0% found this document useful (0 votes)
40 views

CHM 101 General Chemistry I (Physical) Lecture Notes

Huh
Copyright
© © All Rights Reserved
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
40 views

CHM 101 General Chemistry I (Physical) Lecture Notes

Huh
Copyright
© © All Rights Reserved
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 27

CHM 101- PHYSICAL CHEMISTRY I

Course Synopsis:
4. Modern electronic theory of atoms
5. Radioactivity
6. Chemical bonding

4. MODERN ELECTRONIC THEORY OF ATOMS


In our study of atomic structure, we look first at the fundamental particles. These are the basic building
blocks of all atoms. Atoms, and hence all matter, consist principally of three fundamental particles:
electrons, protons, and neutrons. Knowledge of the nature and functions of these particles is essential to
understanding chemical interactions. Many other subatomic particles, such as quarks, positrons, neutrinos,
pions, and muons, have also been discovered.
The relative masses and charges of the three fundamental particles are shown in Table below

Fundamental Particles of Matter


Particle Mass Charge (relative scale)
electron (e─) 0.00054858 amu 1−
proton (p or p+) 1.0073 amu 1+
neutron (n or n0) 1.0087 amu None

The mass of an electron is very small compared with the mass of either a proton or a neutron. The charge
on a proton is equal in magnitude, but opposite in sign, to the charge on an electron. The proton is a
fundamental particle with a charge equal in magnitude but opposite in sign to the charge on the electron.
Its mass is almost 1836 times that of the electron.

The Discovery of Electrons


Some of the earliest evidence about atomic structure was supplied in the early 1800s by the English chemist
Humphry Davy (1778–1829). He found that when he passed electric current through some substances, the
substances decomposed. He therefore suggested that the elements of a chemical compound are held together
by electrical forces. In 1832–1833, Michael Faraday (1791–1867), Davy’s student, determined the
quantitative relationship between the amount of electricity used in electrolysis and the amount of chemical
reaction that occurs. Studies of Faraday’s work by George Stoney (1826–1911) led him to suggest in 1874
that units of electric charge are associated with atoms. In 1891, he suggested that they be named electrons.
1|Page
The most convincing evidence for the existence of electrons came from experiments using cathode-ray
tubes (Figure below).

Figure…….Some experiments with cathode-ray tubes that show the nature of cathode rays. (a) A cathode-
ray (discharge) tube, showing the production of a beam of electrons (cathode rays). The beam is detected
by observing the glow of a fluorescent screen. (b) A small object placed in a beam of cathode rays casts a
shadow. This shows that cathode rays travel in straight lines. (c) Cathode rays have negative electric charge,
as demonstrated by their deflection in an electric field. (The electrically charged plates produce an electric
field.) (d) Interaction of cathode rays with a magnetic field is also consistent with negative charge. The
magnetic field goes from one pole to the other. (e) Cathode rays have mass, as shown by their ability to
turn a small paddle wheel in their path.

Two electrodes are sealed in a glass tube containing gas at a very low pressure. When a high voltage is
applied, current flows and rays are given off by the cathode (negative electrode). These rays travel in
straight lines toward the anode (positive electrode) and cause the walls opposite the cathode to glow. An
object placed in the path of the cathode rays casts a shadow on a zinc sulphide screen placed near the anode.
The shadow shows that the rays travel from the cathode toward the anode. The rays must therefore be

2|Page
negatively charged. Furthermore, they are deflected by electric and magnetic fields in the directions
expected for negatively charged particles.
In 1897 J. J. Thomson (1856–1940) studied these negatively charged particles more carefully. He
called them electrons, the name Stoney had suggested in 1891. By studying the degree of deflections of
cathode rays in different electric and magnetic fields, Thomson determined the ratio of the charge (e) of the
electron to its mass (m). The modern value for this ratio is
e/m = 1.75882 × 108 coulomb (C)/gram
Note: The coulomb (C) is the standard unit of quantity of electric charge. It is defined as the quantity of
electricity transported in one second by a current of one ampere. It corresponds to the amount of electricity
that will deposit 0.00111798 g of silver in an apparatus set up for plating silver.

This ratio is the same regardless of the type of gas in the tube, the composition of the electrodes, or the
nature of the electric power source. The clear implication of Thomson’s work was that electrons are
fundamental particles present in all atoms.
Once the charge-to-mass ratio for the electron had been determined, additional experiments were
necessary to determine the value of either its mass or its charge, so that the other could be calculated (Figure
below). In 1909, Robert Millikan (1868–1953) solved this dilemma with the famous “oil-drop experiment,”
in which he determined the charge of the electron.

Figure….. The Millikan oil-drop experiment.

3|Page
All of the charges measured by Millikan turned out to be integral multiples of the same number. He assumed
that this smallest charge was the charge on one electron. This value is 1.60218 × 10−19 coulomb (modern
value).

Rutherford’s mathematical analysis of his results showed that the scattering of positively charged
α-particles was caused by repulsion from very dense regions of positive charge in the gold foil. He
concluded that the mass of one of these regions is nearly equal to that of a gold atom, but that the diameter
is no more than 1/10,000 that of an atom. Many experiments with foils of different metals yielded similar
results. Realizing that these observations were inconsistent with previous theories about atomic structure,
Rutherford discarded the old theory and proposed a better one. He suggested that each atom contains a tiny,
positively charged, massive centre that he called an atomic nucleus. Most α-particles pass through metal
foils undeflected because atoms are primarily empty space populated only by the very light electrons. The
few particles that are deflected are the ones that come close to the heavy, highly charged metal nuclei.
Rutherford was able to determine the magnitudes of the positive charges on the atomic nuclei. The picture
of atomic structure that he developed is called the Rutherford model of the atom.

Atoms consist of very small, very dense positively charged nuclei surrounded by clouds of electrons at
relatively great distances from the nuclei. All nuclei contain protons; nuclei of all atoms except the common
form of hydrogen also contain neutrons.

ATOMIC NUMBER
Moseley (1887–1915) showed that the X-ray wavelengths could be better correlated with the atomic
number. On the basis of his mathematical analysis of these X-ray data, he concluded that

each element differs from the preceding element by having one more positive charge in its nucleus.

We now know that every nucleus contains an integral number of protons exactly equal to the number of
electrons in a neutral atom of the element. Every hydrogen atom contains one proton, every helium atom
contains two protons, and every lithium atom contains three protons. The number of protons in the nucleus
of an atom determines its identity; this number is known as the atomic number of that element.

NEUTRONS
The neutron is an uncharged particle with a mass slightly greater than that of the proton.

4|Page
MASS NUMBER AND ISOTOPES
Most elements consist of atoms of different masses, called isotopes. The isotopes of a given element contain
the same number of protons (and also the same number of electrons) because they are atoms of the same
element. They differ in mass because they contain different numbers of neutrons in their nuclei.
For example, there are three distinct kinds of hydrogen atoms, commonly called hydrogen, deuterium,
tritium. Each contains one proton in the atomic nucleus. The predominant form of hydrogen contains no
neutrons, but each deuterium atom contains one neutron and each tritium atom contains two neutrons in its
nucleus. All three forms of hydrogen display very similar chemical properties.

Isotopes are atoms of the same element with different masses; they are atoms containing the same
number of protons but different numbers of neutrons.
The mass number of an atom is the sum of the number of protons and the number of neutrons in its
nucleus

Mass number = number of protons + number of neutrons


= atomic number + neutron number
The composition of a nucleus is indicated by its nuclide symbol. This consists of the symbol for the element
(E), with the atomic number (Z) written as a subscript at the lower left and the mass number (A) as a
superscript at the upper left, AZE.

Note:
• The atomic number, Z, is an integer equal to the number of protons in the nucleus of an atom of the
element. It is also equal to the number of electrons in a neutral atom. It is the same for all atoms of an
element.
• The mass number, A, is an integer equal to the sum of the number of protons and the number of neutrons
in the nucleus of an atom of a particular isotope of an element. It is different for different isotopes of
the same element.
• Many elements occur in nature as mixtures of isotopes. The atomic weight of such an element is the
weighted average of the masses of its isotopes. Atomic weights are fractional numbers, not integers.

5|Page
Calculation of Atomic Weight
Three isotopes of magnesium occur in nature. Their abundances and masses, determined by mass
spectrometry, are listed in the following Table below. Use this information to calculate the atomic weight
of magnesium, Mg.
Isotope % Abundance Mass (amu)
78.99 22.98504

10.00 24.98584

11.01 25.98259

one amu is exactly 1/12 of the mass of a carbon-12 atom.

Solution
Multiply the fraction of each isotope by its mass and add these numbers to obtain the atomic weight of
magnesium.
Atomic weight = 0.7899(23.98504 amu) + 0.1000(24.98584 amu) + 0.1101(25.98259 amu)
= 18.946 amu + 2.4986 amu + 2.8607 amu
= 24.30 amu (to four significant figures)

Exercise
1. The atomic weight of gallium is 69.72 amu. The masses of the naturally occurring isotopes are
69 71
68.9257 amu for 31Ga and 70.9249 amu for 31Ga. Calculate the percent abundance of each
isotope.
Solution
Let x = fraction of 6931Ga, then (1−x) = fraction of 7131Ga
x(68.9257 amu) + (1−x)(70.9249 amu) = 69.72 amu
68.9257x amu + 70.9249 amu – 70.9249x = 69.72 amu
−1.9992x = 69.72
x = 0.600
x = 0.600 = fraction of 6931Ga, ⁛ 60% of 6931Ga
(1−x) = 0.400 = fraction of 7131Ga, ⁛ 40% of 7131Ga

6|Page
Quantum Numbers
Quantum numbers can be used to designate the electronic arrangements in all atoms, their so-called
electron configurations. These quantum numbers play important roles in describing the energy levels of
electrons and the shapes of the orbitals that describe distributions of electrons in space.
For now, let’s say that

an atomic orbital is a region of space in which the probability of finding an electron is high.

We define each quantum number and describe the range of values it may take
1. The principal quantum number, n, describes the main energy level, or shell, an electron
occupies. It may be any positive integer:
n = 1, 2, 3, ……
2. The angular momentum quantum number (azimuthal), ℓ, designates the shape of the region in space
that an electron occupies. Within a shell (defined by the value of n, the principal quantum number)
different sublevels or subshells are possible, each with a characteristic shape. The angular momentum
quantum number designates a sublevel, or specific shape of atomic orbital that an electron may occupy.
This number, ℓ, may take integral values from 0 up to and including (n = 1):
ℓ = 0, 1, 2, …….. (n-1)
Each letter corresponds to a different sublevel (subshell).
ℓ = 0, 1, 2, 3…….. (n-1)
s, p, d, f
3. The magnetic quantum number, mℓ, designates the specific orbital within a subshell. Orbitals within a
given subshell differ in their orientations in space, but not in their energies. Within each subshell, mℓ
may take any integral values from −ℓ through zero up to and including +ℓ:
mℓ = (−ℓ), …….., 0, ……. (+ℓ)
for example, when ℓ = 1, which designates the p subshell, there are three permissible values of mℓ: −1, 0,
and +1. Thus, three distinct regions of space, called atomic orbitals, are associated with a p subshell. We
refer to these orbitals as the px, py, and pz orbitals.
4. The spin quantum number, ms, refers to the spin of an electron and the orientation of the magnetic
field produced by this spin. For every set of n, ℓ, and mℓ values, ms can take the value +½ or −½.
It is noteworthy that the electronic structures of atoms are governed by the following:
(a) Pauli Exclusion Principle:

7|Page
No two electrons in an atom may have identical sets of four quantum numbers.

(b) Hund’s Rule


Electrons occupy all the orbitals of a given subshell singly before pairing begins. These unpaired
electrons have parallel spins.

Easier Way to Write Electronic Configuration

1s2, 2s2, 2p6, 3s2, 3p6, 4s2, ……..


Note: The electronic configuration using s, p, d, f- format for neutral atom is quite different from its ionic
specie. For example, the electronic configuration for nitrogen (N = 7) is 1s2, 2s2, 2p3, while its ionic form
N3− is 1s2, 2s2, 2p6.

Things to Remember:
Angular momentum quantum number (ℓ): The quantum mechanical solution to a wave equation that
designates the sub-shell, or set of orbitals (s, p, d, f), within a given main shell in which an electron resides.
Atomic mass unit (amu): An arbitrary mass unit defined to be exactly one-twelfth the mass of the carbon-
12 isotope.
Atomic orbital: The region or volume in space in which the probability of finding electrons is highest.
Electron: A subatomic particle having a mass of 0.00054858 amu and a charge of 1−.
Hund’s Rule: Each orbital of a given subshell is occupied by a single electron before pairing begins.
Isotopes Two or more forms of atoms of the same element with different masses; that is, atoms containing
the same number of protons but different numbers of neutrons.
Magnetic quantum number (mℓ): Quantum mechanical solution to a wave equation that designates the
particular orbital within a given subshell (s, p, d, f) in which an electron resides.

8|Page
Mass number: The integral sum of the numbers of protons and neutrons in an atom.
Neutron: A subatomic nuclear particle having a mass of 1.0087 amu and no charge.
Nucleus: The very small, very dense, positively charged center of an atom containing protons and neutrons,
Pauli Exclusion Principle: No two electrons in the same atom may have identical sets of four quantum
numbers.
Principal quantum number (n): The quantum mechanical solution to a wave equation that designates the
main shell, or energy level, in which an electron resides.
Proton: A subatomic particle having a mass of 1.0073 amu and a charge of 1+, found in the nuclei of atoms.
Quantum numbers: Numbers that describe the energies of electrons in atoms; they are derived from
quantum mechanical treatment.
Spin quantum number (ms): The quantum mechanical solution to a wave equation that indicates the
relative spins of electrons.

Further Exercise
1. Determine the number of protons, neutrons, and electrons in each of the following species. Are the
members within each pair isotopes?

(a) and ; and (b) and


2. How many protons, neutrons and electrons are in the following isotopes

(a) (b) (c) (d)


3. Determine the number of protons, neutrons, and electrons in each of the following species: (a) 4020Ca
(b) 4521Sc (c) 9140Zr (d) 3919K+ (e) 6530Zn2+ (f ) 10847Ag+.
4. Write the composition of one atom of each of the three isotopes of silicon: 28Si, 29Si, 30Si.
5. Write the composition of one atom of each of the four isotopes of sulphur: 32S, 33S, 34S, 36S.
6. The atomic weight of lithium is 6.941 amu. The two naturally occurring isotopes of lithium have
the following masses: 6Li, 6.01512 amu; 7Li, 7.01600 amu. Calculate the percent of 6Li in naturally
occurring lithium.
7. Naturally occurring iron consists of four isotopes with the indicated abundances below. From the
masses and relative abundances of these isotopes, calculate the atomic weight of naturally occurring
iron.

9|Page
Isotope Isotopic Mass (amu) % Natural Abundance
54
Fe 53.9396 5.82
56
Fe 55.9349 91.66
57
Fe 56.9354 2.19
58
Fe 57.9333 0.33

8. The atomic weight of rubidium is 85.4678 amu. The two naturally occurring isotopes of rubidium
have the following masses: 85Rb, 84.9118 amu; 87Rb, 86.9092 amu. Calculate the percent of 85
Rb in
naturally occurring rubidium.
9. Bromine is composed of 3579Br, 78.9183 amu, and 3581Br 80.9163 amu. The percent composition of
a sample is 50.69% Br-79 and 49.31% Br-81. Based on this sample, calculate the atomic weight of
bromine.
10. The atomic weight of copper is 63.546 amu. The two naturally occurring isotopes of copper have the
63 65 63
following masses: Cu, 62.9298 amu and Cu, 64.9278 amu. Calculate the percent of Cu in
naturally occurring copper.
11. Silver consists of two naturally occurring isotopes: 107Ag, which has a mass of 106.90509 amu, and
109
Ag, which has a mass of 108.9047 amu. The atomic weight of silver is 107.8682 amu. Determine
the percent abundance of each isotope in naturally occurring silver.
12. Write the electronic configuration of the following species:
(a) Cl (b) Cl+ (c) Na+ (d) Mg (e) P3− (f) Sc (g) V (h) Fe (i) Fe3+ (j) Cr3+ (k) Co (l) O2− (m) Cu2+

5. RADIOACTIVITY
Chemical properties are determined by electron distributions and are only indirectly influenced by
atomic nuclei. Until now, discussions have always been about ordinary chemical reactions, so we focused
on electron configurations. Nuclear reaction changes in the composition of nuclei. These extraordinary
processes are often accompanied by the release of tremendous amounts of energy and by transmutations of
elements. Some differences between nuclear- and ordinary chemical reactions are as follows:

10 | P a g e
Nuclear Reaction Ordinary Chemical Reaction
1. Elements may be converted from one to 1. No new elements can be produced.
another.
2. Particles within the nucleus are involved 2. Only the electrons participate.
3. Tremendous amounts of energy are 3. Relatively small amounts of energy are released
released or absorbed or absorbed
4. Rate of reaction is not influenced by 4. Rate of reaction depends on factors such as
external factors. concentration, temperature, catalyst, and pressure.

Medieval alchemists spent years trying to convert other metals into gold without success. Years of
failure and the acceptance of Dalton’s atomic theory early in the nineteenth century convinced scientists
that one element could not be converted into another. Then, in 1896 Henri Becquerel discovered
“radioactive rays” (natural radioactivity) coming from a uranium compound. Ernest Rutherford’s study
of these rays showed that atoms of one element may indeed be converted into atoms of other elements by
spontaneous nuclear disintegrations. Many years later it was shown that nuclear reactions initiated by
bombardment of nuclei with accelerated subatomic particles or other nuclei can also transform one element
into another —accompanied by the release of radiation (induced radioactivity).
Becquerel’s discovery led other researchers, including Marie and Pierre Curie, to discover and study new
radioactive elements. Many radioactive isotopes, or radioisotopes, now have important medical,
agricultural, and industrial uses.

The Nucleus
The neutrons and protons together constitute the nucleus, with the electrons occupying essentially
empty space around the nucleus. The nucleus is only a minute fraction of the total volume of an atom, yet
nearly all the mass of an atom resides in the nucleus. Thus, nuclei are very small (10−13 cm) and extremely
dense. It has been shown experimentally that nuclei of all elements have approximately the same density,
2.4 × 10−14 g/cm3.

Nuclear Stability and Radioactive Decay


Many nuclei are radioactive, which means they can decompose, forming another nucleus and
producing one or more other particles.

A typical radioactive decay equation (an electron, beta decay, (β-particle))

11 | P a g e
Types of Radioactive Decay
There are two main categories of radioactive decay.
1. Those that involve a change in the mass number (atomic mass) of the decaying nucleus, and
2. Those that do not

Examples
1. (a) Alpha-particle production (α-particle = a helium nucleus, 42He.

(b) Spontaneous Fission


This involves the splitting of a heavy nuclide into two (2) lighter nuclides with similar (smaller) mass
numbers

2. (a) β−particle Production


During β−particle production, an electron is produced, while the mass number of the decaying
nucleus remains unchanged, as seen below.

(b) Gamma-Ray Production


This occurs when a high-energy (excited) nucleus losses energy in the form of electromagnetic radiation
(EMR) called gamma-rays, γ

(c) Electron Capture

12 | P a g e
(d) Positron Production
Herein, the positron is the same as β+ particle, or 0
+1 e, as the positron emission is commonly encountered
with artificially produced radioactive nucleii of the lighter elements

Note:
Sometimes, nuclear decays and particle reactions, such as α-particle decay (helium particle) are
accompanied by gamma-ray production.

Half-life, t½.
Radionuclides have different stabilities and decay at different rates. Some decay nearly completely in a
fraction of a second and others only after millions of years. The rates of all radioactive decays are
independent of temperature and obey first-order kinetics. The rate law and the integrated rate equation for
a first-order process are:
𝐴
rate of decay = k[A] and 𝑙𝑛 ( 𝐴0 ) = 𝑎𝑘𝑡

Each atom decays independently of the others, hence, the stoichiometric coefficient a is always 1 for
radioactive decay.
The rate of radioactive disintegrations follows first-order kinetics, so it is proportional to the amount
of A present; we can write the integrated rate equation in terms of N, the number of disintegrations per unit
time:
𝑁0
𝑙𝑛 ( ) = 𝑘𝑡
𝑁
N0 = the amount of radionuclide before disintegration
N = the amount of radionuclide after disintegration time, t.
In nuclear chemistry, the decay rate is usually expressed in terms of the half-life, t1/2, of the process.
This refers to the time required for the mass of nuclides to reach half the original value.
ln 2 0.693
𝑡1⁄ = =
2 𝑘 𝑘
13 | P a g e
k = rate constant.
The half-lives of radioactive isotopes (nuclides) can vary tremendously from 109 years to 10−4 seconds

Example 1:
Consider a technetium-99 or 9943Tc. The rate constant for its radioactive decay is 1.16×10−1 h. Calculate
the half-life of this nuclide.
Solution
ln 2 0.693
𝑡1⁄ = =
2 𝑘 𝑘
0.693/0.116/h = 5.98 h

Example 2
Cobalt-60 decays with the emission of beta particles and gamma rays, with a half-life of 5.27 years.

How much of a 3.42 µg sample of cobalt-60 remains after 30.0 years?


Solution
The value of the specific rate constant is determined first.
ln 2 0.693
𝑡1⁄ = = , so k = 0.693/t1/2 = 0.693/5.27 y = 0.131 y−1
2 𝑘 𝑘

This value can now be used to determine the ratio of N0 to N after 30.0 years.
𝑁
𝑙𝑛 ( 𝑁0 ) = 𝑘𝑡 = 0.131 y−1 (30.0) = 3.93

Taking the inverse ln of both sides, N0/N = 51


N0 = 3.42µg, so
𝑁0
𝑁= = 3.42 µg/51 = 0.067µg of 6027Co after 30.0 years
51

Detection of Radioactivity
1. Photographic Detection: Emanations from radioactive substances affect photographic plates just as
ordinary visible light does. Becquerel’s discovery of radioactivity resulted from the unexpected
exposure of such a plate, wrapped in black paper, by a nearby enclosed sample of a uranium-containing
compound, potassium uranyl sulfate. After a photographic plate has been developed and fixed, the
intensity of the exposed spot is related to the amount of radiation that struck the plate. Quantitative
detection of radiation by this method is difficult and tedious.

14 | P a g e
2. Detection by Fluorescence: Fluorescent substances can absorb high-energy radiation such as gamma
rays and subsequently emit visible light. As the radiation is absorbed, the absorbing atoms jump to
excited electronic states. The excited electrons return to their ground states through a series of
transitions, some of which emit visible light. This method may be used for the quantitative detection
of radiation, using an instrument called a scintillation counter.
3. Cloud Chambers: A chamber contains air saturated with vapor. Particles emitted from a radioactive
substance ionize air molecules in the chamber. Cooling the chamber causes droplets of liquid to
condense on these ions. The paths of the particles can be followed by observing the fog-like tracks
produced. The tracks may be photographed and studied in detail.
4. Gas Ionisation Counters: Radioactive decay processes generate high energy particles such as α-, β- and
gamma particles. These particles can produce ions when traveling through matter. Geiger Muller
Counter takes advantage of this. Geiger Muller Counter measures radioactivity by “capturing” particles
through a window opening in the counter. Once inside, these radioactive particles ionises Ar(g) to Ar+(g),
and the amount of the ionised particle is then measured.
5. Carbon-14 Dating: This is a method used for dating ancient artefacts made from wood or cloth.
Radioactive carbon dating is based on

14
Carbon-14 is continually produced in our atmosphere, so the concentration of 6C in the atmosphere
remains constant over time because the above two processes (reactions) occur at the same rate.

NUCLEAR FISSION
This involves the splitting of a heavy nucleus into two (2) nuclei with smaller mass numbers. For example,
a Uranium-235 nuclide after being bombarded with neutrons, were observed to split into 2 lighter elements

15 | P a g e
This nuclear fission process releases 2.1×1013 J/mol (26 million times larger than burning methane, CH4)
of uranium-235. Furthermore, the production of more neutrons in the reaction males it possible to have a
self-sustaining chain reaction

NUCLEAR FUSION
This reaction involves the combination of two (2) lighter nuclei to form a heavier, more stable nucleus,
with enormous release of energy. Stars, including the Sun, produce energy via nuclear fusion processes.

There is one major problem that makes fusion reactions difficult as potential sources of energy for us on
earth, as the protons (11H) each other electrostatically, so it is very difficult to get them to “fuse”, however

16 | P a g e
to make fusion occur, scientists have to “shoot” protons at each other at very high speeds to overcome the
huge p+−p+ repulsion, and this requires temperatures of 4×107 K.

STABILITY OF THE NUCLEUS


Mass Deficiency and Nuclear Binding Energy
The Mass Deficiency, Δm
The mass deficiency, Δm, for a nucleus is the difference between the sum of the masses of electrons, protons,
and neutrons in the atom (calculated mass) and the actual measured mass of the atom.
Δm = (sum of masses of all e−, p+ and n0) ─ (actual mass of atom).
The mass deficiency for most naturally occurring isotopes is only about 0.15% or less of the calculated
mass of an atom.
For example,
1. Calculate the mass deficiency for chlorine-35 atoms in amu/atom and in g/mol atoms. The actual mass
of a chlorine-35 atom is 34.9689 amu.
Solution
Each atom of 3517Cl contains 17 protons, 17 electrons, and (35 − 17) = 18 neutrons.
protons: 17 × 1.0073 amu = 17.124 amu
electrons: 17 × 0.00054858 amu = 0.0093 amu
neutrons: 18 × 1.0087 amu = 18.157 amu
sum = 35.290 amu* (calculated mass)
Thereafter, we subtract the actual mass from the “calculated” mass to obtain the Δm
Δm = 35.290 amu – 34.9689 amu = 0.321 amu (mass deficiency in one atom)
Recall that 1g is 6.022×1023 amu. It has been shown via calculations that a number expressed in amu/atom
is the same as the number in g/mol of atoms, hence
0.321 amu = 0.321 g/mol of 35Cl atoms (mass deficiency in a mole of Cl atoms).
Mass defect (Δm) can be used to calculated the corresponding nuclear binding energy, BE, which
is the energy required to decompose a nucleus into its constituent nucleons (protons and neutrons). The
overall loss of mass is compensated for, by a tremendous release of energy.
BE = mc2 or ΔE = (Δm)c2.
Nuclear binding energies may be expressed in many different units, including kilo-joules/mole of atoms,
kilojoules/gram of atoms, and megaelectron volts/nucleon. Some useful equivalences are
1 megaelectron volt (MeV) = 1.60 × 10−13 J
1 joule (J) = 1 kg∙m2/s2

17 | P a g e
Exercise
For example, let us use the value of Δm for 35Cl atoms to calculate their nuclear binding energy. Calculate
the nuclear binding energy 35Cl in (a) kilojoules per mole of Cl atoms, (b) kilo-joules per gram of Cl atoms,
and (c) megaelectron volts per nucleon.
Solution
The mass deficiency is 0.321 g/mol = 3.21 × 10−4 kg/mol.
(a) BE = (Δm)c2 = 3.21 × 10−4 kg × (3.00 × 108 m/s)2 = 2.89 × 1013 kg∙m2/s2
mol 35Cl atoms mol 35Cl atoms

= 2.89 × 1013 J/mol 35Cl atoms = 2.89 × 1010 kJ/mol 35Cl atoms
(b) From the previous example of 35Cl, the actual mass is
34.9689 amu or 34.9689 g
35
Cl atom mol 35Cl atom
This mass can be used to do the required conversion factor
BE = 2.89 × 1010 kJ × 1 mol 35Cl atoms = 8.26 × 108 kJ/g 35Cl atoms
mol of mol 35Cl atoms 34.9689 g 35Cl atoms
(c) The number of nucleons in one atom of 35Cl is 17 protons + 18 neutrons = 35 nucleons.
BE = 2.89 × 1010 kJ × 1000J × 1 MeV × 1 mol 35Cl atoms × 1 35Cl atom
mol of mol 35Cl atoms kJ 1.60 × 10−4 kg 6.022 .9689 g 35Cl atoms 35 nucleons
= 8.57 MeV/nucleon

Further Exercise
(1) The half-life of 198O is 29 s. What fraction of the isotope originally present would be left after 10.0 s?
(2) The half-life of 116C is 20.3 min. How long will it take for 95.0% of a sample to decay? How long will
it take for 99.5% of the sample to decay?
(3) The activity of a sample of tritium decreased by 5.5% over the period of a year. What is the half-life
of 31H?
(4) A very unstable isotope of beryllium, 8Be, undergoes emission with a half-life of 0.07 fs. How long
does it take for 99.99% of a 1.0- µg sample of 8Be to undergo decay?
(5) The actual mass of a 62Ni atom is 61.9283 amu. (a) Calculate the mass deficiency in amu/atom and in
g/mol for this isotope. (b) What is the nuclear binding energy in kJ/mol for this isotope?
(6) The actual mass of a 108Pd atom is 107.90389 amu. (a) Calculate the mass deficiency in amu/atom and
in g/mol for this isotope. (b) What is the nuclear binding energy in kJ/mol for this isotope?
18 | P a g e
(7) Calculate the following for 64Zn (actual mass = 63.9291 amu). (a) mass deficiency in amu/atom; (b)
mass deficiency in g/mol; (c) binding energy in J/atom; (d) binding energy in kJ/mol; (e) binding
energy in MeV/nucleon.
(8) Calculate the following for 49Ti (actual mass = 48.94787 amu). (a) mass deficiency in amu/atom; (b)
mass deficiency in g/mol; (c) binding energy in J/atom; (d) binding energy in kJ/mol; (e) binding
energy in MeV/nucleon.
(9) Calculate the nuclear binding energy in kJ/mol for each of the following (a) 12753I; (b) 8135Br; (c) 3517Cl.
The atomic masses are 126.9044 amu, 80.9163 amu, and 34.96885 amu, respectively.

6. CHEMICAL BONDING
Chemical bonding refers to the attractive forces that hold atoms together in compounds. There are
two major classes of bonding.
(1) Ionic bonding results from electrostatic interactions among ions, which often results from the net
transfer of one or more electrons from one atom or group of atoms to another. Ionic bonds are stronger than
covalent bonds and have a higher bond energy (the energy required to break a bond).
(2) Covalent bonding results from sharing one or more electron pairs between two atoms.
These two classes represent two extremes; all bonds between atoms of different elements have at
least some degree of both ionic and covalent character. Compounds containing predominantly ionic
bonding are called ionic compounds. Those that are held together mainly by covalent bonds are called
covalent compounds. Some nonmetallic elements, such as H2, Cl2, N2, and P4, also involve covalent
bonding. Some properties usually associated with many simple ionic and covalent compounds are
summarized in the following list. The differences in these properties can be accounted for by the differences
in bonding between the atoms or ions.

19 | P a g e
Ionic Compounds Covalent Compounds
1. They are solids with high melting points 1. They are gases, liquids, or solids with low
(typically >400°C). melting points (typically <300°C).
2. Many are soluble in polar solvents such as 2. Many are insoluble in polar solvents.
water
3. Most are insoluble in nonpolar solvents, such 3. Most are insoluble in nonpolar solvents, such as
as hexane, C6H14 and carbon tetrachloride, hexane, C6H14 and carbon tetrachloride, CCl4.
CCl4.
4. Molten compounds conduct electricity well 4. Liquid and molten compounds do not
because they contain mobile charged conduct electricity.
particles (ions).
5. Aqueous solutions conduct electricity well 5. Aqueous solutions are usually poor conductors
because they contain mobile charged of electricity because most do not contain
particles (ions). charged particles.
6. They are often formed between two elements 6. They are often formed between two elements
with quite different electronegativities, with similar electronegativities, usually
usually a metal and a nonmetal. nonmetals.

Lewis Dot Formulas of Atoms


The number and arrangements of electrons in the outermost shells of atoms determine the chemical
and physical properties of the elements as well as the kinds of chemical bonds they form. We write Lewis
dot formulas (or Lewis dot representations, or just Lewis formulas) as a convenient bookkeeping
method for keeping track of these “chemically important electrons.” We now introduce this method for
atoms of elements; in our discussion of chemical bonding in subsequent sections, we will frequently use
such formulas for atoms, molecules, and ions. Chemical bonding usually involves only the outermost
electrons of atoms, also called valence electrons. In Lewis dot representations, only the electrons in the
outermost occupied s and p orbitals are shown as dots. Paired and unpaired electrons are also indicated.
The table below shows Lewis dot formulas for the representative elements

20 | P a g e
Ionic Bonding
Ionic bonding is the attraction of oppositely charged ions (cations and anions) in large numbers to form a
solid. Such a solid compound is called an ionic solid.
As can be recalled, an ion is an atom or a group of atoms that carries an electrical charge. An
ion in which the atom or group of atoms has fewer electrons than protons is positively charged, and is
called a cation; one that has more electrons than protons is negatively charged, and is called an anion. An
ion that consists of only one atom is described as a monatomic ion. Examples include the chloride ion, Cl−,
and the magnesium ion, Mg2+. An ion that contains more than one atom is called a polyatomic ion.
Examples include the ammonium ion, NH4+; hydroxide ion, OH−; and sulphate ion, SO42−. The atoms of a
polyatomic ion are held together by covalent bonds.
Ionic bonding can occur easily when elements that have low ionization energies (metals) react with
elements having high electronegativities and very negative electron affinities (non-metals).
Many metals are easily oxidized—that is, they lose electrons to form cations; and many nonmetals are
readily reduced—that is, they gain electrons to form anions, i.e.

When the electronegativity difference, Δ(EN), between two elements is large, as between a metal and a
nonmetal, the elements are likely to form a compound by ionic bonding (transfer of electrons).

Let us describe some combinations of metals with nonmetals to form ionic compounds.

21 | P a g e
Group IA Metals and Group VIIA Nonmetals
2Na(s) + Cl2(g) → 2NaCl(s)
sodium chlorine sodium chloride

In this reaction, Na atoms lose one electron each to form Na+ ions, which contain only ten electrons, the
same number as the preceding noble gas, neon. We say that sodium ions have the neon electronic structure:
Na+ is isoelectronic with Ne. In contrast, Cl atoms gain one electron each to form Cl− ions, which contain
18 electrons. This is the same number as the following noble gas, argon; Cl− is isoelectronic with Ar. These
processes can be represented compactly as

Using Lewis dot formulas to represent the reaction, we have:

We can represent the general reaction of the IA metals with the VIIA elements as follows:

The Lewis dot representation for the generalized reaction is

Group IA Metals and Group VIA Nonmetals


Consider the reaction of lithium (Group IA) with oxygen (Group VIA) to form lithium oxide, a solid ionic
compound (mp > 1700°C)

22 | P a g e
In a compact representation,

The Lewis dot formulas for the atoms and ions are

Group IIA Metals and Group VIA Nonmetals


Consider the reaction of calcium (Group IIA) with oxygen (Group VIA). This reaction forms calcium
oxide, a white solid ionic compound with a very high melting point, 2580°C.

The Lewis dot notation for the reacting atoms and the resulting ions is

COVALENT BONDING
A covalent bond is formed when two atoms share one or more pairs of electrons. Covalent bonding occurs
when the electronegativity difference, Δ(EN), between elements (atoms) is zero or relatively small.
In predominantly covalent compounds the bonds between atoms within a molecule (intramolecular bonds)
are relatively strong, but the forces of attraction between molecules (intermolecular forces) are relatively
weak. As a result, covalent compounds have lower melting and boiling points than ionic compounds.

23 | P a g e
Formation of Covalent Bonds
Let us look at the simplest case of covalent bonding, the reaction of two hydrogen atoms to form the
diatomic molecule H2. As you recall, an isolated hydrogen atom has the ground state electron configuration
1s1, with the probability density for this one electron spherically distributed about the hydrogen nucleus (a).
As two hydrogen atoms approach each other, the electron of each hydrogen atom is attracted by the nucleus
of the other hydrogen atom as well as by its own nucleus (b). If these two electrons have opposite spins so
that they can occupy the same region (orbital), both electrons can now preferentially occupy the region
between the two nuclei, because they are attracted by both nuclei (c). The electrons are shared between the
two hydrogen atoms, and a single covalent bond is formed. We say that the 1s orbitals overlap so that both
electrons are now in the orbitals of both hydrogen atoms. The closer together the atoms come, the more
nearly this is true. In that sense, each hydrogen atom then has the helium configuration 1s2.

Other pairs of nonmetal atoms share electron pairs to form covalent bonds. The result of this sharing
is that each atom attains a more stable electron configuration—frequently the same as that of the nearest
noble gas. This results in a more stable arrangement for the bonded atoms. Most covalent bonds involve
sharing of two, four, or six electrons—that is, one, two, or three pairs of electrons. Two atoms form a single
covalent bond when they share one pair of electrons, a double covalent bond when they share two electron
pairs, and a triple covalent bond when they share three electron pairs. These are usually called simply single,
double, and triple bonds. Covalent bonds that involve sharing of one and three electrons are known, but are
relatively rare.
In a Lewis formula, we represent a covalent bond by writing each shared electron pair either as a
pair of two dots between the two atom symbols or as a dash connecting them.
Thus, the formation of H2 from H atoms could be represented as

where the dash represents a single bond.


Similarly, the combination of two fluorine atoms to form a molecule of fluorine, F2, can be shown as

24 | P a g e
The formation of a hydrogen fluoride, HF, molecule from a hydrogen atom and a fluorine atom can be
shown as

Also, we can use Lewis formulas to show the valence electrons in simple molecules. For examples,
(a) An H2O molecule has two shared electron pairs, that is, two single covalent bonds. The O atom also has
two unshared pairs.

(b) A CO2 molecule has four shared electron pairs in two double bonds. The central atom (C) has no
unshared pairs.

The Octet Rule


Representative elements usually attain stable noble gas electron configurations when they share
electrons. In the water molecule eight electrons are in the outer shell of the O atom, and it has the neon
electron configuration; two electrons are in the valence shell of each H atom, and each has the helium
electron configuration. Likewise, the C and O of CO2 and the N of NH3 and the NH4+ ion each have a share
in eight electrons in their outer shells. The H atoms in NH3 and NH4+ each share two electrons. Many Lewis
formulas are based on the idea that
in most of their compounds, the representative elements achieve noble gas configurations.

This statement is usually called the octet rule, because the noble gas configurations have 8 e− in their
outermost shells (except for He, which has 2 e−).
The octet rule alone does not let us write Lewis formulas. We still must decide how to place the electrons
around the bonded atoms−that is, how many of the available valence electrons are bonding electrons
(shared) and how many are unshared electrons (associated with only one atom). A pair of unshared
electrons in the same orbital is called a lone pair. A simple mathematical relationship is helpful here:
S=N–A
S is the total number of electrons shared in the molecule or polyatomic ion.

25 | P a g e
N is the total number of valence shell electrons needed by all the atoms in the molecule or ion to achieve
noble gas configurations (N = 8 × number of atoms that are not H, plus 2 × number of H atoms).
A is the number of electrons available in the valence shells of all of the (representative) atoms. This is equal
to the sum of their periodic group numbers. We must adjust A, if necessary, for ionic charges. We add
electrons to account for negative charges and subtract electrons to account for positive charges.
For example,
(a) For F2,
N = 2 × 8 (for two F atoms) = 16 e− needed
A = 2 × 7 (for two F atoms) = 14 e− available
S = N − A = 16 − 14 = 2 e− shared
The Lewis formula for F2 shows 14 valence electrons total, with 2 e− shared in a single bond.

(b) For HF,


N = 1 × 2 (for one H atom) + 1 × 8 (for one F atom) = 10 e− needed
A = 1 × 1 (for one H atom) + 1 × 7 (for one F atom) = 8 e− available
S = N − A = 10 − 8 = 2 e− shared
The Lewis formula for HF shows 8 valence electrons total, with 2 e− shared in a single bond.

(c) For H2O,


N = 2 × 2 (for two H atoms) + 1 × 8 (for one O atom) = 12 e− needed
A = 2 × 1 (for two H atoms) + 1 × 6 (for one O atom) = 8 e− available
S = N − A = 12 − 8 = 4 e− shared
The Lewis formula for H2O shows 8 valence electrons total, with a total of 4 e− shared, 2 e− in each single
bond.

(d) For CO2,


N = 1 × 8 (for one C atom) + 2 × 8 (for two O atoms) = 24 e− needed
A = 1 × 4 (for one C atom) + 2 × 6 (for two O atoms) = 16 e− available
S = N − A = 24 − 16 = 8 e− shared
The Lewis formula for CO2 shows 16 valence electrons total, with a total of 8 e− shared, 4 e− in each double
bond.
(e) For NH4+,
N = 1 × 8 (for one N atom) + 4 × 2 (for four H atoms) = 16 e− needed

26 | P a g e
A = 1 × 5 (for one N atom) + 4 × 1 (for four H atoms) − 1 (for 1+ charge) = 8 e− available
S = N − A = 16 − 8 = 8 e− shared
The Lewis formula for NH4+ shows 8 valence electrons total, with all 8 e− shared, 2 e− in each single bond.

Note: You should practice writing many Lewis formulas. A few common types of organic compounds and
their Lewis formulas are also identified. Each follows the octet rule. Methane, CH4, is the simplest of a
huge family of organic compounds called hydrocarbons (composed solely of hydrogen and carbon). Ethane,
C2H6, is another hydrocarbon that contains only single bonds. Ethylene, C2H4, has a carbon–carbon double
bond, and acetylene, C2H2, contains a carbon–carbon triple bond.

Further Exercise
Calculate S, A and N, for the following species.
(a) H2SO4 (b) For ClO4− (c) NO3− (d) PO43− (e) SO42− (f) P2O74− (g) H3PO3 (h) C2H4 (i) O22− (j) CS2

27 | P a g e

You might also like