

286
If we are given the actual concentrations of all substances in a system,
and we need to know in which direction the reaction would go, we
follow the procedure described below. Suppose the reaction is:
a A(g) + b B(g)c C(g) + d D(g)
We are told that a container has some concentrations of each of the
substances. What will be the direction of the reaction? To answer
that we define a quantity called the reaction quotient, Q. Q has the
same form as Kc , but it uses the actual concentrations rather than the
equilibrium ones. For the above reaction, Qc = [C]c [D]d/ [A]a [B]b,
where the concentrations are the actual ones present.

It is clear that at equilibrium Qc = Kc. At any point other than


equilibrium, if Qc < Kc, the reaction will go to right and C and D will
tend to form, in order for Qc to increase and become equal to Kc. If on
the other hand, Qc is greater than Kc, the reaction will go backwards in
order for Qc to decrease and become equal to Kc. In short if:
Qc < Kc; the reaction will go forward.
Qc > Kc; the reaction will go backwards.
Qc = Kc; the reaction is at equilibrium.

Example 14.5:
For the reaction N2O4 (g) 2NO2(g), Kc = 0.125 at 25 oC.
If a vessel has [N2O4] = 0.0200 M and [NO2] = 0.0580 M, in which
direction will the reaction go?
Solution:
Qc = [NO2]2 / [N2O4] = (0.0580)2/(0.0200) = 0.168
Qc> Kc; so the reaction will go backwards and N2O4 will tend to form.
If it is Kp that we are dealing with rather than Kc, the treatment is similar,
as in the following example.

Example 14.6:
The equilibrium constant for the oxidation of sulphur dioxide is Kp = 0.14
at 900 K: 2 SO2(g) + O2(g) 2 SO3(g)

287
If a reaction vessel is filled with SO3 at a partial pressure of 0.10 atm and
with O2 and SO2 each at a partial pressure of 0.20 atm, what can you
conclude about whether, and in which direction, any net change in
composition will take place?
Solution:
The value of the equilibrium quotient Q for the initial conditions is

Since Q > K, the reaction is not at equilibrium, so a net change will occur
in a direction that decreases Q. This can only occur if some of the SO3 is
converted back into products. In other words, the reaction will "shift to
the left".

Predicting equilibrium concentrations


This is by far the most common kind of equilibrium problem you will
encounter: starting with an arbitrary number of moles of each
component, how many moles of each will be present when the system
comes to equilibrium?
The principal source of confusion and error for beginners relates to the
need to determine the values of several unknowns (a concentration or
pressure for each component) from a single equation, the equilibrium
expression. The key to this is to make use of the stoichiometric
relationships between the various components, which usually allow us
to express the equilibrium composition in terms of a single variable.
The easiest and most error-free way of doing this is adopt a systematic
approach in which you create and fill in a small table as shown in the
following problem example. You then substitute the equilibrium
values into the equilibrium constant expression, and solve it for the
unknown.
This very often involves solving a quadratic or higher-order equation.
Quadratics can of course be solved by using the familiar quadratic
formula, but it is often easier to use an algebraic or graphical
approximation, and for higher-order equations this is the only practical
approach. There is almost never any need to get an exact answer,
since the equilibrium constants you start with are rarely known all that
precisely anyway.

288
Example 14.7:
Phosgene (COCl2) is a poisonous gas that dissociates at high temperature
into two other poisonous gases, carbon monoxide and chlorine. The
equilibrium constant Kp = 0.0041 at 600K. Find the equilibrium
composition of the system after 0.124 atm of COCl2 is allowed to reach
equilibrium at this temperature.
Solution:
Start by drawing up a table showing the relationships between the
components:
COCl2 CO Cl2
initial pressures: 0.124 atm 0 0
change: x +x +x
equilibrium pressures: 0.124 x x x

Substitution of these pressures into the equilibrium expression gives

This expression can be rearranged into standard polynomial form


x2 + 0.0041 x 0.00054 = 0 and solved by the quadratic formula, but we
will simply obtain an approximate solution by iteration. Because the
equilibrium constant is small, we know that x will be rather small
compared to 0.124, so the above relation can be approximated by

which gives x = 0.0225. To see how good this is, substitute this value of x
into the denominator of the original equation and solve again:

This time, solving for x gives 0.0204. Iterating once more, we get

and x = 0.0206 which is sufficiently close to the previous to be considered


the final result. The final partial pressures are then 0.104 atm for COCl2,
and 0.0206 atm each for CO and Cl2.
289
Comment:
Using the quadratic formula to find the exact solution yields the two roots
0.0247 (which we ignore) and 0.0206, which shows that our
approximation is quite good.
Example 14.8:
The gas-phase dissociation of phosphorus pentachloride to the trichloride
has Kp = 3.60 at 540C:
PCl5(g) PCl3(g) + Cl2(g)
What will be the partial pressures of all three components if 0.200 mole of
PCl5 and 3.00 moles of PCl3 are combined and brought to equilibrium at
this temperature and at a total pressure of 1.00 atm?
Solution:
As always, set up a table showing what you know (first two rows) and
then expressing the equilibrium quantities:
PCl5 PCl3 Cl2
initial moles 0.200 3.00 0
change x +x +x
equilibrium
0.200 x 3.00 + x x
moles
equilibrium
(2.00-x)/(3.20+x) (3.00+x)/(3.20+x) x/(3.20 + x)
partial pressures

The partial pressures in the bottom row were found by multiplying the
mole fraction of each gas by the total pressure: Pi = XiPt. The term in the
denominator of each mole fraction is the total number of moles of gas
present at equilibrium: (0.200 x) + (3.00 + x) + x = 3.20 + x.
Substituting the equilibrium partial pressures into the equilibrium
expression, we have

whose polynomial form is 4.60 x2 + 13.80 x 2.304 = 0. This equation


can be solved for x using the quadratic formula, or it can be solved
graphically as shown below.

290
Plotting this on a graphical calculator yields x = 0.159 as the positive root.
Substitution of this root into the expressions for the equilibrium partial
pressures in the table yields the following values: P(PCl5) = 0.012 atm,
P(PCl3) = 0.94 atm, P(Cl2) = 0.047 atm.

14.5 Le Chtelier Principle

If a reaction is at equilibrium and we alter the conditions so as to


create a new equilibrium state, then the composition of the system will
tend to change until that new equilibrium state is attained. (We say
"tend to change" because if the reaction is kinetically inhibited, the
change may be too slow to observe or it may never take place.) In
1884, the French chemical engineer and teacher Henri Le Chtelier
(1850-1936) showed that in every such case, the new equilibrium state
is one that partially reduces the effect of the change that brought it
about.
This law is known as the Le Chtelier principle. His original
formulation was somewhat complicated, but a reasonably useful
paraphrase of it reads as follows:
Le Chtelier principle: If a system at equilibrium is subjected to a
change of pressure, temperature, or the number of moles of a
component, there will be a tendency for a net reaction in the direction
that reduces the effect of this change.
To see how this works (and you must do so, as this is of such
fundamental importance that you simply cannot do any meaningful
chemistry without a thorough working understanding of this
principle), look again at the hydrogen iodide dissociation reaction
2 HI(g) H2 (g) + I2 (g)
Consider an arbitrary mixture of these three components at
equilibrium, and
assume that we
inject more
hydrogen gas into
the container.
Because the H2
concentration
now exceeds its

291
new equilibrium value, the
system is no longer in its
equilibrium state, so a net
reaction now ensues as the
system moves to the new state
The Le Chtelier principle states
that the net reaction will be in a direction that tends to reduce the
effect of the added H2. This can occur if some of the H2 is consumed
by reacting with I2 to form more HI; in other words, a net reaction
occurs in the reverse direction. Chemists usually simply say that "the
equilibrium shifts to the left".

To get a better idea of how this works, carefully examine the diagram
below which follows the concentrations of the three components of
this reaction as they might change in time (the time scale here will
typically be about an hour).

Disruption and restoration of equilibrium


At the left, the concentrations of the three components do not change
with time because the system is at equilibrium. We then add more
hydrogen to the system, disrupting the equilibrium. A net reaction
then ensures that moves the system to a new equilibrium state (right)
in which the quantity of hydrogen iodide has increased; in the process,
some of the I2 and H2 are consumed. Notice that the new equilibrium
state contains more hydrogen than did the initial state, but not as much
as was added; as the le Chtelier principle predicts, the change we
292
made (addition of H2) has been partially counteracted by the "shift to
the right".
The following table contains several examples showing how changing
the quantity of a reaction component can shift an established
equilibrium.

Table. 14. 2
system change result
CO2 + H2 H2O absorbed Shift to the right. Continuous
H2O(g) + CO on drying removal of a product will
agent force any reaction to the right
H2(g) + I2(g) 2HI(g) Some No change; N2 is not a
nitrogen gas component of this reaction
is added system.
NaCl(s) + H2SO4(l) HCl (g) escapes from the
Na2SO4(s) + HCl(g) system, the reaction is forced
Container
to the right. This is the basis
opened
for the commercial production
of hydrochloric acid.
H2O(l) H2O(g) Removal of water vapor
Container forces the reaction to the
opened right, so equilibrium is never
achieved.
AgCl(s) some NaCl is
Shift to left due to increase in
Ag+(aq) + Cl(aq) Cl concentration. This is
added to the
known as the common ion
solution
effect on solubility.
N2 + 3 H2 2 NH3 a catalyst is
No change. Catalysts affect
only the rate of a reaction; the
added to
have no effect at all on the
speed up this
composition of the
reaction
equilibrium state.

How do changes in temperature affect equilibria?

293
Virtually all chemical reactions are accompanied by the liberation or
uptake of heat. If we regard heat as a "reactant" or "product" in an
endothermic or exothermic reaction respectively, we can use the Le
Chtelier principle to predict the direction in which an increase or
decrease in temperature will shift the equilibrium state. Thus for the
oxidation of nitrogen, an endothermic process, we can write
[heat] + N2(g) + O2(g) 2 NO (g)
Suppose this reaction is at equilibrium at some temperature T1 and we
raise the temperature to T2. The Le Chtelier principle tells us that a
net reaction will occur in the direction that will partially counteract
this change. Since the reaction is endothermic, a shift of the
equilibrium to the right will take place.
Nitric oxide, the product of this reaction, is a major air pollutant which
initiates a sequence of steps leading to the formation of atmospheric
smog. Its formation is an unwanted side reaction which occurs when
the air (which is introduced into the combustion chamber of an engine
to supply oxygen) gets heated to a high temperature. Designers of
internal combustion engines now try, by various means, to limit the
temperature in the combustion region, or to restrict its highest-
temperature part to a small volume within the combustion chamber.

How do changes in pressure affect equilibria?


You will recall that if the pressure of a gas is reduced, its volume will
increase; pressure and volume are inversely proportional. With this in
mind, suppose that the reaction
2 NO2(g) N2O4(g)
is in equilibrium at some arbitrary temperature and pressure, and that
we double the pressure, perhaps by compressing the mixture to a
smaller volume. From the Le Chtelier principle we know that the
equilibrium state will change to one that tends to counteract the
increase in pressure. This can occur if some of the NO2 reacts to form
more of the dinitrogen tetroxide, since two moles of gas are being
removed from the system for every mole of N2O4 formed, thereby
decreasing the total volume of the system. Thus increasing the
pressure will shift this equilibrium to the right.

294
It is important to understand that changing the pressure will have a
significant effect only on reactions in which there is a change in the
number of moles of gas as a result of the reaction.
For the above reaction, this change
ng = (nproducts nreactants) = 1 2 = 1.
In the case of the nitrogen oxidation reaction
N2(g) + O2(g) 2 NO (g), ng = 0 and pressure will have no effect.
For reactions involving gases, only changes in the partial pressures of
those gases directly involved in the reaction are important; the
presence of other gases has no effect.
The volumes of solids and liquids are hardly affected by the pressure
at all, so for reactions that do not involve gaseous substances, the
effects of pressure changes are ordinarily negligible. Exceptions arise
under conditions of very high pressure such as exist in the interior of
the Earth or near the bottom of the ocean. A good example is the
dissolution of calcium carbonate CaCO3(s) Ca2+(aq) + CO32(aq).
There is a slight decrease in the volume when this reaction takes
place, so an increase in the pressure will shift the equilibrium to the
right, with the results that calcium carbonate becomes more soluble at
higher pressures.
The skeletons of several varieties of microscopic organisms that
inhabit the top of the ocean are made of CaCO3, so there is a continual
rain of this substance toward the bottom of the ocean as these
organisms die. As a consequence, the floor of the Atlantic Ocean is
covered with a blanket of calcium carbonate. This is not true for the
Pacific Ocean, which is deeper; once the skeletons fall below a certain
depth, the higher pressure causes them to dissolve. Some of the
seamounts (undersea mountains) in the Pacific extend above the
solubility boundary so that their upper parts are covered with CaCO 3
sediments.
The effect of pressure on a reaction involving substances whose
boiling points fall within the range of commonly encountered
temperature will be sensitive to the states of these substances at the
temperature of interest.

Example 14.9:

295
The commercial production of hydrogen is carried out by treating
natural gas with steam at high temperatures and in the presence of a
catalyst (steam reforming of methane):
CH4 + H2O CH3OH + H2
Given the following boiling points: CH4 (methane) = 161C, H2O =
100C, CH3OH = 65, H2 = 253C, predict the effects of an increase
in the total pressure on this equilibrium at 50, 75 and 120C.
Solution:
Calculate the change in the moles of gas for each process:

temp equation ng shift


50 CH4(g) + H2O(l) CH3OH(l) + H2(g) 0 none
75 CH4(g) + H2O(l) CH3OH(g) + H2(g) +1 to left
120 CH4(g) + H2O(g) CH3OH(g) + H2(g) 0 none

What is the Haber process and why is it important?


The Haber process for the synthesis of ammonia is based on the
exothermic reaction
N2(g) + 3 H2(g) 2 NH3(g) H = 92 kJ/mol
The Le Chtelier principle tells us that in order to maximize the
amount of product in the reaction mixture, it should be carried out at
high pressure and low temperature. However, the lower the
temperature, the slower the reaction (this is true of virtually all
chemical reactions.) As long as the choice had to be made between a
low yield of ammonia quickly or a high yield over a long period of
time, this reaction was infeasible economically.
Nitrogen is available for free, being the major component of air, but
the strong triple bond in N2 makes it extremely difficult to incorporate
this element into species such as NO3 and NH4+ which serve as the
starting points for the wide variety of nitrogen-containing compounds
that are essential for modern industry. This conversion is known as
nitrogen fixation, and because nitrogen is an essential plant nutrient,
modern intensive agriculture is utterly dependent on huge amounts of
fixed nitrogen in the form of fertilizer. Until around 1900, the major
source of fixed nitrogen was the NaNO3 (Chile saltpeter) found in
extensive deposits in South America. Several chemical processes for
296
obtaining nitrogen compounds were developed in the early 1900's, but
they proved too inefficient to meet the increasing demand.
Although the direct synthesis of ammonia from its elements had been
known for some time, the yield of product was found to be negligible.
In 1905, Fritz Haber (1868-1934) began to study this reaction,
employing the thinking initiated by Le Chtelier and others, and the
newly-developing field of thermodynamics that served as the basis of
these principles. From the Le Chtelier law alone, it is apparent that
this exothermic reaction is favored by low temperature and high
pressure. However, it was not as simple as that: the rate of any
reaction increases with the temperature, so working with temperature
alone, one has the choice between a high product yield achieved only
very slowly, and a very low yield quickly. Further, the equipment and
the high-strength alloy steels need to build it did not exist at the time.
Haber solved the first problem by developing a catalyst that would
greatly speed up the reaction at lower temperatures.
The second problem, and the development of an efficient way of
producing hydrogen, would delay the practical implementation of the
process until 1913, when the first plant based on the Haber-Bosch
process (as it is more properly known, Carl Bosch being the person
who solved the major engineering problems) came into operation. The
timing could not have been better for Germany, since this country was
about to enter the First World War, and the Allies had established a
naval blockade of South America, cutting off the supply of nitrate for
the German munitions industry.
Bosch's plant operated the ammonia reactor at 200 atm and 550C.
Later, when stronger alloy steels had been developed, pressures of
800-1000 atm became common. The source of hydrogen in modern
plants is usually natural gas, which is mostly methane:

CH4 + H2O CO + 3 H2 formation of synthesis gas from methane


CO + H2O CO2 + H2 shift reaction carried out in reformer

The Haber-Bosch process is considered the most important chemical


synthesis developed in the 20th century.
Besides its scientific importance as the first large-scale application of
the laws of chemical equilibrium, it has had tremendous economic and
social impact; without an inexpensive source of fixed nitrogen, the
297
intensive crop production required to feed the world's growing
population would have been impossible. Haber was awarded the 1918
Nobel Prize in Chemistry in recognition of his work. Carl Bosch, who
improved the process, won the Nobel Prize in 1931.
14.6 Exercises
14.1 Write the expressions for Kc for the following reactions:
(a) N2(g) + 3H2(g) 2NH3(g)
(b) CaCO3(s) CaO(s) + CO2(g)
(c) 2ZnS(s) + 3O2(g) 2ZnO(s) + 2SO2(g)
(d) H2O (g) + C(s) CO(g) + H2(g)
(e) H2O(l) H+(aq) + OH- (aq)

14.2 The following equilibrium was established at a certain temperature:


CH4(g) + H2O (g) CO(g) + 3H2(g).
The concentrations at equilibrium are:
[CH4] = 0.280 M, [H2O] = 0.0840 M, [CO] = 0.486 M, [H2] = 0.680
M. What is Kc at this temperature?

14.3 For the reaction CO(g) + Cl2(g) COCl2 (g), Kc = 3.4 at 800 K.
Calculate Kp.

14.4 For the reaction 2H2 (g) + S2 (g) 2H2S(g), Kc = 1.8 x 107 at 1000
K. Calculate Kc for the reaction H2S(g) H2 (g) + S2 (g).

14.5 The equilibrium constant Kp for the reaction


PCl3(g) + Cl2 (g) PCl5(g)
is 0.952 at 250 oC. If a reaction mixture has partial pressures of
PCl3(g), Cl2 (g) and PCl5(g) equal to 0.120 atm, 0.246 atm and
0.180 atm respectively, when the reaction reaches equilibrium will
the pressure of PCl5 have decreased or increased over the starting
pressure?

14.6 A 0.084 mol sample of C2H5OH was placed in a 2.00 L vessel at a


certain temperature and the following equilibrium was reached.
C2H5OH(g) C2H4 (g) + H2O (g).
At equilibrium the concentration of ethanol was 0.018 M. What is
the concentration of each species at equilibrium? What is the value
of Kc?

14.7 The equilibrium constant for the reaction:

298
2NO(g) N2 (g) + O2(g) is 20.1 at 2200 oC. Starting with 2.80
moles of NO in a 14.0 L vessel, calculate the concentration of each
of the gases at equilibrium.

14.8 Consider the following endothermic reaction at equilibrium:


C(s) + CO2(g) 2CO(g)
If the reaction at equilibrium is subjected to the following changes,
what would the result be?
(a) If the volume of the vessel is decreased what happens to the
amount of CO?
(b) If some C(s) is added from outside what will happen to the
concentration of CO2?
(c) If some CO is removed, what happens to the concentration of
CO2?
(d) If some CO2 is removed from the mixture, what happens to the
amount of C(s)?
(e) If the temperature is increased what happens to amount of
carbon dioxide.
(f) If the temperature is decreased what happens to Kp?

14.9 A mixture of 5.75 atm of H2 and 5.75 atm of I2 is contained in a


1.0 L reaction vessel at 430oC. The equilibrium constant, Kc =
54.3 for the reaction
H2(g) + I2(g) 2HI(g),
Make an equilibrium table showing the initial partial pressures, the
change in partial pressures and the equilibrium partial pressures for
all the gases. Calculate the values of the equilibrium partial
pressure of each gas.
14.10 A mixture containing 2.4 moles of NO2 and 1.7 moles of CO was
allowed to react at a certain temperature, and the following
equilibrium was reached:
NO2(g) + CO(g) NO(g) + CO2(g)
At equilibrium, 1.8 mol NO2 was present. What is Kc?

299
CHAPTER 15
Acids and Bases
15.1 What is an Acid and what is a Base?

Acids and bases touch upon virtually all areas of chemistry,


biochemistry, and physiology. The concepts of an acid, a base, and a
salt are very old ones that have undergone several major refinements
as chemical science has evolved.

Acids
The term acid was first used in the seventeenth century; it comes from
the Latin root ac-, meaning sharp, as in acetum, vinegar. Some early
writers suggested that acidic molecules might have sharp corners or
spine-like projections that irritate the tongue or skin.
Acids have long been recognized as a distinctive class of compounds
whose aqueous solutions

taste sour
turn blue litmus to red,
react with active metals to give hydrogen gas
react with carbonates to produce a salt, water and CO2.

Acids and the hydrogen ion


The key to understanding acids (as well as bases and salts) had to
await Michael Faradays mid-nineteenth century discovery that
solutions of salts (known as electrolytes) conduct electricity. This
implies the existence of charged particles that can migrate under the
influence of an electric field. Faraday named these particles ions
(wanderers). Later studies on electrolytic solutions suggested that
the properties we associate with acids are due to the presence of an
excess of hydrogen ions in the solution. By 1890 the Swedish chemist
Svante Arrhenius (1859-1927) was able to formulate the first useful
theory of acids.

Arrhenius definition of acids


An acid is a substance that releases H+ in water.

300
We know now that H+ does not exist in solution as such; it exists
associated with H2O molecules, mostly as H3O+. This species is called
hydronium ion. For convenience sometimes it is still written as
H+(aq), but it is to be understood that it is the hydronium ion that is
being referred to.
Thus HCl(g) and HNO3 (l) which are molecular substance, ionize in
water to give hydronium ion.
HCl(g) + H2O H3O+ (aq) + Cl(aq)
HNO3 (l) + H2O H3O+ (aq) + NO3 (aq)
In general, for the acid molecule HA, we can write the ionization
process as:
HA + H2O H3O+ (aq) + A(aq), or simply as
HA H+(aq) + A- (aq)
There are three important points to understand about hydrogen in
acids. Although all Arrhenius acids contain hydrogen, not all
hydrogen atoms in a substance are capable of dissociating; thus the
CH3 hydrogens of acetic acid are non-acidic.
Those hydrogens that do dissociate can do so to different degrees. The
strong acids such as HCl and HNO3 are effectively 100% dissociated
in solution. Most organic acids, such as acetic acid, are weak; only a
small fraction of the acid is dissociated in most solutions. HF and
HCN are examples of weak inorganic acids.
Acids that possess more than one dissociable hydrogen atom are
known as polyprotic acids; H2SO4 and H3PO4 are well-known
examples. Intermediate forms such as HPO42, being capable of both
accepting and losing protons, are called ampholytes.

Hydrogen ions cannot exist in water


The hydrogen ion is no more than a proton, a bare nucleus. Although
it carries only a single unit of positive charge, this charge is
concentrated into a volume of space that is only about a hundred-
millionth as large as the volume occupied by the smallest atom.
(Think of a pebble sitting in the middle of a sports stadium!) The
resulting extraordinarily high charge density of the proton strongly
attracts it to any part of a nearby atom or molecule in which there is an
excess of negative charge. In the case of water, this will be the lone
301
pair (unshared) electrons of the oxygen atom; the tiny proton will be
buried within the lone pair and will form a shared-electron
(coordinate) bond with it, creating a hydronium ion, H3O+. Owing to
the overwhelming excess of H2O molecules in aqueous solutions, a
bare hydrogen ion has no chance of surviving in water.
The equation "HA H+ + A" is so much easier to write that
chemists still use it to represent acid-base reactions in contexts in
which the proton donor-acceptor mechanism does not need to be
emphasized. Thus it is permissible to talk about hydrogen ions and
use the formula H+ in writing chemical equations as long as you
remember that they are not to be taken literally in the context of
aqueous solutions.

Bases
The name base has long been associated with a class of compounds
whose aqueous solutions are characterized by:
a bitter taste;
a soapy feeling when applied to the skin;
ability to restore the original blue color of litmus that has been
turned red by acids;
ability to react with acids to form salts.
react with certain metals to produce gaseous H2;

The word alkali is often applied to strong inorganic bases. It is of


Arabic origin, from al-kali ("the ashes") which refers to the calcined
wood ashes that were boiled with water to obtain potash which
contains the strong base KOH, used in soap making. The element
name potassium and its symbol K (from the Latin kalium) derive from
these sources.

Arrhenius definition of a base


Just as an acid is a substance that liberates hydrogen ions into
solution, a base yields hydroxide ions when dissolved in water:
NaOH(s) Na+ (aq) + OH(aq)
Sodium hydroxide is an Arrhenius base because it contains hydroxide
ions. However, other substances which do not contain hydroxide ions

302
can nevertheless produce them by reaction with water, and are
therefore also classified as bases. Two classes of such substances are
the metal oxides and the hydrogen compounds of certain nonmetals:
Na2O(s) + H2O [2 NaOH] 2 Na+(aq) + 2 OH(aq)
NH3 + H2O NH4+(aq) + OH(aq)
In general, if we represent a base by B, the process is:
B + H2O B+ (aq) + OH(aq)

Brnsted and Lowry definition of acids and bases


Arrhenius definition did not include reactions such as that between
NH3(g) and HCl(g) as being acid-base reactions. The reaction:
NH3(g) + HCl(g) NH4Cl(s)
takes place without the present of water and without any OH being
involved. If NH3 is to be classified as a base in this reaction, there has
to be another definition of a base.
Brnsted and Lowry independently extended the definition of acids
and bases to cover systems where there is no water.
According to Brnsted and Lowry definition:
An acid is a substance that gives a proton (H+) to another substance
A base is a substance that accepts a proton.
This definition covers all the substances that fall under the Arrhenius
classification in addition to other substances. Thus the HCl-NH3
reaction mentioned above is an acid-base reaction according to this
definition.
When HCl(g) dissolves in water the following reaction takes place.
HCl(g) + H2O(l)H3O+(aq) + Cl-(aq)
Here HCl is an acid according to both the definitions, because it gives
H+. But in addition, H2O acts as a Brnsted-Lowry base in the
reaction, since it accepts an H+.
Similarly in the reaction
NH3(aq) + H2O NH4+ (aq) + OH(aq)

303
in addition to NH3 being a base (as it is also according to the
Arrhenius definition), H2O is an acid according to Brnsted and
Lowry definition, since it gives an H+.

Conjugate Acid-Base Pairs


Any reaction in which there is a transfer of H+ is an acid-base reaction.
So there must be an acid and a base taking part in the reaction. In the
general reaction (written as an equilibrium, because both reactants and
products might be present in significant amounts),
HA(aq) + H2O(l) A (aq) + H3O+(aq)
In the forward reaction HA donates a proton to H2O. So, HA is an
acid. H2O accepts a proton so it is a base. In the reverse reaction H 3O+
gives a proton to A-, while A- accepts that proton. Hence A- is a base
and H3O+ is an acid. Thus HA and A are an acid-base pair and H2O
and H3O+ the other acid-base pair. Such pairs are called conjugate
acid-base pairs. Conjugate base of HA is A- and the conjugate acid of
H2O is H3O+..
Some common conjugate acid-base pairs
Acid Base
Ethanoic acid; HC2H3O2 Ethanoate ion: C2H3O2
Hydrogen sulfate ion, HSO4 Sulfate ion: SO42
Ammonium ion, NH4+ Ammonia: NH3
Water, H2O Hydroxide ion: OH
Hydronium ion, H3O+ water: H2O
Hydrogen carbonate, HCO3 Carbonate ion: CO32

Relative Strengths of Acids and Bases


Some acids are better proton donors than others. Acids and bases can
be broadly divided into categories: strong and weak.
Strong acids are those that are fully ionized in water, leaving no
undissociated acid in the solution. Similarly strong bases are
completely ionized. Weak acids and bases are only partially ionized.

304
Examples of strong acids are: HNO3, H2SO4, HCl, HI, HBr, HClO4.
Examples of strong bases are alkali metal hydroxides, such as NaOH
and KOH.
Weak acids are partially dissociated in water. Examples are HNO2,
HF, CH3COOH, and H3PO4. Examples of weak bases are NH3,
CH3NH2, CH3NHC2H5 and other amines, whether primary, secondary
or tertiary.
The relative strength of weak acids depends on how much to the right
the following equilibrium lies:
HA(aq) + H2O(l) A (aq) + H3O+(aq). The equilibrium constant for
such a reaction is denoted by the symbol Ka, and is called acid
dissociation constant.
The larger the Ka for an acid, stronger it is. Similarly the equilibrium
constant for a base:
NH3(aq)+ H2O(l) NH4+(aq) + OH(aq) is given the symbol Kb, and
called a base dissociation constant.
Larger the kb for a base, stronger it is.

15.2 Acidity of a Solution: pH

The Autoionization of Water


We have seen that water can act as an acid and a base depending on
whether it is in the presence of an acid or a base. One water molecule
can donate a proton to another water molecule as well.
H2O(l) + H2O(l) OH (aq) + H3O+ (aq)
This is self-ionization or autoionization of water. The equilibrium
constant is low, so only about 2 out of every 10 9 molecules are ionized
at any given time. Even then this self-ionization is very important.
The equilibrium constant of the above reaction is given a special
symbol, Kw:
Kw = [H3O+][OH] = 1.0 x 10-14 at 25 oC.
This equilibrium can also be written as:
H2O(l) H+(aq) + OH(aq)
Kw = [H+][OH] = 1.0 x 1014 at 25 oC.

305
A solution in which [H+] = [OH] is said to be neutral. In a neutral
solution, [H+] = [OH] = 1.0 x 107 at 25oC. This is applicable to all
aqueous solutions, not just to pure water. If in a solution [H +] > [OH],
it is acidic; if [H+] < [OH] it is basic.

pH
When dealing with a range of values (such as the variety of hydrogen
ion concentrations encountered in chemistry) that spans many powers
of ten, it is convenient to represent them on a more compressed
logarithmic scale. By convention, we use the pH scale to denote
hydrogen ion concentrations: pH = log10 [H+]; or conversely [H+] =
10pH.
This notation was devised by the Danish chemist Sren Srensen
(1868-1939) in 1909. There are several accounts of why he chose
"pH"; a likely one is that the letters stand for the French term pouvoir
hydrogne, meaning "power of hydrogen" "power" in the sense of
an exponent. It has since become common to represent other small
quantities in "p-notation". Two that you need to know are the
following:
pOH = log10 [OH]
pKw = log Kw (= 14 when Kw = 1.00 1014)
Note that pH and pOH are expressed as numbers without any units,
since logarithms must be dimensionless.
Recall that [H+][OH] = 1.00 1014.
Taking negative logarithms of both sides in the last equation, we find
that:
pKa + pKb = pKw = 14.00 at 25oC.
In a neutral solution at 25C, pH = pOH = 7.0. As pH increases, pOH
diminishes; a higher pH corresponds to an alkaline solution, a lower
pH to an acidic solution. In a solution with [H+] = 1 M, the pH would
be 0; in a 0.00010 M solution of H+, it would be 4.00. Similarly, a
0.00010 M solution of NaOH would have a pOH of 4.00, and thus a
pH of 10.00. A change is the concentration by a factor of 10 causes
the pH to change by 1.

306
Example 15.1:
The pH of blood must be held very close to 7.40. Find the hydroxide ion
concentration that corresponds to this pH.
Solution:
pOH = 14.007.40 = 6.60
[OH] = 10pOH = 106.6 = 2.51 x 107 M

The pH scale

The range of possible pH values runs from about 0 to 14. The word
"about" in the above statement reflects the fact that at very high
concentrations (10 M hydrochloric acid or sodium hydroxide, for
example) a significant fraction of the ions will be associated into
neutral pairs such as H+Cl, thus reducing the concentration of
available ions to a smaller value which we will call the effective
concentration. It is the effective concentration of H+ and OH that
determines the pH and pOH.
For solutions in which ion concentrations don't exceed 0.1 M, the
formulas pH = log [H+] and pOH = log [OH] are generally reliable,
but don't expect a 10.0 M solution of a strong acid to have a pH of
exactly 1.00!

307
Figure 15.1: pH scale
Figure 15.1 gives a general feeling for where common substances fall
on the pH scale. Notice especially that
most foods are slightly acidic;
the principal "bodily fluids" are slightly alkaline, as is seawater
not surprising, since early animal life began in the oceans.
the pH of freshly-distilled water will drift downward as it takes
up carbon dioxide from the air; CO2 reacts with water to produce
carbonic acid, H2CO3.
the pH of water that occurs in nature varies over a wide range.
Groundwater often picks up additional CO2 respired by
organisms in the soil, but can also become alkaline if it is in
contact with carbonate-containing sediments. "Acid" rain is by
definition more acidic than pure water in equilibrium with

308
atmospheric CO2, owing mainly to nitric and sulfuric acids that
originate from fossil-fuel emissions of nitrogen oxides and SO2.

pH indicators

The colors of many dye-like compounds depend on the pH, and can
serve as useful indicators to determine whether the pH of a solution is
above or below a certain value.
Universal indicators
Most indicator dyes show only one color change, and thus are only
able to determine whether the pH of a solution is greater or less than
the value that is characteristic of a particular indicator. By combining
a variety of dyes whose color changes occur at different pH values, a
"universal" indicator can be made. Commercially-prepared pH test
papers of this kind are available for both wide and narrow pH ranges.
pH can be measured using a pH meter. Acid-base indicators can be
used to measure pH as well. Indicators give different colours in
solutions of different pH values.

15.3 Calculations involving Ka and pH

Calculating Ka from pH
Ka can be calculated from pH by solving equilibrium problems using
techniques learned in the chapter on equilibrium.
Example 15.2:
A 0.030 M solution of ethanoic acid has pH = 3.13 at 25 oC. Calculate Ka
of ethanoic acid.
Solution:
Represent the acid by HA; for the ionization, we can write:
HA(aq) + H2O(l) A (aq) + H3O+(aq).
HA(aq) H+(aq) + A (aq).
Ka = [H+] [A] / [HA]
Since pH = -log[H+] ; [H+] = antilog [-pH] = 103.13 = 7.41 x 104M
Let us make an equilibrium table:

309
HA(aq) H+(aq) + A- (aq).
Initial 0.030M 0 0
Change 7.41 x 10-4M 7.41 x 10-4M 7.41 x 10-4M
Equil. 0.030- 7.41 x 10-4M 7.41 x 10-4M 7.41 x 10-4M

We now make the approximation 0.030 - 7.41 x 10-4 M 0.030 M,


since to the number of significant figures we are using, the answer on
subtraction is 0.030.
Ka = (7.41 x 10-4) (7.41 x 10-4) / 0.030 = 1.8 x 105
Percent Ionization
Sometimes percent ionization is used to indicate the relative strengths
of acids and bases.
% ionization:
= (concentration ionized / original concentration) x 100 %
= {[H+]/ [HA]} x 100 %

Example 15.3:
What is the percent ionization of ethanoic acid in the solution in the
example above?
% ionization:
= {[H+]/ [HA]} x 100 = {7.41 x 10-4M/0.030M} x 100 % = 2.5%
The stronger the acid, the greater the percent ionization.
Base dissociation constants of some weak bases at 25 oC
BASE FORMULA Kb
(CH3)2NH dimethylamine 9.6 x 104
CH3NH2 methylamine 4.4 x 104
(CH3)3N trimethylamine 7.4 x 105
NH3 ammonia 1.8 x 105
N2H4 hydrazine 9.2 x 107
C5H5N pyridine 1.5 x 109
C6H5NH2 aniline 4.1 x 1010

310
Acid dissociation constants of some weak acids at 25 oC

ACID FORMULA Ka
Benzoic C6H5COOH 6.28x10-5
1-Butanoic CH3CH2CH2COOH 1.52x10-5
Chloroacetic ClCH2COOH 1.36x10-3
Ethanoic acid CH3COOH 1.75x10-5
Hydrogen Cyanide HCN 6.2x10-10
Hydrofluoric HF 6.8x10-4
Hypochlorous HOCl 3.0 x 10-8
Lactic CH3CHOHCOOH 1.38x10-4
Methanoic HCOOH 1.80x10-4
Nitrous HNO2 7.1 x 10-4
Propanoic CH3CH2COOH 1.34x10-5
Trichloroacetic Cl3CCOOH 3
Using Ka to calculate pH

If we know Ka of a weak acid and its initial concentration, we can


calculate the pH of the solution. An example of how it is done
follows.
Example 15.4:
Calculate pH of a 0.40M solution of methanoic acid, HCOOH.
Ka = 1.8 x 104 at 25oC.
Solution:
HCOOH(aq) H+(aq) + HCOO(aq)
Ka = [H+][HCOO]/[HCOOH] = 1.8 x 10-4

Equilibrium table:
HCOOH (aq) H+(aq) + HCOO-(aq)
Initial 0.40 M 0 0
Change -x +x +x
Equilibrium 0.40-x x x

311
Ka = x . x / (0.40 x) = 1.8 x 10-4
Now we make an assumption (the validity of which will be tested later):
0.40 x 0.40
Ka = x . x / 0.40 = 1.8 x 10-4;
x2 = (1.8 x 10-4)(0.40) ;
hence x = (1.8 x 10-4)(0.40) =8.49 x 10-3 = [H+]
pH = -log[H+] = -log(8.49 x 10-3) =2.07
Checking the validity of the assumption:
% ionization = 0.00849/0.40 x 100% = 2.12%.
If x is less than 5% of the initial concentration, it is permissible to ignore
it compared to the initial concentration; otherwise the quadratic formula
should be used. The assumption made in the above example is thus valid.

Polyprotic Acids
These are the acids that have more than one ionizable atom. For
example for H3PO4, phosphoric acid, we have
H3PO4(aq) H+(aq) +H2PO4-(aq) Ka1 = 7.5 x 10-3
H2PO4-(aq) H+(aq)+ HPO42-(aq) Ka2 = 6.2 x 10-8
HPO42-(aq) H+(aq) + PO43-(aq) Ka3 = 4.2 x 10-13

Here Ka2 and Ka3 are much smaller than Ka1 since it is more difficult
to remove a proton from a charged species than a neutral one. It
becomes progressively more difficult to remove subsequent protons
after the first one because of the charge on the remaining ion.
For most polyprotic acids Ka1 is much larger than subsequent Ka
values so only the first ionization constant is considered and the acid
is treated as if it is monoprotic.

Weak Bases
Weak bases take part in the equilibrium:
B(aq) + H2O(l) HB+(aq) + OH(aq)
Kb = [HB+][OH]/[B]
Example is: NH3(aq)+ H2O(l) NH4+(aq) + OH(aq) ;
312
Kb= [NH4+][OH]/[NH3]
There are two types of weak bases. One is those containing N with a
non-bonding pair that can accept a proton. These include NH3 and
organic amines, such as C2H5NH2.The other class is the anions of
weak acids, such as BrO (the conjugate base of HBrO). Thus an
aqueous solution of the salt KBrO dissociates to give BrO and K+
ions. K+ is a spectator ion in acid-base reactions, but BrO acts as a
base in water.
BrO(aq) + H2O(l) HBrO (aq) + OH(aq) Kb = 4.0 x 10-6

Relationship between Ka and Kb


Let us consider the acid-base pair, NH3(aq) and NH4+(aq). The
equilibrium of each of these is:
NH3(aq) + H2O(l) NH4+(aq) + OH (aq);
Kb = [NH4+][OH]/[NH3]
and NH4+(aq) + H2O(l) NH3(aq) + H3O+(aq).
Or, in simpler form: NH4+(aq) NH3(aq) + H+(aq).
Ka = [NH3][H+]/[NH4+]
If we multiply Ka with Kb, we get:
Ka x Kb ={ [NH3][H+]/[NH4+]} x {[NH4+][OH]/[NH3]}
= [H+][OH] = Kw
So, Ka x Kb = Kw
This relation shows that if Ka is large, Kb will be small. In other
words, a stronger acid has a weaker conjugate base; a weaker acid has
a stronger conjugate acid.

15.4 Titrations

Since acids and bases readily react with each other, it is


experimentally quite easy to find the amount of acid in a solution by
determining how many moles of base are required to neutralize it.
This operation is called titration, and you should already be familiar
with it from your work in the Laboratory.

313
We can titrate an acid with a base, or a base with an acid. The
substance whose concentration we are determining (the analyte) is the
substance being titrated; the substance we are adding in measured
amounts is the titrant. The idea is to add titrant until the titrant has
reacted with all of the analyte; at this point, the number of moles of
titrant added tells us the concentration of base (or acid) in the solution
being titrated.

Example 15.5:
36.00 ml of a solution of HCl was titrated with 0.44 M KOH. The volume
of KOH solution required to neutralize the acid solution was 27.00 mL.
What was the concentration of the HCl?
Solution:
The number of moles of titrant added was (0.027 L)(0.44 mol L1) =
0.0119 mol. Because one mole of KOH reacts with one mole of HCl, this
is also the number of moles of HCl; its concentration is therefore (0.0119
mol) (0.036 L) = 0.33 M.

Titration curves

The course of a titration can be followed by plotting the pH of the


solution as a function of the quantity of titrant added. Fig. 15.2 shows
such curves, one for a strong acid (HCl) and the other for a weak acid,
acetic acid, denoted by HAc. Looking first at the HCl curve, notice
how the pH changes very slightly until the acid is almost neutralized.
At that point, which corresponds to the vertical part of the plot, just
one additional drop of NaOH solution will cause the pH to jump to a
very high value almost as high as that of the pure NaOH solution.

314
Fig. 15.2: Titration curves of weak and strong acids titrated with OH

Compare the curve for HCl with that of HAc. For a weak acid, the pH
jump near the neutralization point is less steep. Notice also that the pH
of the solution at the neutralization point is greater than 7.

Figure 15.3: Titration


curves for a polyprotic
acid

If the acid or base is polyprotic, there will be a jump in pH for each


proton that is titrated. In the example shown in Fig. 15.3, a solution of
carbonic acid H2CO3 is titrated with sodium hydroxide. The first
equivalence point (at which the H2CO3 has been converted entirely
into bicarbonate ion HCO3) occurs at pH 8.3. The solution is now
identical to one prepared by dissolving an identical amount of sodium
hydrogen carbonate in water.
315
Addition of another mole equivalent of hydroxide ion converts the
hydrogen carbonate (bicarbonate) into carbonate ion and is complete
at pH 10.3; an identical solution could be prepared by dissolving the
appropriate amount of sodium carbonate in water.

Finding the equivalence point: indicators

When enough base has been added to react completely with the
hydrogens of a monoprotic acid, the equivalence point has been
reached. If a strong acid and strong base are titrated, the pH of the
solution will be 7.0 at the equivalence point. However, if the acid is a
weak one, the pH will be greater than 7; the neutralized solution
will not be neutral in terms of pH. For a polyprotic acid, there will
be an equivalence point for each titratable hydrogen in the acid. These
typically occur at pH values that are 4-5 units apart, but they are
occasionally closer, in which case they may not be readily apparent in
the titration curve.
The key to a successful titration is knowing when the equivalence
point has been reached. The easiest way of finding the equivalence
point is to use an indicator dye. One such indicator that is commonly
encountered in the laboratory is phenolphthalein; it is colorless in
acidic solution, but turns intensely red when the solution becomes
alkaline. If an acid is to be titrated, you add a few drops of
phenolphthalein to the solution before beginning the titration. As the
titrant is added, a local red color appears, but quickly dissipates as the
solution is shaken or stirred. Gradually, as the equivalence point is
approached, the color dissipates more slowly; the trick is to stop the
addition of base after a single drop results in a permanently pink
solution.
Different indicators change color at different pH values. Since the pH
of the equivalence point varies with the strength of the acid being
titrated, one tries to fit the indicator to the particular acid. One can
titrate polyprotic acids by using a suitable combination of several
indicators, as is illustrated above for carbonic acid.

316
15.5 Exercises
15.1 Calculate [OH-] and pH for (a) 1.5 x 10-3 M Ba(OH)2 (b) 1.380 g of
NaOH in 200.0 mL of solution, (c) 2.00 mL of 0.235 M KOH
diluted to 2.00 L.
15.2 Calculate [OH-] and pH for a solution formed by adding 10.00 mL
of 0.120 M KOH to 15.0 mL of 8.5 x 10-3 M Sr(OH)2.
15.3 Calculate the pH and pOH of 1.0 x 10-3 M HClO4(aq).
15.4 A 0.20 M solution of chloroacetic acid, HC2H2O2Cl, has a pH =
3.56. Calculate Ka.
15.5 Ka for hydrazoic acid, HN3, is 1.8 x 10-5. Calculate equilibrium
concentrations of H3O+, N3-, and HN3 if the initial concentration of
HN3 is 0.040.
15.6 Calculate the molar concentration of OH- ions in a 0.0030 M
solution of dimethylamine (CH3)2NH (Kb = 9.6 x 10-4). What is the
pH of this solution?
15.7 The percentage protonation of the base octylamine in a 0.100 M
aqueous solution is 6.7%. What is the pH of the solution and the Kb
of the base?
15.8 Calculate the percentage ionization of 0.0045 M chloroacetic acid
(Ka = 1.36 x 10-3)
15.9 Calculate the pH of 0.25 M HCN (aq). Ka for HCN = 6.2x10-10 at
25 oC.

317
CHAPTER 16

Aqueous Equilibria
16.1 pH of Salt Solutions

Solutions of salts
In the previous chapter we came across acidic ions that are conjugate
acids of weak bases, such as NH4+ and CH3NH3+. Examples of basic
ions, which are conjugate bases of weak acids, are ClO, F, BrO,
CN, CO32, CH3COO. Solutions of salts could be acidic, basic or
neutral depending on the aqueous ions they provide. It has to be
emphasized that salts have ions even in solid state. NH4Cl has NH4+
and Cl ions. In solutions these ions take part in equilibrium with
water.

NH4Cl gives NH4+ (aq) and Cl(aq) ions. We recognize NH4+(aq) as


being acidic from the fact that it is the conjugate acid of the weak base
NH3, according to the equilibrium:
NH3(aq) + H2O NH4+(aq) + OH(aq)
Being an acid, NH4+ gives H+ ions in solution by the equilibrium:
NH4+(aq) + H2O NH3(aq) + H3O+(aq)
The other ion in NH4Cl, the chloride ion, does not take part in any
equilibrium in aqueous solution. We can recognize this if we realize
that HCl being a strong acid, is fully ionized in water. In other words
the equilibrium:
HCl + H2O H3O+(aq) + Cl-(aq)
lies entirely to the right, and there is no tendency by Cl to react with
H3O+ to give back HCl. Hence Cl is neutral.
This argument makes it apparent that a solution of NH4Cl will be
acidic and will have pH less than 7.
A salt solution is acidic, basic or neutral depending on what acid and
base the salt is derived from. See Table 16.1.

318
Table 16.1: Salt solution and pH

Salt derived from Example pH Reaction


strong acid + strong base KCl 7 None
weak acid + strong base NaF >7 F + H2O HF + OH
strong acid + weak base NH4Cl < 7 NH4+ + H2O NH3 + H3O+
depends on the relative
weak acid + weak base NH4F ? strengths of ions as acid and
bases.

Metal cations as acids


When sodium chloride is dissolved in pure water, the pH remains
unchanged because neither ion reacts with water. However, a solution
of magnesium chloride will be faintly acidic, and a solution of
iron(III) chloride FeCl3 will be distinctly so. How can this be? Since
none of these cations contains hydrogen, we can only conclude that
the protons come from the water.
The water molecules in question are those that find themselves close
to any cation in aqueous solution; the positive field of the metal ion
interacts with the polar H2O molecule through ion-dipole attraction,
and at the same time increases the acidity of these loosely-bound
waters by making facilitating the departure H+ ion. In general, the
smaller and more highly charged the cation, the more acidic will it be;
the acidity of the alkali metals and of ions like Ag+(aq) is negligible,
but for more highly-charged ions such as Mg2+, Pb2+ and Al3+, the
effect is quite noticeable.
Most of the transition-metal cations form
organized coordination complexes in
which four or six H2O molecules are
chemically bound to the metal ion where
they are well within the influence of the
electric field of the cation, and thus
subject to losing a proton. Thus an
aqueous solution of "Fe3+" is really a
solution of the ion hexaaquo iron III,

319
whose first stage of "dissociation" can be represented as Fe(H2O)63+ +
H2O Fe(H2O)5(OH)2+ + H3O+
As a consequence of this reaction, a solution of FeCl 3 turns out to be a
stronger acid than an equimolar solution of acetic acid. A solution of
FeCl2, however, will be a much weaker acid; the +2 charge is
considerably less effective in easing the loss of the proton.
It should be possible for a hydrated cation to lose more than one
proton. For example, an Al(H2O)63+ ion should form, successively, the
following species:
AlOH(H2O)52+ Al(OH)2(H2O)4+ Al(OH)3(H2O)30
Al(OH)4(H2O)2 Al(OH)5(H2O)2 Al(OH)63
However, removal of protons becomes progressively more difficult as
the charge decreases from a high positive value to a negative one; the
last three species have not been detected in solution. In dilute
solutions of aluminum chloride the principal species are actually
Al(H2O)63+ (commonly represented simply as Al3+) and AlOH(H2O)52+
("AlOH2+").

Example 16.1:
What is the pH of 0.20 NH4Br(aq)? Ka for NH4+ = 5.6 x 10-10
Solution:
The solution has the ions NH4+ and Br. The former is acidic and
takes part in the equilibrium with water. Br is derived from a strong
acid HBr, and so it is neutral and does not react with water. The
equilibrium is:
NH4+(aq) + H2O(l) NH3(aq) + H3O+(aq).
Ka = [NH3] [H3O+] /[NH4+]
Equilibrium table:
NH4+(aq) + H2O(l) NH3(aq) + H3O+(aq).
Start 0.20
Change -x +x +x
Eq. conc 0.20-x x x
Ka = x2/(0.20 x) = (5.6)(10-10)

320
Make the approximation, 0.20 x 0.20, since x is expected to be
very small as Ka is small.
Ka= x2/0.20 = (5.6)(10-10); solve for x
x = 1.1 x 10-6 = [H3O+]
pH = -log(1.06 x 10-5) = 4.97

Example 16.2:
Calculate the pH of 0.20 M solution of HCOOK.
Kb for HCOO- = 5.56 x 10-11
Solution:
The equilibrium table:
HCOO- (aq) + H2O HCOOH(aq) + OH- (aq)
Start 0.20
Change -x +x +x
Equil. 0.20 x x x

Kb = [HCOOH] [OH-] / [HCOO-] = x.x / (0.20 x) = 5.56 x 10-11


Approximation 0.20 x = 0.20; x2 = 0.20 x 5.56 x 10-11
x = 3.33 x 10-6 = [OH-]
pOH = -log(3.33 x 10-6 ) = 5.48; pH = 14.00 5.48 = 8.52

pH of a mixed solution
What is pH of a solution containing mixture of a weak acid and its
conjugate base or a mixture of weak base and its conjugate acid? The
procedure for calculating the pH is similar to the examples done
earlier, except that now starting concentration of the conjugate base
will need to be taken into account. The examples below will illustrate
this.

Example 16.3:
Calculate pH of a solution that is 0.20 M HNO2 and 1.5 x 10-3 M NaNO2.
Given Ka of HNO2 = 4.5 x 10-4
Solution:
The species taking part in the equilibrium are HNO2 and NaNO2. The

321
equilibrium table is:
HNO2(aq) + H2O H3O+(aq)+ NO2(aq)
Start 0.20M 1.5 x 10-3 M
Change -x +x +x
Equil. 0.20 x x 1.5 x 10-3+ x

Ka = [H3O+] [NO2] / [HNO2] = (x) (1.5 x 10-3+ x )/(0.20 x)


= 4.5 x 10-4
Make the approximation 0.20 x = 0.20
(x) (1.5 x 10-3+ x ) /0.20 = 4.5 x 10-4. This equation can be solved by the
quadratic formula:
x2 + 1.5 x 10-3x 9.0 x 10-5 = 0
x = [- 1.5 x 10-3 {(1.5 x 10-3)2 + 4(9.0 x 10-5)}]/2
x = 8.77 x 10-3 M = [H3O+]
pH = 2.06

16.2 Buffers

Buffers are solutions that resist change in pH on addition of a small


amount of acid or base. They are a mixture of comparable amounts of
a weak acid and its conjugate base; or of a weak base and its conjugate
acid. An example is a solution having 0.1 M each in HClO and ClO -
(as NaClO, for example) or a solution of NH3 and NH4Cl in roughly
equal amounts.
How do buffers resist change in pH?
Consider a buffer system containing 0.1 M each of NH3 and NH4Cl
buffer. The equilibrium is:
NH3(aq) + H2O(l) NH4+ (aq) + OH-(aq)
If to this solution a small amount of OH- gets added, the added OH-
tends to get used up to shift the equilibrium to the left in accordance
with Le Chateliers principle, keeping pH unchanged. If some acid
gets added, it is removed by reaction with the OH- present. The loss
of OH- is made up by the equilibrium shifting to right in accordance
with Le Chateliers principle. Thus any large change in pH is resisted.

322
pH of buffers
pH of a buffer solution can be calculated by first setting up an
equilibrium table and proceeding as in the following example.

Example 16.4:
Calculate pH of a buffer solution comprising 0.10 M HNO2 and 0.15 M
NaNO2. Given Ka of HNO2 = 4.5 x 10-4
Solution:
The species taking part in the equilibriumareHNO2 and NaNO2. The
equilibrium table is:
HNO2(aq) + H2O H3O+(aq) + NO2(aq)
Start 0.10M 0.15 M
Change -x +x +x
Equil. 0.10 x x 0.15+ x
Ka = [H3O+] [NO2] / [HNO2] = (x) (0.15+ x ) / (0.10 x) = 4.5 x 10-4
Make the approximation 0.10 x = 0.10; 0.15 + x = 0.15, by neglecting x
compared to 0.10 and 0.15.
(x) (0.15 ) / 0.10 = 4.5 x 10-4
[H3O+] =x = (4.5 x 10-4) (0.10/0.15)
pH = -log [H3O+] = 3.52.

Henderson-Hasselbalch Equation
Often a quick way of calculating pH is to use the formula derived
below using the above example of a buffer.
Ka = [H3O+] [NO2] / [HNO2]
Taking negative logarithms of both sides,
-logKa = -log [H3O+]- log( [NO2] / [HNO2])
pKa = pH - log( [NO2] / [HNO2])
pH = pKa + log([NO2] / [HNO2])
In general pH = pKa + log([base] / [acid]), which is the Henderson-
Hasselbalch equation.

323
16.3 Solubility Equilibria

Examples of sparingly soluble salts coming to equilibrium with their


aqueous solutions are:
AgCl(s) Ag+ (aq) + Cl-(aq)
The equilibrium constant of such reactions is called the solubility
product and given the special symbol Ksp.
Ksp = [Ag+] [Cl-] = 1.6 x 10-10 for AgCl at 25 oC.
Another example:
Sb2S3(s) 2Sb3+ (aq) + 3S2-(aq)
Ksp = [Sb3+]2 [S2-]3 = 1.7 x 10-93

Determination of Ksp from molar solubility


If the solubility of a sparingly salt is known its Ksp can be calculated as
shown in the exercise below.
Example 16.5:
The molar solubility of CaF2 is 2.2 x 10-4 M at 25 oC. Calculate Ksp
of the salt.
Solution:
The equilibrium is CaF2(s) Ca2+ (aq) + 2F- (aq).
Assume the molar solubility is S. Then [Ca2+] = S; and [F-] = 2S
Ksp = [Ca2+] [F-]2 = (S) (2S)2 = 4S3 = 4(2.2 x 10-4 )3 =4.3 x 10-11.

Determination of molar solubility from Ksp


If Ksp is known, molar solubility can be calculated from the
equilibrium equation.
Example 16.6:
Ksp of Fe(OH)2 is 1.6 x 10-14. What is the molar solubility of Fe(OH)2?
Solution:
Fe(OH)2 (s) Fe+2 (aq) + 2OH- (aq)
Assume the solubility is S; [Fe2+] = S and [OH-]= 2S.
Ksp= [Fe2+] [OH-]2 = (S) (2S)2 = 4S3 = 1.6 x 10-14

324
S = (1.6 x 10-14)1/3 /4 = 6.3 x 10-6 M.

Determination of molar solubility from Ksp.


The solubility of a sparingly soluble salt can be calculated from a
known value of solubility product. The following exercise illustrates
this.
Example 16.7:
Calculate the solubility of Ca3(PO4)2. Ksp = 1.2 x 10 -26
Solution:
Ca3(PO4)2 (s) 3Ca2+ (aq) + 2PO4-3(aq)
Ksp = [Ca2+]3[PO4-3]2 = (3S)3 (2S)2 = 108 S5
S5 = (1.2 x 10 -26/108);
S = 2.6 x 10-6 M

The Common-ion effect


A consequence of Le-Chateliers principle as applied to solubility
equilibrium involving a sparingly soluble salt is that the solubility of
the salt decreases in a solution in which there is one of the constituent
ions of the salt from another source. Consider the case of Fe(OH)2
treated in the preceding exercise. If the solution contains OH- from
another source, the equilibrium shifts to the left and the solubility
decreases. The solubility of Fe(OH)2 in water was found to be 6.3 x
10-6 M.
Let us calculate the solubility of Fe(OH)2 in a 0.0010 M OH- solution.
Ksp= [Fe2+] [OH-]2 = [Fe2+] (0.0010)2 = 1.6 x 10-14
[Fe2+] = 1.6 x 10-14 / (0.0010)2 = 1.6 x 10-11 M
The solubility in the solution containing the common ion OH- is
drastically lower than in water. This is only a rough estimate, because
of the interactions of ions in the more concentrated solution.
Will a precipitate form?
The value of Ksp can be used to predict if a precipitate would form
when two types of ions are mixed at certain concentrations. Here we
use our knowledge of how to predict direction of a reaction using Q c
and Kc dealt with in the chapter on chemical equilibrium.
325
Suppose solutions of Ca(NO3)2 and Na2SO4 are mixed such that the
instant they are mixed, there concentration is 0.010 M each. Will a
precipitate of CaSO4 form? In other words will the following reaction
go to the right or left?
Ca2+ (aq) + SO42- (aq) CaSO4 (s)
Let us write the reaction as:
CaSO4 (s) Ca2+ (aq) + SO42- (aq), Ksp =[Ca2+] [SO42-] = 2.4x10-5
Using the actual concentrations, we find that Qsp = (0.010)(0.010) =
1.0 x 10-4
Since Qsp > Ksp, we can conclude that the reaction will go to the left
and a precipitate of CaSO4 will form.

16.4 Exercises

16.1 Arrange the following 0.10 M solutions in order of increasing


pH: (i) NH4Cl (ii) KCl, (iii) HCOONa, (iv) KNO2,
(v) CH3COOK.

16.2 Calculate the pH of 0.25 M KCN(aq). Ka for HCN = 6.2 x 10-10


at 25 oC.

16.3 Arrange the following acids in order of increasing strength


using the data for their conjugate bases given below: HNO2,
HCN, NH4+, C6H5COOH.
The following pKb values are given. NO-2 (10.8), CN- (4.79),
NH3(4.75), C6H5COO- (9.80).

16.4 Calculate the pH of a solution that is 0.15 M (CH3)2NH and


0.10 M (CH3)2NH2+Cl-. Given Kb of (CH3)2NH = 9.6 x 10-4

16.5 Calculate Ksp for MgF2; given its solubility is 1.2 x 10-3 mol/L.

16.6 Find the solubility of lead(II) chloride in 0.10 M CaCl 2 (aq).


Ksp = 1.6 x 10-5 for PbCl2.

16.7 Will a precipitate form in a solution containing 2.5 x 10 -3 M


Pb(NO3)2 and 4.3 x 10-3 M NaI? Ksp of PbI2 = 7.9 x 10-9.

326
16.8 What is the molar solubility of CaSO4 in a 0.015 M solution of
K2SO4? Ksp = 2.45 x 10-5 for CaSO4.

16.9 Calculate the pH of a buffer solution containing 0.25 M


NaCOOH and 0.19 M HCOOH.
For HCOOH, Ka = 1.80 x 10-4.

16.10 It is desired to prepare a solution buffered at pH of 8.00 in a


lab. How many moles of NH4Cl should be added to 1.0 L of a
solution that has 1.0 mol NH3 to make the required buffer.

327
INDEX
A
accuracy, 19, 20
Acid, 38, 65, 300, 304, 308, 309
Acid dissociation constants, 311
actinides, 28, 117, 118
activation energy, 273, 274
Alcohols, 185, 191
Aldehydes, 191
aliphatic, 165
alkali metals, 28, 69, 319
alkaline earth metals, 28, 69
Alkanes, 165, 167, 169, 173, 174
Alkenes, 173, 178, 180, 181
Amides, 193
Amines, 193
amino acids, 158
ammonia, 30, 39, 54, 60, 61, 67, 68, 76, 79, 193, 228, 254, 276, 296,
297, 310
ammonium ion, 33, 35, 37, 62
amu, 23, 26, 40, 45, 46, 48
angular momentum, 104
anions, 31, 32, 33, 34, 35, 36, 63, 78, 129, 131, 313
antimony, 29
Arenes, 189
aromatic hydrocarbons, 165, 189
Arrhenius, 268, 269, 300, 301, 302, 303, 304
Arrhenius equation, 268
arsenic, 29
atomic mass unit, 23
Atomic number, 24, 39, 40
Atomic radii, 119
atomic spectrum, 98
atomic weight, 26, 27, 40
Atoms, 23, 24, 95, 119, 190
Aufbau Principle, 107, 111
Avogadros Law, 84
328
Avogadros number, 46, 49, 50, 53, 57

B
Balancing chemical reaction equations, 42
Balancing redox reactions, 71
Balmer series, 99
Base dissociation constants, 310
Bases, 66, 300, 302, 304, 312
benzene, 30, 172, 180, 189, 190, 191, 240, 247
Bohr model, 100, 103, 112
Boiling point elevation, 237, 238
Bomb calorimeter, 201
bond enthalpy, 213, 214, 215
Bond enthalpy, 213
Bond lengths, 163
bond strength, 163
boron, 29, 40, 109
Boyles Law, 81
Brackett, 99
Brnsted and Lowry, 303, 304
Buffers, 257, 322
Butane, 43, 165

C
calorimeter constant, 201
Calorimetry, 200
carbohydrates, 158
carbon, 9, 10, 11, 20, 22, 29, 35, 38, 39, 42, 43, 50, 57, 58, 65, 87, 88,
93, 94, 109, 110, 134, 135, 136, 147, 149, 158, 160, 161, 162, 163,
164, 165, 166, 167, 168, 169, 171, 173, 174, 176, 177, 178, 179,
180, 181, 183, 184, 186, 188, 189, 191, 194, 206, 207,210, 211, 212,
238, 248, 268, 288, 299, 308
carbon dioxide, 42, 212
carbonate, 33, 35, 78, 80, 93, 138, 160, 281, 295, 304, 308, 315, 316
Carboxylic acids, 192
catalyst, 181, 182, 185, 187, 188, 266, 274, 293, 295, 297
catalysts, 171, 181, 190, 253, 274, 275
Catenation, 160
329
cations, 31, 32, 33, 34, 35, 62, 63, 70, 122, 123, 129, 319
Charless Law, 82
Chemical Bonding, 129
chemical equilibrium, 61, 278, 297, 325
Chemical properties, 12, 173, 177
chromium, 28, 39, 111
cobalt(III) chloride, 39
Coffee cup calorimeter, 200
Colligative properties, 251
Colligative Properties, 235
collision, 270
Collision, 270
Collision theory, 269
Combined gas law, 83
Combustion, 42, 173, 181
Common-ion effect, 325
compound, 8, 9, 11, 20, 22, 30, 33, 36, 37, 38, 39, 40, 41, 44, 45, 46,
49, 50, 51, 52, 53, 57, 58, 63, 64, 65, 66, 67, 114, 119, 129, 130,
139, 144, 158, 159, 161, 164, 165, 167, 178, 180, 186, 192, 194,
208, 210, 232, 242
Concentration, 219, 225, 253
concentration units, 223
Conjugate Acid-Base Pairs, 304
continuous spectra, 98
cooking gas, 171
copper, 20, 24, 26, 28, 36, 39, 48, 57, 60, 71, 78, 111, 119, 128, 216
Cracking, 171
Cycloalkanes, 176, 177

D
Daltons Law, 87, 219
de Broglie, 101, 102
density, 12, 14, 15, 17, 20, 21, 59, 85, 86, 93, 119, 148, 222, 223, 224,
225, 235, 241, 251, 285, 301
deuterium, 26
diatomic molecules, 30, 155
diesel, 171
differential rate laws, 260
330
dihydrogen phosphate ion, 37
dihydrogen sulphide, 38
dilution, 76, 77, 244
Dilution calculations, 225
diprotic acid, 65
Dissociation, 60

E
Effusion, 90
Electrical work, 197
electrolysis, 9
electrolyte, 60, 61, 67, 76, 251
electrolytes, 60, 61, 63, 64, 251, 300
Electromagnetic Radiation, 95
electromagnetic spectrum, 96, 112
Electronegativity, 141
Electronic Structure, 95
Emission and absorption spectra, 98
empirical formula, 30, 33, 41, 44, 50, 51, 52, 57, 58, 93
endothermic, 127, 198, 204, 205, 209, 210, 226, 227, 234, 272, 293,
294, 298
enthalpy, 203, 204, 205, 206, 207, 208, 210, 211, 212, 213, 214, 215,
216, 217, 233, 234
Enthalpy of formation, 208
enthalpy of fusion, 205, 206, 216
enthalpy of neutralization, 207
entropy, 227, 231
enzyme, 206, 266, 267, 275
enzyme-substrate complex, 267
Equilibrium Constant, 279
Esters, 192
Ethers, 191
excited states, 100, 108, 110
exothermic, 123, 126, 198, 200, 201, 204, 205, 209, 210, 211, 216,
226, 234, 272, 293, 296
Exothermic reactions, 198
Extensive properties, 12
extensive property, 12, 199
331
F
first-order reaction, 262, 263, 264, 265
Formal Charge, 139
Formula Weights, 44
fractional distillation, 170
Freezing, 205, 240, 241
Freezing point depression, 238
frequency, 95, 96, 97, 112
functional group, 164, 169, 190

G
Gasoline, 171
Gay-Lussac, 84
geometrical isomers, 169, 180, 194, 195
germanium, 29
gold, 11, 21, 28, 276
Grahams Law, 90, 94
graphite, 11, 207, 208, 210, 211, 212, 214
ground state, 100, 101, 108, 110, 112, 113, 114, 123

H
Haber process, 296
half-life, 260, 262, 263, 264, 265, 276
half-reaction method, 72, 79
Halogenation, 173, 174
halogens, 10, 28, 36, 70, 129, 173, 182, 187
Heat capacity, 198
heat of atomization, 211
Henderson-Hasselbalch Equation, 323
Henry's Law, 229
Heptane, 165, 171
Hesss law, 207, 208, 211, 212
heterogeneous catalysis, 266
Heterogeneous Equilibrium, 283
Homogeneous, 20
Homologous series, 160, 166
Hunds rule, 110, 111

332
hybridization, 157, 160, 161, 162, 163, 184, 194
Hydration energy, 232
hydrocarbons, 162, 164, 165, 166, 167, 170, 171, 172, 178, 182, 186,
189, 227, 231
hydrochloric acid, 37, 38, 59, 61, 67, 68, 79, 80, 93, 94, 225, 293, 307
hydrocyanic acid, 37
hydrogen bonding, 227
hypochlorous acid, 37, 79

I
Ideal Gas Law, 21, 84, 85, 87, 92, 93, 224
indicator, 309, 316
inert gases, 28, 118
Initial rate method, 258
integral rate laws, 260
Intensive properties, 12
ionic compounds, 11, 29, 33, 36, 41, 44, 52, 60, 62, 130
Ionic Compounds, 30
Ionic radii, 121
ionization, 60, 123, 124, 125, 126, 128, 129, 135, 141, 301, 305, 309,
310, 312, 317
Ions, 23, 31
iron, 9, 11, 15, 28, 35, 37, 58, 73, 319
Isomerism, 160, 169
isooctane, 171
isotopes, 24, 25, 26, 40
IUPAC, 34, 115, 167, 190

K
Kelvin scale, 14
kerosene, 171
Ketones, 191
kinetic energy, 92, 97, 98
Kp and Kc, 282
Ksp, 324, 325, 326

L
lanthanides, 28, 117
333
Lattice Energies, 130
Lattice energy, 130, 232
Law of Constant Composition, 11
Le Chtelier Principle, 291
Lewis Dot Symbols, 131
Lewis structures, 134, 136, 137, 142, 143
Limiting Reactants, 54
line formula, 166
Line Spectrum, 98
lock and key mechanism, 275
lone pairs, 134, 143, 147, 149, 150, 151, 152, 153, 157
lubricating oils, 171
Lyman series, 99

M
magnesium, 12, 27, 59, 63, 68, 78, 121, 131, 132, 319
magnesium oxide, 12
Magnetic quantum number, 105
mass number, 24, 25, 26
Mass percent, 220
mass percentage, 45, 46, 50, 57
matter, 8, 9, 10, 11, 46, 101, 102, 197, 199, 219, 229, 241, 260, 279
Maxwell-Boltzman distribution law, 273
Mechanical work, 197
mechanism, 174, 175, 184, 275, 302
Mechanism, 174, 184
metalloids, 28, 29, 115
Metals, 28, 29, 141
mineral acids, 65
Molality, 223, 241, 242
Molar Mass, 46
Molarity, 75, 76, 77, 221, 225, 258
Mole, 46, 57, 88, 222, 224
Mole fraction, 222
molecular formula, 30
Molecular Geometry, 146
Molecules, 8, 23, 142, 143, 146
monoprotic acid, 65, 316
334
N
natural gases, 170, 171
Net ionic equations, 64
Neutralization, 67
neutrons, 23, 24, 25, 39, 40, 97
nickel, 28, 187, 199
nitrogen, 10, 20, 35, 39, 43, 54, 57, 59, 88, 90, 93, 127, 131, 132, 133,
134, 139, 149, 158, 193, 277, 293, 294, 296, 297, 309
nomenclature, 11, 167, 191
nonelectrolyte, 60
non-metals, 28, 29, 31, 38, 115, 129, 141
non-polar molecules, 227
nucleic acids, 158
nucleus, 23, 24, 100, 103, 104, 107, 119, 120, 123, 124, 126, 127, 150,
301

O
octane rating, 171
Octane Rating, 171
octet rule, 129, 132, 137, 138, 139, 142, 144
orbital, 103, 104, 105, 106, 107, 108, 109, 111, 113, 114, 115, 117,
120, 121, 123, 126, 127, 149, 150, 161, 162, 189
Orbital diagram, 109, 110
Organic Chemistry, 158
Orientation Factor, 270
Osmosis, 244, 245, 250
Osmotic Pressure, 244
Oxidation numbers, 69
Oxidation-Reduction, 68
oxidising agent, 70, 71, 73, 74
oxyanions, 35, 36, 70

P
paraffin, 171
Partial pressure, 87
Paschen, 99, 112
Pauli Exclusion Principle, 106, 115

335
Paulings Rule, 214, 215
Percent Ionization, 310
perchloric acid, 37
Periodic Table, 10, 27, 28, 29, 31, 32, 38, 40, 44, 48, 104, 107, 115,
117, 120, 122, 124, 126, 142
petrol, 171
Petroleum, 170
Pfund series, 99
pH, 305, 306, 307, 308, 309, 311, 312, 314, 315, 316, 317, 318, 319,
320, 321, 322, 323, 326, 327
Phase diagram, 243
Phenols, 191
phosphorus, 10, 20, 36, 52, 57, 151, 158, 290
photoelectric effect, 97, 98
Photons, 96
Physical properties, 12
pi bonds, 162, 163, 186, 189, 194
pKw, 306
Plancks constant, 96, 102
pOH, 306, 307, 317, 321
Polar molecules, 227
polonium, 29
Polyatomic molecules, 30
Polyprotic Acids, 312
potassium, 11, 20, 28, 35, 36, 67, 78, 79, 87, 129, 131, 220, 222, 302
Precision, 19
Pressure, 81, 93
Protons, 23, 24

Q
quantum, 97, 100, 103, 104, 105, 106, 107, 113, 114, 117, 120, 121
quantum number, 107, 120, 125
quantum numbers, 103, 107

R
radioactive decay, 264
Raoult's law, 236, 239
rare earth elements, 28
336
rate constant, 256, 263, 264, 267, 268, 274, 276
rate law, 256, 257, 260, 261, 262, 263, 264, 265, 266, 276
reaction kinetics, 255
Reaction Order, 256
reaction quotient, 287
Reaction Rate, 253
redox reaction, 69, 70
reducing agent, 70, 71, 73, 74
Resonance, 137
Reverse osmosis, 248
Rydberg constant, 99
Rydberg equation, 99

S
Saturated solution, 219
Schrdinger equation, 103
second order reaction, 256, 265
Second-order reactions, 265
Semi permeable membranes, 244
semiconductors, 29
SI Units, 13
sigma bonds, 161, 163
Significant figures, 18
silicate, 35
silicon, 10, 11, 20, 29, 35
silver, 28, 62, 63, 71, 78
skeletal isomerism, 169
sodium, 9, 11, 25, 28, 30, 31, 37, 49, 59, 60, 66, 67, 78, 79, 115, 119,
130, 219, 220, 232, 250, 307, 315, 316, 319
Solubility, 62, 78, 172, 219, 229, 231
Solubility Equilibria, 324
solute, 60, 75, 76, 218, 219, 220, 221, 223, 224, 226, 228, 231, 235,
236, 237, 238, 239, 242, 243, 244, 246, 247, 248, 251
Solutions, 60, 218, 226, 231, 235, 318
solvent, 60, 76, 159, 218, 219, 221, 222, 223, 226, 230, 231, 235, 236,
237, 238, 239, 242, 243, 244, 245, 246, 247, 251, 284
sp Hybridization, 162
sp2 Hybridization, 161
337
sp3 Hybridization, 161
specific heat, 199, 200, 201, 216
spectator ions, 64, 65, 78
Spin quantum number, 106
standard state, 207, 209
State property, 204
Stereoisomerism, 169
Stoichiometry, 42
strong acid, 65, 66, 67, 207, 307, 314, 316, 318, 319, 320
structural formula, 30, 166, 169, 194, 196
Structural isomers, 169
subatomic particles, 23, 97
Sublimation, 205
subshells, 104, 105, 106, 107, 108, 113, 125
Subshells, 104
sulphur, 10, 11, 30, 35, 36, 39, 57, 59, 94, 158, 286, 287
sulphuric acid, 37, 65, 79, 185, 286
sulphurous acid, 37, 79
surface area, 253
suspensions, 20

T
tellurium, 29
Temperature, 13, 14, 253, 267, 277
theoretical yield, 56
thermal energy, 198
thermal equilibrium, 199
thermochemical equation, 206, 207, 210, 216, 217
Thermochemistry, 197
titration, 80, 218, 313, 314, 316
Titrations, 77, 313
transition metal, 36, 70, 112
Transition metal, 34
transition metals, 28, 32, 35
triprotic acid, 65
tritium, 26

338
U
Units of Measurement, 12
urea, 158, 159, 224, 251

V
van't Hoff equation, 246, 247, 248
Vaporization, 205
Vapour pressure lowering, 235
volume percent, 220
VSEPR, 146, 147

W
Wavelength, 95
Wohler, 158
work, 97, 98, 123, 197, 198, 215, 256, 297, 313

Y
Yield, 56

Z
Zero-order processes, 265

339

You might also like