Sie sind auf Seite 1von 80

CHEMIST

RY

2MATTER AND ITS FORM


Matter is anything that has mass and occupies space, we were taught in
school. True enough, but not very satisfying. A really complete answer is
unfortunately beyond the scope of this course, but we will offer a hint of it in
a later chapter on atomic structure.
For the moment, lets put off trying to define matter and focus on the
chemists view: matter is what chemical substances are composed of. But
what do we mean by chemical substances? How do we organize our view of
matter and its properties? These very practical questions will be the subjects
of this lesson.

2.1 OBSERVABLE PROPERTIES OF MATTER


The science of chemistry developed from observations made about the
nature and behavior of different kinds of matter, which we refer to
collectively as the properties of matter.
The properties we refer to in this lesson are all macroscopic properties:
those that can be observed in bulk matter. At the microscopic level, matter is
of course characterized by its structure: the spatial arrangement of the
individual atoms in a molecular unit or an extended solid.
The study of matter begins with the study of its properties
By observing a sample of matter and measuring its various properties, we
gradually acquire enough information to characterize it; to distinguish it from
other kinds of matter. This is the first step in the development of chemical
science, in which interest is focussed on specific kinds of matter and the
transformations between them.

2.2 EXTENSIVE AND INTENSIVE PROPERTIES


If you think about the various observable properties of matter, it will
become apparent that these fall into two classes. Some properties, such as
mass and volume, depend on the quantity of matter in the sample we are
studying. Clearly, these properties, as important as they may be, cannot by
themselves be used to characterize a kind of matter; to say that water has a
mass of 2 kg is nonsense, although it may be quite true in a particular
instance. Properties of this kind are called extensive properties of matter.
This definition of the density illustrates an important general rule: the
ratio of two extensive properties is always an intensive property.
Suppose we make further measurements, and find that the same
quantity of water whose mass is 2.0 kg also occupies a volume of 2.0 liters.

We have measured two extensive properties (mass and volume) of the same
sample of matter. This allows us to define a new quantity, the quotient m/V
which defines another property of water which we call the density. Unlike the
mass and the volume, which by themselves refer only to individual samples
of water, the density (mass per unit volume) is a property of all samples of
pure water at the same temperature. Density is an example of an intensive
property of matter.
Intensive properties are extremely important, because every possible
kind of matter possesses a unique set of intensive properties that
distinguishes it from every other kind of matter. Some intensive properties
can be determined by simple observations: color (absorption spectrum),
melting point, density, solubility, acidic or alkaline nature, and density are
common examples. Even more fundamental, but less directly observable, is
chemical composition.
The more intensive properties we know, the more precisely we can
characterize a sample of matter. Intensive properties are extremely
important, because every possible kind of matter possesses a unique set of
intensive properties that distinguishes it from every other kind of matter. In
other words, intensive properties serve to characterize matter. Many
of the intensive properties depend on such variables as the temperature and
pressure, but the ways in which these properties change with such variables
can themselves be regarded as intensive properties.
Problem Example:
Classify each of the following as an extensive or intensive property.
1. The volume of beer in a mug
2. The percentage of alcohol in
the beer
3. The number of calories of
energy you derive from
eating a banana
4. The number of calories of
energy made available to
your body when you
consume 10.0 g of sugar
5. The mass of iron present in
your blood
6. The mass of iron present in 5
mL of your blood
7. The electrical resistance of a
piece of 22-gauge copper

ext; depends on size of


the mug.
int; same for any samesized sample.
ext; depends on size and
sugar content of the
banana.
int; same for any 10-g
portion of sugar.
ext; depends on volume
of blood in the body.
int; the same for any 5-mL
sample.
ext; depends on length of
the wire.

wire.
8. The electrical resistance of a
1-km length of 22-gauge
copper wire
9. The pressure of air in a
bicycle tire

int; same for any 1-km


length of the same wire.
pressure itself is
intensive, but is also
dependent on the
quantity of air in the tire.

The last example shows that not everything is black or white!


But we often encounter matter that is not uniform throughout, whose
different parts exhibit different sets of intensive properties. This brings up
another distinction that we address immediately below.
How Do We Classify Matter?
One useful way of organizing our understanding
of matter is to think of a hierarchy that extends down
from the most general and complex to the simplest
and most fundamental. The orange-colored boxes
represent the central realm of chemistry, which deals
ultimately with specific chemical substances, but as a
practical matter, chemical science extends both above
and below this region.

Alternatively, when we are thinking about specific


samples of matter, it may be more useful to re-cast
our classification in two dimensions:

Notice, in the bottom line of boxes above, that "mixtures" and "pure
substances" can fall into either the homogeneous or heterogeneous
categories.
Since this latter schema covers virtually all matter that we encounter in the
world, it is important that you thoroughly understand the meaning of each
part of it. We will begin by looking at the distinction represented in the top
line of the diagram.

2.3 HOMOGENEOUS AND HETEROGENEOUS: IT'S A MATTER OF


PHASES

Homogeneous matter (from the Greek homo =


same) can be thought of as being uniform and
continuous, whereas heterogeneous matter (hetero
=
different)
implies
non-uniformity
and
discontinuity. To take this further, we first need to
define "uniformity" in a more precise way, and this
takes us to the concept of phases.
A phase is a region of matter that possesses
uniform intensive properties throughout its volume.
A volume of water, a chunk of ice, a grain of
sand, a piece of copper each of these constitutes a
single phase, and by the above definition, is said to
be homogeneous.
A sample of matter can contain more than a
single phase; a cool drink with ice floating in it
consists of at least two phases, the liquid and the
ice. If it is a carbonated beverage, you can probably
see gas bubbles in it that make up a third phase.

The phase
boundaries in an oilvinegar salad
dressing are easy to
see...

but those in milk are


so small that they

2.4 PHASE BOUNDARIES


Each phase in a multiphase system is separated from its neighbors by a
phase boundary, a thin region in which the intensive properties change
discontinuously. Have you ever wondered why you can easily see the ice
floating in a glass of water although both the water and the ice are
transparent? The answer is that when light crosses a phase boundary, its

direction of travel is slightly bent, and a portion of the light gets reflected
back; it is these reflected and distorted light
rays emerging from that reveal the chunks
of ice floating in the liquid.
If, instead of visible chunks of material, the
second phase is broken into tiny particles, the light rays usually bounce off
the surfaces of many of these particles in random directions before they
emerge from the medium and are detected by the eye. This phenomenon,
known as scattering, gives multiphase systems of this kind a cloudy
appearance, rendering them translucent instead of transparent. Two very
common examples are ordinary fog, in which water droplets are suspended
in the air, and milk, which consists of butterfat globules suspended in an
aqueous solution.
Getting back to our classification,
we can say that
Homogeneous
matter
consists of a single phase
throughout
its
volume;
heterogeneous
matter
contains two or more phases.
Dichotomies
("either-or"
classifications) often tend to break down when closely examined, and the
distinction between homogeneous and heterogeneous matter is a good
example; this is really a matter of degree, since at the microscopic level all
matter is made up of atoms or molecules separated by empty space! For
most practical purposes, we consider matter as homogeneous when any
discontinuities it contains are too small to affect its visual appearance.
How large must a molecule or an agglomeration of molecules be before
it begins to exhibit properties of a being a separate phase? Such particles
span the gap between the micro and macro worlds, and have been known as
colloids since they began to be studied around 1900. But with the
development of nanotechnology in the 1990s, this distinction has become
even more fuzzy.

2.5

PURE SUBSTANCES AND MIXTURES

The air around us, most of the liquids and solids we encounter, and all
too much of the water we drink consists not of pure substances, but of
mixtures. You probably have a general idea of what a mixture is, and how it
differs from a pure substance; what is the scientific criterion for making this
distinction?

To a chemist, a pure substance usually refers to a sample of matter


that has a distinct set of properties that are common to all other samples of
that substance. A good example would be ordinary salt, sodium chloride. No
matter what its source (from a mine, evaporated from seawater, or made in
the laboratory), all samples of this substance, once they have been purified,
possess the same unique set of properties.
A pure substance is one whose intensive properties are the same in
any purified sample of that same substance.
A mixture, in contrast, is composed of two or more substances, and it
can exhibit a wide range of properties depending on the relative amounts of
the components present in the mixture. For example, you can dissolve up to
357 g of salt in one litre of water at room temperature, making possible an
infinite variety of "salt water" solutions. For each of these concentrations,
properties such as the density, boiling and freezing points, and the vapor
pressure of the resulting solution will be different.
Is anything really pure?
"9944100% Pure: It Floats"
This description of Ivory Soap is
a classic example of junk science from
the 19th century. Not only is the term
"pure" meaningless when applied to an
undefined mixture such as hand soap,
but the implication that its ability to
float is evidence of this purity is
deceptive. The low density is achieved
by beating air bubbles into it, actually
reducing the "purity" of the product and in a sense cheating the consumer.
Similarly, those of us who enjoy peanut butter would never willingly
purchase a brand advertised as "impure". But a Consumer Reports article
published some years ago showed a table listing the
number of "mouse droppings" and "insect parts"
(presumably from peanut storage silos) they found in
samples of all the major brands. Bon appetit!
Finally, we all prefer to drink "pure" water, but we
don't usually concern ourselves with the dissolved
atmospheric gases and ions such as Ca2+ and HCO3 that
are present in most drinking waters. But these harmless
"impurities" and are always present in those "pure"
spring waters.

However, these and similar substances could seriously interfere with


certain uses to which we put water in the laboratory, were we customarily
use distilled or deionized water.
Even these still contain some dissolved gases and occasionally some
silica, but their small amounts and relative inertness make these impurities
insignificant for most purposes. When water of the highest obtainable purity
is required for certain types of exacting measurements, it is commonly
filtered, de-ionized, and triple-vacuum distilled. But even this "chemically
pure" water is a mixture of isotopic species: there are two stable isotopes of
both hydrogen (H1 and H2, often denoted by D) and oxygen (O16 and O18)
which give rise to combinations such as H 2O18, HDO16, etc., all of which are
readily identifiable in the infrared spectra of water vapor. (Interestingly, the
ratio of O18/O16 in water varies enough from place to place that it is now
possible to determine the source of a particular water sample with some
precision.) And to top this off, the two hydrogen atoms in water contain
protons whose magnetic moments can be parallel or antiparallel, giving rise
to ortho- and para-water, respectively.
The bottom line: To a chemist, the term "pure" has meaning only in the
context of a particular application or process.

2.6 OPERATIONAL AND CONCEPTUAL CLASSIFICATIONS


Since chemistry is an experimental science, we need a set of
experimental criteria for placing a given sample of
matter in one of these categories. There is no
single experiment that will always succeed in
unambiguously deciding this kind of question.
However, there is one principle that will always
work in theory, if not in practice. This is based
on the fact that the various components of a
mixture can, in principle, always be separated
into pure substances.

the

Consider a heterogeneous mixture of salt water and


sand. The sand can be separated from the salt water by
mechanical process of filtration.
Similarly, the butterfat contained in milk may be
separated from the water by a process known as
churning, in which mechanical agitation forces the butterfat
droplets to coalesce into the solid mass we know as butter.
These examples illustrate the general principle that

Heterogeneous matter may be separated into homogeneous matter by


mechanical means.
Turning this around, we have an operational definition of
heterogeneous matter: if, by some mechanical operation we can separate a
sample of matter into two or more other kinds of matter, then our original
sample was heterogeneous.
To find a similar operational definition for homogeneous mixtures,
consider how we might separate the two components of a solution of salt
water. The most obvious way would be to evaporate off the water, leaving
the salt as a solid residue. Thus a homogeneous mixture can be separated
into pure substances by undergoing appropriate partial changes of state
that is, by evaporation, freezing, etc.
Note the term partial in the above sentence; in the last example, we
evaporate only the water, not the salt (which would be very difficult to do
anyway!) The idea is that one component of the mixture is preferentially
affected by the process we are carrying out. This principle will be
emphasized in the following examples.

2.7 ORGANIC COMPOUNDS


The chemical compounds of living things are known as organic
compounds because of their association with organisms and because they
are carbon-containing compounds. Organic compounds, which are the
compounds associated with life processes, are the subject matter of organic
chemistry. Among the numerous types of organic compounds, four major
categories are found in all living things: carbohydrates, lipids, proteins, and
nucleic acids.

2.7.1

CARBOHYDRATES

Almost all organisms use carbohydrates as sources of energy. In


addition, some carbohydrates serve as structural materials. Carbohydrates
are molecules composed of carbon, hydrogen, and oxygen; the ratio of
hydrogen atoms to oxygen and carbon atoms is 2:1.
Simple carbohydrates, commonly referred to as sugars, can be
monosaccharides if they are composed of single molecules, or
disaccharides if they are composed of two molecules. The most important
monosaccharide is glucose, a carbohydrate with the molecular formula
C6H12O6. Glucose is the basic form of fuel in living things. In multicellular
organisms, it is soluble and is transported by body fluids to all cells, where it
is metabolized to release its energy. Glucose is the starting material for
cellular respiration, and it is the main product of photosynthesis.

Three important disaccharides are also found in living things: maltose,


sucrose, and lactose. Maltose is a combination of two glucose units
covalently linked. The table sugar sucrose is formed by linking glucose to
another monosaccharide called fructose. Lactose is composed of glucose and
galactose units.
In the figure, Glucose and fructose
molecules
combine
to
form
the
disaccharide sucrose.
Complex carbohydrates are known as
polysaccharides. Polysaccharides are
formed
by
linking
innumerable
monosaccharides.
Among
the
most
important polysaccharides is starch, which
is composed of hundreds or thousands of
glucose units linked to one another. Starch
serves as a storage form for carbohydrates.
Much of the worlds human population
satisfies its energy needs with starch in the
form of rice, wheat, corn, and potatoes.

Two other important polysaccharides are glycogen and cellulose.


Glycogen is also composed of thousands of glucose units, but the units are
bonded in a different pattern than in starch. Glycogen is the form in which
glucose is stored in the human liver. Cellulose is used primarily as a
structural carbohydrate. It is also composed of glucose units, but the units
cannot be released from one another except by a few species of organisms.
Wood is composed chiefly of cellulose, as are plant cell walls. Cotton fabric
and paper are commercial cellulose products.

2.7.2

LIPIDS

Lipids are organic molecules composed of carbon, hydrogen, and


oxygen atoms. The ratio of hydrogen atoms to oxygen atoms is much higher
in lipids than in carbohydrates. Lipids include steroids (the material of which
many hormones are composed), waxes, and fats.
Fat molecules are composed of a glycerol molecule and one, two, or
three molecules of fatty acids. A glycerol molecule contains three hydroxyl (
OH) groups. A fatty acid is a long chain of carbon atoms (from 4 to 24) with
a carboxyl (COOH) group at one end. The fatty acids in a fat may all be alike
or they may all be different. They are bound to the glycerol molecule by a
process that involves the removal of water.

Certain fatty acids have one or more double bonds in their molecules.
Fats that include these molecules are unsaturated fats. Other fatty acids
have no double bonds. Fats that include these fatty acids are saturated fats.
In most human health situations, the consumption of unsaturated fats is
preferred to the consumption of saturated fats.
Fats stored in cells usually form clear oil droplets called globules
because fats do not dissolve in water. Plants often store fats in their seeds,
and animals store fats in large, clear globules in the cells of adipose tissue.
The fats in adipose tissue contain much concentrated energy. Hence, they
serve as a reserve energy supply to the organism. The enzyme lipase breaks
down fats into fatty acids and glycerol in the human digestive system.
A fat molecule is constructed by combining a
glycerol molecule with three fatty acid molecules.
(Two saturated fatty acids and one unsaturated
fatty acid are shown for comparison.) The
constructed molecule is at the bottom.

2.7.3PROTEINS
Proteins, among the most complex of all
organic compounds, are composed of amino
acids (see Figure 2-4), which contain carbon,
hydrogen, oxygen, and nitrogen atoms. Certain amino acids also have sulfur
atoms, phosphorus, or other trace elements such as iron or copper.

The structure and chemistry of amino acids. When two amino acids
are joined in a dipeptide, the OH of one amino acid is removed, and the H
of the second is removed. So, water is removed. A dipeptide bond (right)
forms to join the amino acids together.

Many proteins are immense and extremely complex. However, all


proteins are composed of long chains of relatively simple amino acids. There
are 20 kinds of amino acids. Each amino acid (see the left illustration in
Figure 2-4) has an amino (NH2) group, a carboxyl (COOH) group, and a
group of atoms called an R group (where R stands for radical). The amino
acids differ depending on the nature of the R group, as shown in the middle
illustration of Figure 2-4. Examples of amino acids are alanine, valine,
glutamic acid, tryptophan, tyrosine, and histidine.
The removal of water molecules links amino acids to form a protein.
The process is called dehydration synthesis, and a by-product of the
synthesis is water. The links forged between the amino acids are peptide
bonds, and small proteins are often called peptides.
All living things depend on proteins for their existence. Proteins are the
major molecules from which living things are constructed. Certain proteins
are dissolved or suspended in the watery substance of the cells, while others
are incorporated into various structures of the cells. Proteins are also found
as supporting and strengthening materials in tissues outside of cells. Bone,
cartilage, tendons, and ligaments are all composed of proteins.
One essential function of proteins is as an enzyme. Enzymes catalyze
the chemical reactions that take place within cells. They are not used up in a
reaction; rather, they remain available to catalyze succeeding reactions.
Every species manufactures proteins unique to that species. The
information for synthesizing the unique proteins is located in the nucleus of
the cell. The so-called genetic code specifies the amino acid sequence in
proteins. Hence, the genetic code regulates the chemistry taking place within
a cell. Proteins also can serve as a reserve source of energy for the cell.
When the amino group is removed from an amino acid, the resulting
compound is energy-rich.

2.8 SEPARATING HOMOGENEOUS MIXTURES


Some common methods of separating homogeneous mixtures into their
components are outlined below.

2.8.1

Distillation

A mixture of two volatile liquids is partly boiled away; the first portions
of the condensed vapor will be enriched in the component having the lower
boiling point.
Note that if all the liquid were boiled away, the distillate would be
identical with the original liquid. But if, say, half of the liquid is distilled, the
distillate would contain a larger fraction of the more volatile component.
If the distillate is then re-distilled, it can be further enriched in the lowboiling liquid. By repeating this process many times (aided by the
fractionating column above the boiling vessel), a high degree of separation
can be achieved.

2.7.2

Fractional crystallization

A hot saturated solution containing two or more dissolved solids is


allowed to cool slowly; the least-soluble material crystallizes out first, and
can be separated by filtration. This process is widely employed both in
the laboratory and, on a much larger scale, in industry.

Similarly, a molten mixture of several components, when slowly


cooled, will first yield crystals of the material having the highest melting
point. This process occurs on a huge scale in nature when molten magma
from the earth's mantle rises into the lithosphere and cools underground a
process that can take up to a million years.
This is how the common rock known as
granite is formed. Eventually these rocks
rise and become exposed on the earth's
surface.
Granite is a major component of California's
Sierra Nevada mountains, where it forms
the well-known peak Half Dome in Yosemite
Valley.

A magnified image of granite shows the


various components that have crystallized
out over time as the melt gradually cooled.

Note that because the various components of granite are not removed
as they crystallize out, they remain embedded in the solidified material.

2.7.3

Liquid-liquid extraction

Two mutually-insoluble liquids, one


containing two or more solutes (dissolved
substances), are shaken together in a
separatory funnel. Each solute will concentrate
the liquid in which it is more soluble. The two
solutions are then separated by opening the
stopcock at the bottom, allowing the more
dense solution to drain out.

in

2.7.4

Solid-liquid extraction

In working with natural products such as plant materials, a first step is


often to extract soluble substances from the plant parts.
This, and similar extractions of the soluble components of complex
solids, is carried out in an apparatus known as a Soxhlet extractor.
The idea is to continuously percolate an appropriate hot solvent
through the material, which is contained in a porous paper "thimble". Hot
vapor from the boiling flask bypasses the extraction chamber through the
arm at the left (labeled "vapor" in the illustration ) and into the condenser,
from which it drips down into the extraction chamber, where a portion of the
soluble material mixes with the solvent. When the condensate reaches the
top of the chamber, it flows out through the siphon arm, emptying its
contents into the boiling flask, which becomes increasingly concentrated in
the extracted material.
The advantage of this arrangement is that the percolation-andextraction process can be repeated indefinitely (usually hours to days)
without much attention.

2.7.5Chromatography

As a liquid or gaseous mixture flows along a column containing an


adsorbant material, the more strongly-adsorbed components tend to move
more slowly and emerge later than the less-strongly adsorbed components.
In this example, an extract made from plant leaves is separated into its
principle components: carotene, xanthophyll, and chlorophylls A and B.
Although chromatography originated in the mid-19th Century, it was
not widely employed until the 1950's. Since that time, it has encompassed a
huge variety of techniques and is no longer limited to colored substances.
Chromatography is now one of the most widely-employed methods for the
analysis and separation of complex mixtures of liquids and gases.

2.8 Physical and chemical properties

Since chemistry is partly the study of the transformations that matter


can undergo, we can also assign to any substance a set of chemical
properties that express the various changes of composition the substance is
known to undergo. Chemical properties also include the conditions of
temperature, etc., required to bring about the change, and the amount of
energy released or absorbed as the change takes place.
The properties that we described above are traditionally known as
physical properties, and are to be distinguished from chemical properties
that usually refer to changes in composition that a substance can undergo.
For example, we can state some of the more distinctive physical and
chemical properties of the element sodium:

PHYSICAL PROPERTIES
(25C)

CHEMICAL PROPERTIES
forms an oxide Na2O and a hydride NaH

appearance: a soft, shiny


metal
density: 0.97 g cm3
melting point: 97.5C
boiling point: 960C

burns in air to form sodium peroxide


Na2O2
reacts violently with water to release
hydrogen gas
dissolves in liquid ammonia to form a
deep blue solution

3GASES: MOVING MOLECULES

3.1 The Gas Laws


All of the gas laws rely on some basic assumptions that are made about
gases, and together they constitute what it means for a gas to be in an ideal
state. In an ideal state
1. All gas particles are in constant, random motion.

2. All collisions between gas particles are perfectly elastic (meaning that
the kinetic energy of the system is conserved).
3. The volume of the gas molecules in a gas is negligible.
4. Gases have no intermolecular attractive or repulsive forces.
5. The average kinetic energy of the gas is directly proportional to its
Kelvin temperature and is the same for all gases at a specified
temperature.
Only four measurable properties are used to describe a gas: its quantity,
temperature, volume, and pressure. The quantity (amount) of the gas is
usually expressed in moles (n). The temperature, T, of gases must always be
converted to the Kelvin temperature scale (the absolute temperature scale).
The volume, V, of a gas is usually given in liters. Finally, the pressure, P, of a
gas is usually expressed in atmospheres. Gases are often discussed in terms
of standard temperature and pressure (STP), which means 273K (or
0C) and 1 atm.
Example
Which of the following statements is not true of ideal gases?
1. The volume occupied by gas particles is only significant at very low
pressures.
2. Gas molecules occupy an insignificant volume compared to the volume
of the container that holds them.
3. The particles of a gas move in random straight line paths until a
collision occurs.
4. The collisions that occur between gas particles are considered elastic.
5. At a given temperature, all gas molecules within a sample possess the
same average kinetic energy.
Explanation
In this example, choice 1 is incorrect. Choices 2, 3, 4, and 5 all describe
an ideal gas. Choice 1 makes an incorrect assumption: it begins with a true
statement about volume not being very significant but then turns around and
gives the incorrect scenarioif the pressure is low, then gas particles
undergo very few collisions, so the volume is insignificant. The volume only

becomes significant if gas particles collide often, increasing the chances that
intermolecular forces will hold them together.
Measuring the Pressure of a Gas
Gas pressure is a gauge of the number and force of collisions between
gas particles and the walls of the container that holds them. The SI unit for
pressure is the pascal (Pa), but other pressure terms include atmospheres
(atms), millimeters of mercury (mmHg), and torr. The following is a list
of all of the standard pressure in every unit for pressure. Memorize these for
the exam so you can convert units where necessary:
760 mmHg
760 torr
1.00 atm
101,325 Pa
101.325 kPa
The piece of lab equipment specifically designed to measure the
pressure of gases is known as the barometer. A barometer uses the height
of a column of mercury to measure gas pressure in millimeters of mercury or
torr (1 mmHg = 1 torr). The mercury is pushed up the tube from the dish
until the pressure at the bottom of the tube (due to the mass of the mercury)
is balanced by the atmospheric pressure.

When using a barometer, you calculate gas pressure with the following
equation:
Gas pressure = atmospheric pressure - h (height of the mercury)
The open-tube manometer is another device that can be used to
measure pressure. The open-tube manometer is used to measure the
pressure of a gas in a container.

The pressure of the gas is given by h (the difference in mercury levels)


in units of torr or mmHg. Atmospheric pressure pushes on the mercury from
one direction, and the gas in the container pushes from the other direction.
In a manometer, since the gas in the bulb is pushing more than the
atmospheric pressure, you add the atmospheric pressure to the height
difference:
gas pressure = atmospheric pressure + h
There is one other possibility for a manometer question that could
appear on the SAT II Chemistry test: they could ask you about a closed-tube
manometer. Closed-tube manometers look similar to regular manometers
except that the end thats open to the atmospheric pressure in a regular
manometer is sealed and contains a vacuum. In these systems, the
difference in mercury levels (in mmHg) is equal to the pressure in torr.

3.2 BOYLES LAW

A graph of Boyle's
original data
This relationship between pressure and volume was first noted by two
new scientists, Richard Towneley and Henry Power. Robert Boyle confirmed
their discovery through experiments and published the results. According to
Robert Gunther and other authorities, it was Boyle's assistant, Robert Hooke,
who built the experimental apparatus. Boyle's law is based on experiments
with air, which he considered to be a fluid of particles at rest in between
small invisible springs. At that time, air was still seen as one of the four
elements, but Boyle disagreed. Boyle's interest was probably to understand

air as an essential element of life; for example, he published works on the


growth of plants without air. Boyle used a closed J-shaped tube and after
pouring mercury from one side he forced the air on the other side to contract
under the pressure of mercury. After repeating the experiment several times
and using different amount of mercury he found that under controlled
conditions, the pressure of a gas is inversely proportional to the volume
occupied by it.
The French physicist Edme Mariotte (16201684) discovered the same
law independent of Boyle in 1676, but Boyle had already published it in
1662. Thus this law is sometimes referred to as Mariotte's law or the Boyle
Mariotte law. Later, in 1687 in the Philosophi Naturalis Principia
Mathematica, Newton showed mathematically that if an elastic fluid
consisting of particles at rest, between which are repulsive forces inversely
proportional to their distance, the density would be directly proportional to
the pressure, but this mathematical treatise is not the physical explanation
for the observed relationship. Instead of a static theory a kinetic theory is
needed, which was provided two centuries later by Maxwell and Boltzmann.
This law was the first physical law to be expressed in the form of an
equation describing the dependence of two variable quantities.
Boyles law simply states that the volume of a confined gas at a fixed
temperature is inversely proportional to the pressure exerted on the gas.
This can also be expressed as PV = a constant. This makes sense if you think
of a balloon. When the pressure around a balloon increases, the volume of
the balloon decreases, and likewise, when you decrease the pressure around
a balloon, its volume will increase.
Boyles law to can also be expressed in the following way, and this is the
form of the law that you should memorize:
P1 V 1 = P2 V 2
Example:
Sulfur dioxide (SO2) gas is a component of car exhaust and power plant
discharge, and it plays a major role in the formation of acid rain. Consider a
3.0 L sample of gaseous SO 2 at a pressure of 1.0 atm. If the pressure is
changed to 1.5 atm at a constant temperature, what will be the new volume
of the gas?
Explanation
If P1V1 = P2V2, then (1.0 atm) (3.0 L) = (1.5 atm) (V2), so V2 = 2.0 L.
This answer makes sense according to Boyles lawas the pressure of the
system increases, the volume should decrease.

3.3 CHARLESS LAW


It was first published by French natural
philosopher Joseph Louis Gay-Lussac in 1802,
although he credited the discovery to
unpublished work from the 1780s by Jacques
Charles.
The
law
was
independently
discovered by British natural philosopher John
Dalton by 1801, although Dalton's description
was less thorough than Gay-Lussac's. The
basic principles had already been described a
century earlier by Guillaume Amontons.
Gay-Lussac was the first to demonstrate
that the law applied generally to all gases,
and to the vapours of volatile liquids if the
temperature was more than a few degrees above the boiling point. His
statement of the law can be expressed mathematically as:

where V100 is the volume occupied by a given sample of gas at 100 C; V0 is


the volume occupied by the same sample of gas at 0 C; and k is a constant
which is the same for all gases at constant pressure. Gay-Lussac's value for k
was 12.6666, remarkably close to the present-day value of 12.7315. This law was
first given by J. Charles in 1787.
Charless law states that if a given quantity of gas is held at a constant
pressure, its volume is directly proportional to the absolute temperature.
Think of it this way. As the temperature of the gas increases, the gas
molecules will begin to move around more quickly and hit the walls of their
container with more forcethus the volume will increase. Keep in mind that
you must use only the Kelvin temperature scale when working with
temperature in all gas law formulas! Heres the expression of Charless law
that you should memorize:

Try using Charless law to solve the following problem.

Example
A sample of gas at 15C and 1 atm has a volume of 2.50 L. What
volume will this gas occupy at 30C and 1 atm?
Explanation
The pressure remains the same, while the volume and temperature change
this is the hallmark of a Charless law question.

So,

, then 2.50 L/288K = V2/303K, and V2 = 2.63 L

This makes sensethe temperature is increasing slightly, so the


volume should increase slightly. Be careful of questions like thisits
tempting to just use the Celsius temperature, but you must first convert to
Kelvin temperature (by adding 273) to get the correct relationships!

3.4 AVOGADROS LAW

Avogadro's Law (Avogadro's theory; Avogadro's hypothesis) is a


principle stated in 1811 by the Italian chemist Amedeo Avogadro (17761856) that "equal volumes of gases at the same temperature and pressure
contain the same number of molecules regardless of their chemical nature
and physical properties". This number (Avogadro's number) is 6.022 X 10 23. It
is the number of molecules of any gas present in a volume of 22.41 L and is

the same for the lightest gas (hydrogen) as for a heavy gas such as carbon
dioxide or bromine.
The law can be stated mathematically

.
where:
V is the volume of the gas.
n is the amount of substance of the gas.
k is a proportionality constant.
The most important consequence of Avogadro's law is that the ideal gas
constant has the same value for all gases. This means that the constant

where:
p is the pressure of the gas
T is the temperature of the gas
has the same value for all gases, independent of the size or mass of the gas
molecules.
One mole of an ideal gas occupies 22.4 liters (dm) at STP, and
occupies 24.45 litres at SATP (Standard Ambient Temperature and Pressure =
273K and 1 atm or 101.325 kPa). This volume is often referred to as the
molar volume of an ideal gas. Real gases may deviate from this value.
Or to put it another way "the principle that equal volumes of all gases
at the same temperature and pressure contain the same number of
molecules. Thus, the molar volume of all ideal gases at 0 C and a pressure
of 1 atm. is 22.4 liters"
Avogadro's number is one of the fundamental constants of chemistry. It
permits calculation of the amount of pure substance (mole), the basis of
stoichiometric relationships. It also makes possible determination of how
much heavier a simple molecule of one gas is than that of another, as a
result the relative molecular weights of gases can be ascertained by
comparing the weights of equal volumes.

Avogadro's number (conventionally represented by N' in chemical


calculations) is now considered to be the number of atoms present in 12
grams of the carbon-12 isotope (one mole of carbon 12) and can be applied
to any type of chemical entity.

3.5 THE IDEAL GAS LAW


The ideal gas law is the most important gas law for you to know: it
combines all of the laws you learned about in this chapter thus far, under a
set of standard conditions. The four conditions used to describe a gas
pressure, volume, temperature, and number of moles (quantity)are all
related, along with R, the universal gas law constant, in the following
formula:

PV = nRT

where
P = pressure (atm), V = volume (L),
n = number of moles (mol),
R = 0.08206 L atm/mol K, and
T = temperature (K).
Now try an example using the ideal gas law equation.

Example
A 16.0 g sample of methane gas, CH4, the gas used in chemistry lab, has a
volume of 5.0 L at 27C. Calculate the pressure.
Explanation
Looking at all the information given, you have a mass, a volume, and a temperature, and
you need to find the pressure of the system. As always, start by checking your units. You must
first convert 16.0 g of CH4 into moles: 16.0 g CH4
1 mol CH4/16.0 g CH4 = 1 mol of methane.
The volume is in the correct units, but you must convert the temperature into Kelvins: 27 + 273 =
300K. Now youre ready to plug these numbers into the ideal gas law equation:
PV = nRT
(P) (5.0 L) = (1.0 mol) (0.0821 L

atm/mol

K) (300K), so P = 4.9 atm

3.6 DALTONS LAW OF PARTIAL PRESSURE

Among the experiments that led John Dalton to propose the atomic
theory were his studies of mixtures of gases. In 1803 Dalton summarized his
observations as follows: For a mixture of gases in a container, the total
pressure exerted is the sum of the pressures that each gas would exert if it
were alone. This statement, known as Daltons law of partial pressures,
can be expressed as follows:

Ptotal = P1 + P2 + . . . Pn
where the subscripts refer to the individual gases (gas 1, gas 2, and so on).
The symbols
P1, P2, P3, and so on represent each partial pressure, the pressure that a
particular gas
would exert if it were alone in the container.
Assuming that each gas behaves ideally, the partial pressure of each
gas can be calculated from the ideal gas law:

The total pressure of the mixture PTOTAL can be represented as

where nTOTAL is the sum of the numbers of moles of the various


gases. Thus, for a mixture of ideal gases, it is the total number of moles of

particles that is important, not the identity or composition of the involved


gas particles.

3.7 GRAHAMS LAW OF DIFFUSION AND EFFUSION


When examining the ideal gas laws in conjunction with the kinetic
theory of gases, we gain insights into the behavior of ideal gas. We can then
predict how gas particles behaviors such as gas molecular speed, effusion
rates, distances traveled by gas molecules. Graham's Law, which was
formulated by the Scottish physical chemist Thomas Graham, is an
important law that connects gas properties to the kinetic theory of gases.
Kinetic Molecular Theory
The Kinetic Molecular Theory states that the average energy of
molecules is proportional to absolute temperature as illustrated by the
following equation:

Where:
ek is the average translation kinetic energy,
R is the gas constant,
NA is Avogadro's number, and
T is temperature in Kelvins.

Since R and NA are constants, this means that the Kelvin temperature
(T) of a gas is directly proportional to the average kinetic energy of its
molecules. This means that at a given temperature, different gases (for
example He or O2) will the same average kinetic energy.
Graham's Law
Gas molecules move constantly and randomly throughout the volume
of the container they occupy. When examining the gas molecules
individually, we see that not all of the molecules of a particular gas at a
given temperature move at exactly the same speed. This means that each
molecule of a gas have slightly different kinetic energy. To calculate the

average kinetic energy (eK) of a sample of a gas, we use an average speed of


the gas, called the root mean square speed (urms).

with

eK is the kinetic energy measures in Joules

m is mass of a molecule of gas (kg)

urms is the root mean square speed (m/s)

The root mean square speed, urms, can be determined from the
temperature and molar mass of a gas.

with

R the ideal gas constant (8.314 kg*m2/s2*mol*K)

T temperature (Kelvin)

M molar mass (kg/mol)

When examining the root mean square speed equation, we can see that
the changes in temperature (T) and molar mass (M) affect the speed of the
gas molecules. The speed of the molecules in a gas is proportional to the
temperature and is inversely proportional to molar mass of the gas. In other
words, as the temperature of a sample of gas is increased, the molecules
speed up and the root mean square molecular speed increases as a result.

Graham's Law states that the rate of effusion of two different gases at the
same conditions are inversely proportional to the square roots of their molar
masses as given by the following equation:

In according with the Kinetic Molecular Theory, each gas molecule


moves independently. However, the net rate at which gas molecules move
depend on their average speed. By examining the equation above, we can
conclude that the heavier the molar mass of the gas molecules slower the
gas molecules move. And conversely, lighter the molar mass of the gas
molecules the faster the gas molecules move.
Molecular Diffusion
Similar to effusion, the process of
diffusion is the spread of gas molecules through
space or through a second substance such as the
atmosphere.
In the figure, the passage of one gas
through another substance. In this example, the
other substance is another gas.

Diffusion has many useful applications.


Here is an example of diffusion that is use in
everyday households. Natural gas is odorless
and used commercially daily. An undetected
leakage can be very dangerous as it is highly
flammable and can cause an explosion when it
comes in contact with an ignition source. In
addition, the long term breathing of natural gas
can lead to asphyxiation.

Fortunately, chemists have discovered a way to easily detect natural


gas occur leak by adding a small quantity of a gaseous organic sulfur
compound named methyl mercaptan, CH3SH, to the natural gas. When a leak
happens, the diffusion of the odorous methyl mercaptan in the natural gas
will serve as a sign of warning.

Example:
Calculate the root mean square speed, urms , in m/s of helium at 30oC.
Solution:
Start by converting the molar mass for helium from g/mol to kg/mol.

Now, using the equation for urms substitute in the proper values for each
variable and perform the calculation.

THE DEVELOPMENT OF THE ATOMIC


THEORY

4.1 THE ATOM IS PROPOSED


A few decades after Empedocles, Democritus, another Greek who lived
from 460 BCE to 370 BCE, developed a new
theory of matter that attempted to overcome
the problems of his predecessor. Democritus's
ideas were based on reasoning rather than
science, and drew on the teachings of two
Greek philosophers who came before him:
Leucippus and Anaxagoras. Democritus knew
that if you took a stone and cut it in half, each
half had the same properties as the original
stone. He reasoned that if you continued to cut
the stone into smaller and smaller pieces, at
some point you would reach a piece so tiny that
it could no longer be divided. Democritus called
these infinitesimally small pieces of matter
atomos, meaning 'indivisible'. He suggested
that atomos were eternal and could not be
destroyed. Democritus theorized that atomos
were specific to the material that they made
up, meaning that the atomos of stone were unique to stone and different
from the atomos of other materials, such as fur. This was a remarkable
theory that attempted to explain the whole physical world in terms of a small
number of ideas.

Stone

Fur

Ultimately, though, Aristotle and Plato, two of the best-known


philosophers of Ancient Greece, rejected the theories of Democritus. Aristotle
accepted the theory of Empedocles, adding his own (incorrect) idea that the
four core elements could be transformed into one another. Because of
Aristotle's great influence, Democritus's theory would have to wait almost
2,000 years before being rediscovered.

4.2 DALTONS ATOMIC THEORY


Although the concept of atoms is now widely
accepted, this wasnt always the case.
Scientists didnt always believe that everything was
composed of small particles called atoms. The work
of several scientists and their experimental data gave
evidence for what is
now called the atomic theory. In the late 1700s,
Antoine Lavoisier, a French scientist, experimented
with the reactions of many metals. He carefully
measured the mass of a substance before reacting
and again measured the mass after a reaction had
occurred in a closed system (meaning that nothing
could enter or leave the container). He found that no
matter what reaction he looked at, the mass of the
starting materials was always equal to the mass of
the ending materials.

Unlike the Greek


philosophers, John
Dalton believed in
both logical
thinking and
experimentation.

This is now called the law of conservation of


mass. This went contrary to what many scientists at the time thought. For
example, when a piece of iron rusts, it appears to gain mass. When a log is
burned, it appears to lose mass. In these examples, though, the reaction
does not take place in a closed container and substances, such as the gases
in the air, are able to enter or leave. When iron rusts, it is combining with
oxygen in the air, which is why it seems to gain mass. What Lavoisier found
was that no mass was actually being gained or lost. It was coming from the
air. This was a very important first step in giving evidence for the idea that
everything is made of atoms. The atoms (and mass) are not being created or
destroyed. The atoms are simply reacting with other atoms that are already
present.

In the late 1700s and early 1800s, scientists began noticing that when
certain substances, like hydrogen and oxygen, were combined to produce a
new substance, like water, the reactants (hydrogen and oxygen) always
reacted in the same proportions by mass. In other words, if 1 gram of
hydrogen reacted with 8 grams of oxygen, then 2 grams of hydrogen would
react with 16 grams of oxygen, and 3 grams of hydrogen would react with 24
grams of oxygen. Strangely, the observation that hydrogen and oxygen
always reacted in the same proportions by mass wasnt special. In fact, it
turned out that the reactants in every chemical reaction reacted in the same
proportions by mass. This observation is summarized in the law of definite
proportions. Take, for example, nitrogen and hydrogen, which react to
produce ammonia. In chemical reactions, 1 gram of hydrogen will react with
4.7 grams of nitrogen, and 2 grams of hydrogen will react with 9.4 grams of
nitrogen. Can you guess how much nitrogen would react with 3 grams of
hydrogen? Scientists studied reaction after reaction, but every time the
result was the same. The reactants always reacted in the same proportions.
At the same time that scientists were finding this pattern out, a man named
John Dalton was experimenting with several reactions in which the reactant
elements formed more than one type of product, depending on the
experimental conditions he used. One common reaction that he studied was
the reaction between carbon and oxygen. When carbon and oxygen react,
they produce two different substances well call these substances A and
B. It turned out that, given the same amount of carbon, forming B always
required exactly twice as much oxygen as forming A. In other words, if you
can make A with 3 grams of carbon and 4 grams of oxygen, B can be made
with the same 3 grams of carbon, but with 8 grams oxygen. Dalton asked
himself why does B require 2 times as much oxygen as A? Why not 1.21
times as much oxygen, or 0.95 times as much oxygen? Why a whole number
like 2? The situation became even stranger when Dalton tried similar
experiments with different substances.
For example, when he reacted nitrogen and oxygen, Dalton discovered
that he could make three different substances well call them C, D, and
E. As it turned out, for the same amount of nitrogen, D always required
twice as much oxygen as C. Similarly, E always required exactly four times as
much oxygen as C. Once again, Dalton noticed that small whole numbers

(2and 4) seemed to be the rule. This observation came to be known as the


law of multiple proportions.
Dalton thought about his results and tried to find some theory that
would explain it, as well as a theory that would explain the Law of
Conservation of Mass (mass is neither created nor destroyed, or the mass
you have at the beginning is equal to the mass at the end of a change). One
way to explain the relationships that Dalton and others had observed was to
suggest that materials like nitrogen, carbon and oxygen were composed of
small, indivisible quantities which Dalton called atoms (in reference to
Democritus original idea). Dalton used this idea to generate what is now
known as Daltons Atomic Theory which stated the following:
1.
2.
3.
4.
5.

Matter is made of tiny particles called atoms.


Atoms are indivisible (cant be broken into smaller particles).
During a chemical reaction, atoms are rearranged, but they do
not break apart, nor are they created or destroyed.
All atoms of a given element are identical in mass and other
properties.
The atoms of different elements differ in mass and other
properties.
Atoms of one element can combine with atoms of another
element to form compounds new, complex particles. In a
given compound, however, the different types of atoms are
always present in the same relative numbers.

4.3 THOMPSONS PLUM PUDDING MODEL


Dalton's Atomic Theory held up well to a lot of the different chemical
experiments that scientists performed to test it. In fact, for almost 100 years,
it seemed as if Dalton's Atomic Theory was the whole truth. However, in
1897, a scientist named J. J. Thomson conducted some research that

suggested that Daltons Atomic Theory wasnt the


entire story. As it turns out, Dalton had a lot right.
He was right in saying matter is made up of atoms;
he was right in saying there are ifferent kinds of
atoms with different mass and other properties; he
was almost right in saying atoms of a given
element are identical; he was right in saying during
a chemical reaction, atoms are merely rearranged;
he was right in saying a given compound always
has atoms present in the same relative numbers.
But he was WRONG in saying atoms were indivisible
or indestructible. As it turns out, atoms are divisible.
In fact, atoms are composed of even smaller, more
fundamental particles. These particles, called
subatomic particles, are particles that are smaller
than the atom. Well talk about the discoveries of
these subatomic particles next.

J.J. Thomson
conducted
experiments that
suggested that
Daltons atomic
theory wasnt
telling the entire
story.

In the mid-1800s, scientists were beginning to


realize that the study of chemistry and the study of
electricity were actually related. First, a man named Michael Faraday showed
how passing electricity through mixtures of different chemicals could cause
chemical reactions. Shortly after that, scientists found that by forcing
electricity through a tube filled with gas, the electricity made the gas glow!
Scientists didnt, however, understand the relationship between chemicals
and electricity until a British physicist named J. J. Thomson began
experimenting with what is known as a cathode ray tube.

The figure shows a basic diagram of a cathode ray tube like the one J. J.
Thomson would have used. A cathode ray tube is a small glass tube with a

cathode (a negatively charged metal plate) and an anode (a positively


charged metal plate) at opposite ends. By separating the cathode and anode
by a short distance, the cathode ray tube can generate what are known as
cathode rays rays of electricity that flow from the cathode to the anode. J. J.
Thomson wanted to know what cathode rays were, where cathode rays
came from, and whether cathode rays had any mass or charge. The
techniques that J. J. Thomson used to answer these questions were very
clever and earned him a Nobel Prize in physics. First, by cutting a small hole
in the anode, J. J. Thomson found that he could get some of the cathode rays
to flow through the hole in the anode and into the other end of the glass
cathode ray tube. Next, J. J. Thomson figured out that if he painted a
substance known as phosphor onto the far end of the cathode ray tube, he
could see exactly where the cathode rays hit because the cathode rays made
the phosphor glow. J. J. Thomson must have suspected that cathode rays
were charged, because his next step was to place a positively charged metal
plate on one side of the cathode ray tube and a negatively charged metal
plate on the other side of the cathode ray tube, as shown in the figure. The
metal plates didnt actually touch the cathode ray tube, but they were close
enough that a remarkable thing happened! The flow of the cathode rays
passing through the hole in the anode was bent upwards towards the
positive metal plate and away from the negative metal plate. Using the
opposite charges attract, like charges repel rule, J. J. Thomson argued that
if the cathode rays were attracted to the positively charged metal plate and
repelled from the negatively charged metal plate, they themselves must
have a negative charge!
J. J. Thomson then did some rather complex experiments with magnets,
and used his results to prove that cathode rays were not only negatively
charged, but also had mass. Remember that anything with mass is part of
what we call matter. In other words, these cathode rays must be the result of
negatively charged matter flowing from the cathode to the anode. But
there was a problem. According to J. J. Thomsons measurements, either
these cathode rays had a ridiculously high charge, or else had very, very
little mass much less mass than the smallest known atom. How was this
possible? How could the matter making up cathode rays be smaller than an
atom if atoms were indivisible? J. J. Thomson made a radical proposal: maybe
atoms are divisible. J. J. Thomson suggested that the small, negatively
charged particles making up the cathode ray were actually pieces of atoms.
He called these pieces corpuscles, although today we know them as
electrons. Thanks to his clever experiments and careful reasoning, J. J.
Thomson is credited with the discovery of the electron.
Now imagine what would happen if atoms were made entirely of
electrons. First of all, electrons are very, very small; in fact, electrons are
about 2,000 times smaller than the smallest known atom, so every atom

would have to contain a whole lot of electrons. But theres another, even
bigger problem: electrons
are negatively charged.
Therefore, if atoms were
made
entirely
out
of
electrons, atoms would be
negatively
charged
themselves
and
that
would mean all matter was
negatively charged as well.
Of course, matter isnt
Thomsons plum pudding model was much
negatively charged. In fact,
like a chocolate chip cookie. Notice how
most matter is what we call
the chocolate chips are the negatively
neutral it has no charge at
charged electrons, while the positive
all. If matter is composed of
charge is spread throughout the entire
atoms, and atoms are
batter.
composed
of
negative
electrons, how can matter
be neutral? The only possible explanation is that atoms consist of more than
just electrons. Atoms must also contain some type of positively charged
material that balances the negative charge on the electrons. Negative and
positive charges of equal size cancel each other out, just like negative and
positive numbers of equal size. What do you get if you add +1 and -1? You
get 0, or nothing. Thats true of numbers, and thats also true of charges. If
an atom contains an electron with a -1 charge, but also some form of
material with a +1 charge, overall the atom must have a (+1) + (-1) = 0
charge in other words, the atom must be neutral, or have no charge at all.
Based on the fact that atoms are neutral, and based on J. J. Thomsons
discovery that atoms contain negative subatomic particles called
electrons, scientists assumed that atoms must also contain a positive
substance. It turned out that this positive substance was another kind of
subatomic particle, known as the proton. Although scientists knew that
atoms had to contain positive material, protons werent actually discovered,
or understood, until quite a bit later. When Thomson discovered the negative
electron, he realized that atoms had to contain positive material as well
otherwise they wouldnt be neutral overall. As a result, Thomson formulated
whats known as the plum pudding model for the atom. According to the
plum pudding model, the negative electrons were like pieces of fruit and
the positive material was like the batter or the pudding. This made a lot of
sense given Thomsons experiments and observations. Thomson had been
able to isolate electrons using a cathode ray tube; however he had never
managed to isolate positive particles. As a result, Thomson theorized that
the positive material in the atom must form something like the batter in a
plum pudding, while the negative electrons must be scattered through this
batter. (If youve never seen or tasted a plum pudding, you can think of a

chocolate chip cookie instead. In that case, the positive material in the atom
would be the batter in the chocolate chip cookie, while the negative
electrons would be scattered through the batter like chocolate chips.)
Notice how easy it would be to pick the pieces of fruit out of a plum
pudding. On the other hand, it would be a lot harder to pick the batter out of
the plum pudding, because the batter is everywhere. If an atom were similar
to a plum pudding in which the electrons are scattered throughout the
batter of positive material, then youd expect it would be easy to
pick out the electrons, but a lot harder to pick out the positive material.
J.J. Thomson had measured the charge to mass ratio of the electron,
but had been unable to accurately measure the charge on the electron. With
his oil drop experiment, Robert Millikan was able to accurately measure the
charge of the electron. When combined with the charge to mass ratio, he
was able to calculate the mass of the electron. What Millikan did was to put a
harge on tiny droplets of oil and measured their rate of descent. By varying
the charge on different drops, he noticed that the electric charges on the
drops were all multiples of 1.6x10-19C, the charge on a single electron.

4.4 RUTHERFORDS MODEL


Rutherford overturned Thomson's model in 1911 with his well-known
gold foil experiment in which he demonstrated that the
atom has a
tiny,
heavy
nucleus.
Rutherford
designed
an
experiment to use the alpha particles emitted by a
radioactive element as probes to the unseen world of
atomic structure.
Rutherford presented his own physical model for
subatomic structure, as an interpretation for the
unexpected experimental results. In it, the atom is
made
up of a central charge (this is the modern atomic
nucleus, though Rutherford did not use the term
"nucleus" in his paper) surrounded by a cloud of
(presumably) orbiting electrons. In this May 1911 paper, Rutherford only
commits himself to a small central region of very high positive or negative
charge in the atom.
For concreteness, consider the passage of a high speed particle
through an atom having a positive central charge N e, and surrounded by a
compensating charge of N electrons.

From purely energetic considerations of how far particles of known


speed would be able to penetrate toward a central charge of 100 e,
Rutherford was able to calculate that the radius of his gold central charge
would need to be less (how much less could not be told) than 3.4 x 10 14
meters. This was in a gold atom known to be 10 10 meters or so in radiusa
very surprising finding, as it implied a strong central charge less than
1/3000th of the diameter of the atom.
The Rutherford model served to concentrate a great deal of the atom's
charge and mass to a very small core, but didn't attribute any structure to
the remaining electrons and remaining atomic mass. It did mention the
atomic model of Hantaro Nagaoka, in which the electrons are arranged in
one or more rings, with the specific metaphorical structure of the stable rings
of Saturn. The plum pudding model of J.J. Thomson also had rings of orbiting
electrons. Jean Baptiste Perrin claimed in his Nobel Lecture [3] that he was the
first one to suggest the model in his paper dated 1901.
The Rutherford paper suggested that the central charge of an atom
might be "proportional" to its atomic mass in hydrogen mass units u (roughly
1/2 of it, in Rutherford's model). For gold, this mass number is 197 (not then
known to great accuracy) and was therefore modeled by Rutherford to be
possibly 196 u. However, Rutherford did not attempt to make the direct
connection of central charge to atomic number, since gold's "atomic
number" (at that time merely its place number in the periodic table) was 79,
and Rutherford had modeled the charge to be about + 100 units (he had
actually suggested 98 units of positive charge, to make half of 196). Thus,
Rutherford did not formally suggest the two numbers (periodic table place,
79, and nuclear charge, 98 or 100) might be exactly the same.
A month after Rutherford's paper appeared, the proposal regarding the
exact identity of atomic number and nuclear charge was made by Antonius
van den Broek, and later confirmed experimentally within two years, by
Henry Moseley.

Keypoints
The atom's electron cloud does not influence alpha particle scattering.

Much of an atom's charge (specifically, its positive charge) is


concentrated in a relatively tiny volume at the center of the atom,
known today as the nucleus. The magnitude of this charge is
proportional to (up to a charge number that can be approximately half
of) the atom's atomic mass - the remaining mass is now known to be
mostly attributed to neutrons. This concentrated central mass and
charge is responsible for deflecting both alpha and beta particles.

The mass of heavy atoms such as gold is mostly concentrated in the


central charge region, since calculations show it is not deflected or
moved by the high speed alpha particles, which have very high
momentum in comparison to electrons, but not with regard to a heavy
atom as a whole.

The atom itself is about one trillion (10^12) the size of the nucleus.
That's like putting a marble in the center of a football field.

4.5 BOHRS ATOMIC MODEL


To explain the structure of the atom, the Danish physicist Niels Bohr
developed in 1913 a hypothesis known as the Bohr theory of the atom. He
assumed that electrons are arranged in definite
shells, or quantum levels, at a considerable
distance from the nucleus. The arrangement of
these
electrons
is
called
the
electron
configuration. The number of such electrons
equals the atomic number of the atom; hydrogen
has a single orbital electron, helium has 2, and
uranium has 92. The electron shells are built up in
a regular fashion from a first shell to a total of
seven shells, each of which has an upper limit to
the number of electrons that it can accommodate.
The first shell is complete with two electrons, the
second can hold up to eight electrons, and successive shells hold still larger
numbers. The "last" electrons, those which are outermost or added last to
the atom's structure, determine the chemical behavior of the atom.
The inert, or noble, gases (helium, neon, argon, krypton, xenon, and
radon) all have completely filled outer shells. They do not enter into chemical
combinations in nature, although the three heaviest inert gases (krypton,
xenon, and radon) have formed chemical compounds in the laboratory. On
the other hand, the outermost shells of such elements as lithium, sodium,
and potassium contain only one electron. These elements combine readily

with other elements (transferring their outermost electrons to them) to form


a great many chemical compounds.

Atomic shells
do not necessarily
fill up with
electrons
in
consecutive
order.
The
electrons of the first 18
elements
in
the periodic table are added
in a regular manner, each shell being filled to a designated limit before a
new shell is started. Beginning with the 19th element, the outermost electron
starts a new shell before the previous shell is completely filled. A regularity is
still maintained, however, as electrons fill successive shells in a repetitious
back-and-forth pattern. The result is the regular repetition of chemical
properties for atoms of increasing atomic weight that corresponds to the
arrangement of the elements in the periodic table.
It is convenient to visualize the electrons moving about the nucleus of
an atom much as if they were planets moving about the sun. This view is
much more precise than that held by contemporary physicists, however. It is
now known that it is impossible to pinpoint the precise position of an electron
in the atom's space without disturbing its predicted location at some future
time. This uncertainty is resolved by attributing to the atom a cloudlike form,
in which the electron's position is defined in terms of the probability of
finding it at some distance from the nucleus. This rather fuzzy schematic
conception of the atom may be reconciled with the solar-system model by
noting that in the tiny space of the atom the electron, which makes many
billions of orbits around the nucleus in a single second, is everywhere at
once. The cloud view thus gives a form to the atom that is not supplied by a
solar-system model.

4.6 SUBATOMIC PARTICLES

In the years since Thomson and Rutherford, a great deal has been
learned about atomic structure. Because much of this material will be
covered in detail in later chapters, only an introduction will be given here.
The simplest view of the atom is that it consists of a
tiny nucleus (with a diameter of about 10_13 cm) and electrons that move
about the nucleus at an average distance of about 10_8 cm from it.

As we will see later, the chemistry of an atom mainly results from its
electrons. For this reason, chemists can be satisfied with a relatively crude
nuclear model. The nucleus is assumed to contain protons, which have a
positive charge equal in magnitude to the electrons negative charge, and
neutrons, which have virtually the same mass as a proton but no charge.
The masses and charges of the electron, proton, and neutron are shown in
Table.

Two striking things about the nucleus are its small size compared with
the overall size of the atom and its extremely high density. The tiny nucleus
accounts for almost all the atoms mass. Its great density is dramatically
demonstrated by the fact that a piece of nuclear material about the size of a
pea would have a mass of 250 million tons!

An important question to consider at


this point is, If all atoms are composed of
these same components, why do different
atoms
have
different
chemical
properties? The answer to this question
lies in the number and the arrangement of
the electrons. The electrons constitute
most of the atomic volume and thus are
the parts that intermingle when atoms
combine to form molecules. Therefore, the
number of electrons possessed by a given
atom greatly affects its ability to interact
with other atoms. As a result, the atoms of
different elements, which have different
numbers of protons and electrons, show different chemical behavior.
A sodium atom has 11 protons in its nucleus. Since atoms have no net
charge, the number of electrons must equal the number of protons.
Therefore, a sodium atom has 11 electrons moving around its nucleus. It is
always true that a sodium atom has 11 protons and 11 electrons. However,
each sodium atom also has neutrons in its nucleus, and different types of
sodium atoms exist that have different numbers of neutrons. For example,
consider the sodium atoms represented in Fig. 2.15. These two atoms are
isotopes, or atoms with the same number of protons but different numbers
of neutrons. Note that the symbol for one particular type of sodium atom is
written

where the atomic number Z (number of protons) is written as a subscript,


and the mass number A (the total number of protons and neutrons) is
written as a superscript. (The particular atom represented here is called
sodium twenty-three. It has 11 electrons, 11 protons, and 12 neutrons.)
Because the chemistry of an atom is due to its electrons, sotopes show
almost identical chemical properties. In nature most elements contain
mixtures of isotopes.

4.7 THE ATOM AND ITS PARTS

Atoms are the basic units of matter and the defining structure of
elements. Atoms are made up of three particles: protons, neutrons and
electrons.
Protons and neutrons are heavier than electrons and reside in the
center of the atom, which is called the nucleus. Electrons are extremely
lightweight and exist in a cloud orbiting the nucleus. The electron cloud has a
radius 10,000 times greater than the nucleus.
Protons and neutrons have approximately the same mass. However,
one proton weighs more than 1,800 electrons. Atoms always have an equal
number of protons and electrons, and the number of protons and neutrons is
usually the same as well. Adding a proton to an atom makes a new element,
while adding a neutron makes an isotope, or heavier version, of that atom.

4.7.1

Nucleus

The nucleus was discovered in 1911, but its parts were not identified
until 1932. Virtually all the mass of the atom resides in the nucleus. The
nucleus is held together by the "strong force," one of the four basic forces in
nature. This force between the protons and neutrons overcomes the
repulsive electrical force that would, according to the rules of electricity,
push the protons apart otherwise.

4.7.2

Protons

Protons are positively charged particles found within atomic nuclei.


They were discovered by Ernest Rutherford in experiments conducted
between 1911 and 1919.
The number of protons in an atom defines what element it is. For
example, carbon atoms have six protons, hydrogen atoms have one and

oxygen atoms have eight. The number of protons in an atom is referred to as


the atomic number of that element. The number of protons in an atom also
determines the chemical behavior of the element. The Periodic Table of the
Elements arranges elements in order of increasing atomic number.
Protons are made of other particles called quarks. There are three
quarks in each proton two "up" quarks and one "down" quark and they
are held together by other particles called gluons.

4.7.3

Electrons

Electrons were discovered by J. J. Thomson in 1897. They have a negative charge


and are electrically attracted to the positively charged protons. Electrons
surround the atomic nucleus in pathways called orbitals. The inner orbitals
surrounding the atom are spherical but the outer orbitals are much more
complicated.
An atom's electron configuration is the orbital description of the
locations of the electrons in an unexcited atom. Using the electron
configuration and principles of physics, chemists can predict an atom's
properties, such as stability, boiling point and conductivity.
Typically, only the outermost electron shells matter in chemistry. The
inner electron shell notation is often truncated by replacing the long-hand
orbital description with the symbol for a noble gas in brackets. This method
of notation vastly simplifies the description for large molecules.
For example, the electron configuration for beryllium (Be) is 1s22s2, but
it's is written [He]2s2. [He] is equivalent to all the electron orbitals in a
helium atom. The Letters, s, p, d, and f designate the shape of the orbitals
and the superscript gives the number of electrons in that orbital.

4.7.4

Neutrons

Neutrons are uncharged particles found within atomic nuclei. A


neutron's mass is slightly larger than that of a proton. Like protons, neutrons
are also made of quarks one "up" quark and two "down" quarks. Neutrons
were discovered by James Chadwick in 1932.

4.7.5

Isotopes

The number of neutrons in a nucleus determines the isotope of that


element. For example, hydrogen has three known isotopes: protium,
deuterium and tritium. Protium, symbolized as 1H, is just ordinary hydrogen;
it has one proton and one electron and no neutrons. Deuterium (D or 2H) has
one proton, one electron and one neutron. Tritium (T or 3H) has one proton,
one electron and two neutrons.

5ORDER AMONG ELEMENTS

5.1 DEVELOPMENT OF THE PERIODIC TABLE


Before written history, people were aware
of some of the elements in the periodic table.
Elements such as gold (Au), silver (Ag), copper
(Cu), lead (Pb), tin (Sn), and mercury (Hg).
It wasn't until 1649, however, until the first
element was discovered through scientific
inquiry by Hennig Brand . That element was
phosphorous (P).
By 1869, 63 elements had been discovered.
Between 1817-1829, Johann Dobereiner began to group elements with
similar properties in to groups of three or triads. This began in 1817 when he

noticed that the atomic weights of strontium, Sr, was


halfway between the weights of calcium and barium.
These elements possessed similar chemical properties.
By 1829, he had discovered the a halogen triad made
up of chlorine, bromine, and iodine and a alkali metal
triad of lithium, sodium and potassium. He postulated
that nature contained triads of elements in which the
middle element had properties that were an average
of the other two elements. Later, other scientists
found other triads and recognized that elements could
be grouped into set large than three. The poor
accuracy of measurements such as that of atomic
weights hindered grouping more elements.
In 1862, A.E.Beguyer de Chancourtois was the
first person to make use of atomic weights to reveal that
the elements were arranged according to their atomic
weights with similar elements occurring at regular
intervals. He drew the elements as a continuous spiral
around a cylinder divided into 16 parts. A list of elements
was wrapped around a cylinder so that several sets of
similar elements lined up, creating the first geometric
representation of the periodic law
In 1863, John Newlands, an
English chemist, proposed the Law of Octaves which
stated that elements repeated their chemical
properties every eighth element.
The musical analogy was ridiculed at the time, but was
found to be insightful after the work of Mendeleev and
Meyer were published.

5.2 THE FATHERS OF THE PERIODIC TABLE

Lothar Meyer and Dmitri


Mendeleev independently produced remarkably
versions of the periodic table of elements at the
the same time.
Meyer's 1864 textbook included a
abbreviated version of a periodic table used to
about half of the known elements. In 1868,
constructed an extended table which he
to a colleague for evaluation. This table
unfortunately was not published until 1870,
year after
Mendeleev's table was published.

Ivanovich
similar
essentially

classify
Meyer
gave
a

Mendeleev periodic table appeared


in his work "On the Relationship of
the Properties of the Elements to
their Atomic Weights" in 1869.
Mendeleev placed many elements out of
order based on their accepted atomic
weights at the time.
Mendeleev predicted the existence
and properties of unknown elements which he
called eka aluminum, eka-boron, and ekasilicon. The elements gallium, scandium and
germanium were found later to fit his predictions quite well.

Glenn Seaborg discovered the transuranium


elements, atomic numbers 94 to 102. The completion
the actinide series allows Seaborg to redesign the
periodic table into it current form. Both the lanthanide
and actinide series of elements were placed under the
rest of the periodic table. These elements technically
should be placed between the alkaline earth metals
and the transition metals; however, since this would
make the periodic table too wide, they were placed
below the rest of the elements.

of

Dr. Seaborg and his colleagues are also responsible for the identification of
more than 100 isotopes of elements.

5.3 THE MODERN PERIODIC TABLE

The periodic table is a tabular arrangement of the chemical


elements, organized on the basis of their atomic number (number of protons
in the nucleus), electron configurations, and recurring chemical properties.
Elements are presented in order of increasing atomic number, which is
typically listed with the chemical symbol in each box. The standard form of
the table consists of a grid of elements laid out in 18 columns and 7 rows,
with a double row of elements below that. The table can also be
deconstructed into four rectangular blocks: the s-block to the left, the p-block
to the right, the d-block in the middle, and the f-block below that.
The rows of the table are called periods; the columns are called groups,
with some of these having names such as halogens or noble gases. Since, by
definition, a periodic table incorporates recurring trends, the table can be
used to derive relationships between the properties of the elements and
predict the properties of new, yet to be discovered or synthesized, elements.
As a result, a periodic table provides a useful framework for analyzing
chemical behavior, and so the tables, in various forms, are widely used in
chemistry and other sciences.

Although precursors exist, Dmitri Mendeleev is generally credited with


the publication, in 1869, of the first widely recognized periodic table. He
developed his table to illustrate periodic trends in the properties of the thenknown elements. Mendeleev also predicted some properties of then-unknown
elements that would be expected to fill gaps in this table. Most of his
predictions were proved correct when the elements in question were
subsequently discovered. Mendeleev's periodic table has since been
expanded and refined with the discovery or synthesis of further new
elements and the development of new theoretical models to explain
chemical behavior.
All elements from atomic numbers 1 (hydrogen) to 118 (ununoctium)
have been discovered or reportedly synthesized, with elements 113, 115,
117, and 118 having yet to be confirmed. The first 98 elements exist
naturally although some are found only in trace amounts and were
synthesized in laboratories before being found in nature. [n 1] Elements with
atomic numbers from 99 to 118 have only been synthesized, or claimed to
be so, in laboratories. Production of elements having higher atomic numbers
is being pursued, with the question of how the periodic table may need to be
modified to accommodate any such additions being a matter of ongoing
debate. Numerous synthetic radionuclides of naturally occurring elements
have also been produced in laboratories.

5.4 PERIODIC PROPERTIES OF THE ELEMENTS


The elements in the periodic table are arranged in order of increasing atomic number. All
of these elements display several other trends and we can use the periodic law and table
formation to predict their chemical, physical, and atomic properties. Understanding these trends
is done by analyzing the elements electron configuration; all elements prefer an octet formation
and will gain or lose electrons to form that stable configuration.

5.4.1Atomic Radius
We can never determine the atomic radius of an atom because there is
never a zero probability of finding an electron, and thus never a distinct
boundary to the atom. All that we can measure is the distance between two
nuclei (internuclear distance). A covalent radius is one-half the distance
between the nuclei of two identical atoms. An ionic radius is one-half the
distance between the nuclei of two ions in an ionic bond. The distance must
be apportioned for the smaller cation and larger anion. A metallic radius is
one-half the distance between the nuclei of two adjacent atoms in a
crystalline structure. The noble gases are left out of the trends in atomic radii
because there is great debate over the experimental values of their atomic
radii. The SI units for measuring atomic radii are the nanometer (nm) and the
picometer (pm). 1 nm = 1 X 10-9 m; 1 pm = 1 X 10-12 m.

Figure 1:Covalent Radii

Figure 2: Ionic Radii

Figure 3:

Metallic Radii

To explain this trend, the concept of screening and penetration must be


understood. Penetration is commonly known as the distance that an electron
is from the nucleus. Screening is defined as the concept of the inner
electrons blocking the outer electrons from the nuclear charge. Within this
concept we assume that there is no screening between the outer electrons
and that the inner electrons shield the outer electrons from the total positive
charge of the nucleus. In order to comprehend the extent of screening and
penetration within an atom, scientists came up with the effective nuclear
charge, Zeff. The equation for calculating the effective nuclear charge is
shown below.
Zeff=ZS
In the equation S represents the number of inner electrons that screen
the outer electrons. Students can easily find S by using the atomic number of
the noble gas that is one period above the element. For example, the S we
would use for Chlorine would be 10 (the atomic number of Neon). Z is the
total number of electrons in the atom. Since we know that a neutral atom
has an identical number of protons and electrons, we can use the atomic
number to define Z. For example, Chlorine would have a Z value of 17 (the
atomic number of Chlorine). Continuing to use Chlorine as an example, the

10 inner electrons (S) would screen out the positive charge of ten protons.
Therefore there would be and effective nuclear charge of 17-10 or +7. The
effective nuclear charge shows that the nucleus is pulling the outer electrons
with a +7 charge and therefore the outer electrons are pulled closer to the
nucleus and the atomic radii is smaller. In summary, the greater the nuclear
charge, the greater pull the nucleus has on the outer electrons and the
smaller the atomic radii. In contrast, the smaller nuclear charge, the lesser
pull the nucleus has on the outer electrons, and the larger atomic radii.
Additionally, as the atomic number increases, the effective nuclear charge
also increases. Figure 3 depicts the effect that the effective nuclear charge
has on atomic radii.

Now we are ready to describe the atomic radius trend in the periodic
table. The atomic number increases moving left to right across a period and
subsequently so does the effective nuclear charge. Therefore, moving left to
right across a period the nucleus has a greater pull on the outer electrons
and the atomic radii decreases. Moving down a group in the periodic table,
the number of filled electron shells increases. In a group, the valence
electrons keep the same effective nuclear charge, but now the orbitals are
farther from the nucleus. Therefore, the nucleus has less of a pull on the
outer electrons and the atomic radii are larger.

We can now use these concept to explain the atomic radius differences
of cations and anions. A cation is an atom that has lost one of its outer
electrons. Cations have a smaller radius than the atom that they were
formed from. With the loss of an electron, the positive nuclear charge out
powers the negative charge that the electrons exert. Therefore, the positive
nucleus pulls the electrons tighter and the radius is smaller. An anion is an
atom that has gained an outer electron. Anions have a greater radius than
the atom that they were formed from. The gain of an electron does not alter
the nuclear charge, but the addition of an electron causes a decrease in the
effective nuclear charge. Therefore, the electrons are held more loosely and
the atomic radius is increased.

5.4.2Ionization Energy (ionization potential)

Expelling an electron from an atom requires enough energy to


overcome the magnetic pull of the positive charge of the nucleus. Therefore,
ionization energy (I.E. or I) is the energy required to completely remove an
electron from a gaseous atom or ion. The Ionization Energy is always
positive.

The energy required to remove one valence electron is the first


ionization energy, the second ionization energy is the energy required to
remove a second valence electron, and so on.
1st I.E.: Na(g) Na+(g)+ e2nd I.E.: Na(g) Na2+(g) + 2eIonization energies increase relative to high effective charge. The
highest ionization energies are the noble gases because they all have high
effective charge due to their octet formation and require a high amount of
energy to destroy that stable configuration. The highest amount of energy
required occurs with the elements in the upper right hand corner.
Additionally, elements in the left corner have a low ionization energy
because losing an electron allows them to have the noble gas configuration.
Therefore, it requires less energy to remove one of their valence electrons
Ionization Energies of certain elements (1st IE, 2nd IE, etc)
Elem
ent
Na
Mg
Al
Si
P
S
Cl
Ar

1st
496
738
577
786
1060
999.6
1256
1520

2nd
4562
1451
1817
1577
1903
2251
2297
2666

3rd

4th

5th

6th

7th

7733
2745
3232
2912
3361
3822
3931

11580
4356
4957
4564
5158
5771

16090
6274
7013
6542
7238

21270
8496
9362
8781

27110
11020
12000

These are the ionization energies for the period three elements. Notice
how Na after in the second I.E, Mg in the third I.E., Al in the fourth I.E., and so
on, all have a huge increase in energy compared to the proceeding one. This

occurs because the proceeding configuration was in a stable octet formation;


therefore it requires a much larger amount of energy to ionize.
Ionization Energies increase going left to right across a period and
increase going up a group. As you go up a group, the ionization energy
increases, because there are less electron shielding the outer electrons from
the pull of the nucleus. Therefore, it requires more energy to out power the
nucleus and remove an electron. As we move across the periodic table from
left to right, the ionization energy increases, due to the effective nuclear
charge increasing. This is because the larger the effective nuclear charge,
the stronger the nucleus is holding onto the electron and the more energy it
takes to release an electron.

The ionization energy is only a general rule. There are some instances
when this trend does not prove to be correct. These can typically be
explained by their electron configuration. For example, Magnesium has
higher ionization energy than Aluminum. Magnesium has an electron
configuration of [Ne]3s2. Magnesium has a high ionization energy because it
has a filled 3s orbital and it requires a higher amount of energy to take an
electron from the filled orbital.

5.4.3Electron Affinity
Electron affinity (E.A.) is the energy change that occurs when an
electron is added to a gaseous atom. Electron affinity can further be defined
as the enthalpy change that results from the addition of an electron to a
gaseous atom. It can be either positive or negative value. The greater the
negative value, the more stable the anion is.
X(g) + e- ---> X- + Energy (Exothermic) The electron affinity is positive
X(g) + e- + Energy ---> X- (Endothermic) The electron affinity is
negative

It is more difficult to come up with trends that describe the electron


affinity. Generally, the elements on the right side of the periodic table will
have large negative electron affinity. The electron affinities will become less
negative as you go from the top to the bottom of the periodic table.
However, Nitrogen, Oxygen, and Fluorine do not follow this trend. The noble
gas electron configuration will be close to zero because they will not easily
gain electrons.

5.4.4Electronegativity

Electronegativity is the measurement of an atom to compete for


electrons in a bond. The higher the electronegativity, the greater its ability to
gain electrons in a bond. Electronegativity will be important when we later
determine polar and nonpolar molecules. Electronegativity is related with
ionization energy and electron affinity. Electrons with low ionization energies
have low electronegativities because their nuclei do not exert a strong
attractive force on electrons. Elements with high ionization energies have
high electronegativities due to the strong pull exerted by the positive nucleus
on the negative electrons. Therefore the electronegativity increases from
bottom to top and from left to right.

5.4.5Metallic Character
The metallic character is used to define the chemical properties that metallic elements
present. Generally, metals tend to lose electrons to form cations. Nonmetals tend to gain
electrons to form anions. They also have a high oxidation potential therefore they are easily
oxidized and are strong reducing agents. Metals also form basic oxides; the more basic the oxide,
the higher the metallic character.

As you move across the table from left to right, the metallic character decreases, because
the elements easily accept electrons to fill their valance shells. Therefore, these elements take on
the nonmetallic character of forming anions. As you move up the table, the metallic character
decreases, due to the greater pull that the nucleus has on the outer electrons. This greater pull
makes it harder for the atoms to lose electrons and form cations.

5.4.6

Other
Trends

Melting
Points:
According
to General
Chemistry
by Petrucci,
trends
in
melting
points and
molecular
mass
of
binary
carbonhalogen compounds and hydrogen halides are due to intermolecular forces. Melting destroys the

arrangement of atoms in a solid; therefore the amount of heat necessary for melting to occur
depends on the strength of attraction between the atoms. This strength of attraction increases as
the number of electrons increase. Increase in electrons increases bonding.
Example: Melting point of HF should be approximately -145 C based off melting points
of HCl, HBr, and HI, but the observed value is -83.6 (Petrucci)
Heat and electricity conductibility vary regularly across a period. Melting points may
increase gradually or reach a peak within a group then reverse direction.

5.4.7Summary Of Periodic Trends

The Periodic Table of Elements categorizes like elements together. Dmitri Mendeleev, a
Russian scientist, was the first to create a widely accepted arrangement of the elements in 1869.
Mendeleev believed that when the elements are arranged in order of increasing atomic mass,
certain sets of properties recur periodically. Although most modern periodic tables are arranged
in eighteen groups (columns) of elements, Mendeleev's original periodic table had the elements
organized into eight groups and twelve periods (rows).
On the periodic table, elements that have similar properties are in the same groups
(vertical). From left to right, the atomic number (z) of the elements increases from one period to
the next (horizontal). The groups are numbered at the top of each column and the periods on the
left next to each row. The main group elements are groups 1,2 and 13 through 18. These groups
contain the most naturally abundant elements, and are the most important for life. The elements
shaded in light pink in the table above are known as transition metals. The two rows of elements
starting at z=58, are sometimes called inner transition metals and have that have been extracted
and placed at the bottom of the table, because they would make the table too wide if kept
continuous. The 14 elements following lanthanum (z=57) are called lanthanides, and the 14
following actinium (z=89) are called actinides.

Elements in the periodic table can be placed into two broad categories, metals and
nonmetals. Most metals are good conductors of heat and electricity, are malleable and ductile,
and are moderate to high melting points. In general, nonmetals are nonconductors of heat and
electricity, are nonmalleable solids, and many are gases at room temperature. Just as shown in
the table above, metals and nonmetals on the periodic table are often separated by a stairstep
diagonal line, and several elements near this line are often called metalloids (Si, Ge, As, Sb, Te,
and At). Metalloids are elements that look like metals and in some ways behave like metals but
also have some nonmetallic properties. The group to the farthest right of the table, shaded
orange, is known as the noble gases. Noble gases are treated as a special group of nonmetals.

5.5 ATOMIC ORBITAL

The shapes of the first five atomic orbitals: 1s, 2s, 2p x, 2py, and 2pz.
The colors show the wave function phase. These are graphs of (x,y,z)
functions which depend on the coordinates of one electron. To see the
elongated shape of (x,y,z)2 functions that show probability density more
directly, see the graphs of d-orbitals below.
An atomic orbital is a mathematical function that describes the wavelike behavior of either one electron or a pair of electrons in an atom.[1] This
function can be used to calculate the probability of finding any electron of an
atom in any specific region around the atom's nucleus. The term may also
refer to the physical region or space where the electron can be calculated to
be present, as defined by the particular mathematical form of the orbital.[2]
Each orbital in an atom is characterized by a unique set of values of
the three quantum numbers n, , and m, which respectively correspond to
the electron's energy, angular momentum, and an angular momentum
vector component (the magnetic quantum number). Any orbital can be
occupied by a maximum of two electrons, each with its own spin quantum
number. The simple names s orbital, p orbital, d orbital and f orbital
refer to orbitals with angular momentum quantum number = 0, 1, 2 and 3
respectively. These names, together with the value of n, are used to describe
the electron configurations of atoms. They are derived from the description
by early spectroscopists of certain series of alkali metal spectroscopic lines
as sharp, principal, diffuse, and fundamental. Orbitals for > 3 continue
alphabetically, omitting j (g, h, i, k, ...).
Atomic orbitals are the basic building blocks of the atomic orbital
model (alternatively known as the electron cloud or wave mechanics

model), a modern framework for visualizing the submicroscopic behavior of


electrons in matter. In this model the electron cloud of a multi-electron atom
may be seen as being built up (in approximation) in an electron configuration
that is a product of simpler hydrogen-like atomic orbitals. The repeating
periodicity of the blocks of 2, 6, 10, and 14 elements within sections of the
periodic table arises naturally from the total number of electrons that occupy
a complete set of s, p, d and f atomic orbitals, respectively.

5.6 ALKALI METALS/ALKALI EARTH METALS


The Alkali metals are comprised of group 1A of the periodic table and
consist of Lithium, Sodium, Rubidium, Caesium, and Francium. These metals
are highly reactive and form ionic compounds (when a nonmetal and a metal
come together) as well as many other compounds. Alkali metals all have a
charge of +1 and have the largest atom sizes than any of the other elements
on each of their respective periods.
Alkali Earth Metals are located in group 2A and consist of Bereyllium,
Magnesium, Calcium, Strontium, Barium, and Radium. Unlike the Alkali
metals, the earth metals have a smaller atom size and are not as reactive.
These metals may also form ionic and other compounds and have a charge
of +2.

5.7 TRANSITION METALS


The transition metals range from groups IIIB to XIIB on the periodic
table. These metals form positively charged ions, are very hard, and have
very high melting and boiling points. Transition metals are also good
conductors of electricity and are malleable.

5.8 LANTHANIDES AND ACTINIDES


Lanthanides and Actinides, form the block of two rows that are placed
at the bottom of the periodic table for space issues. These are also
considered to be transition metals. Lanthanides are form the top row of this
block and are very soft metals with high boiling and melting points. Actinides
form the bottom row and are radioactive. They also form compounds with
most nonmetals. To find out why these elements have their own section,
check out the electron configurations page.

5.9 METALLOIDS
As mentioned in the introduction, metalloids are located along the
staircase separating the metals from the nonmetals on the periodic table.
Boron, silicon, germanium, arsenic, antimony, and tellurium all have metal
and nonmetal properties. For example, Silicon has a metallic luster but is
brittle and is an inefficient conductor of electricity like a nonmetal. As the

metalloids have a combination of both metallic and nonmetal characteristics,


they are intermediate conductors of electricity or "semiconductors".

5.10 HALOGENS
Halogens are comprised of the five nonmetal elements Flourine,
Chlorine, Bromine, Iodine, and Astatine. They are located on group 17 of the
periodic table and have a charge of -1. The term "halogen" means "saltformer" and compounds that contain one of the halogens are salts. The
physical properties of halogens vary significantly as they can exist as solids,
liquids, and gases at room temperature. However in general, halogens are
very reactive, especially with the alkali metals and earth metals of groups 1
and 2 with which they form ionic compounds.

5.11 NOBLE GASES


The noble gases consist of group 18 (sometimes reffered to as group O) of
the periodic table of elements. The noble gases have very low boiling and
melting points and are all gases at room temperature. They are also very
nonreactive as they already have a full valence shell with 8 electrons.
Therefore, the noble gases have little tendency to lose or gain electrons.

5.12 USEFUL RELATIONSHIPS FROM THE PERIODIC TABLE


The periodic table of elements is useful in determining the charges on
simple monoatomic ions. For main-group elements, those categorized in
groups 1, 2, and 13-18, form ions they lose the same number of electrons as
the corresponding group number to which they fall under. For example, K
atoms (group 1) lose one electron to become K + and Mg atoms (group 2) lose
two electrons to form Mg2+.
The elements in groups 3-12 are called transition elements, or
transition metals. Similar to the main-group elements described above, the
transition metals form positive ions but due to their capability of forming
more than two or more ions of differing charge, a relation between the group
number and the charge is non-existent.

5.13 ELECTRONIC CONFIGURATIONS


The electron configuration of an atom is the representation of the
arrangement of electrons distributed among the orbital shells and
subshells. Commonly, the electron configuration is used to describe the
orbitals of an atom in its ground state, but it can also be used to represent an
atom that has ionized into a cation or anion by compensating with the loss of
or gain of electrons in their subsequent orbitals. Many of the physical and

chemical properties of elements can be correlated to their unique electron


configurations. The valence electrons, electrons in the outermost shell, are
the determining factor for the unique chemistry of the element.
Before assigning the electrons of an atom into orbitals, one must
become familiar with the basic concepts of electron configurations. Every
element on the periodic table consists of atoms, which are composed of
protons, neutrons, and electrons. Electrons exhibit a negative charge and are
found around the nucleus of the atom in electronic orbitals, defined as the
volume of space in which the electron can be found within 95% probability.
The four different types of orbitals s,p,d, and f have different shapes, and one
orbital can hold a maximum of two electrons. The p, d, and f orbitals have
different sublevels thus can hold more electrons.
As stated, the electron configuration of each element is unique to its
position on the periodic table. The energy level is determined by the period
and number of electrons is given by the atomic number of the element.
Orbitals on different energy levels are similar to each other, but they occupy
different areas in space. The 1s orbital and 2s orbital both have the
characteristics of an s orbital (radial nodes, spherical volume probabilities,
can only hold two electrons, etc.) but as they are found in different energy
levels they occupy different spaces around the nucleus. Each orbital can be
represented by specific blocks on the periodic table. The s-block is the region
of the alkali metals including helium (Groups 1 & 2), the d-block is the
transition metals (Groups 3 to 12), the p-block are the main group elements
from Groups 13 to 18, and the f-block are the lanthanides and actinides
series.

Using the periodic table to determine the electron configurations of


atoms is key, but also keep in mind that there are certain rules to follow
when assigning electrons to different orbitals. The periodic table is an
incredibly helpful tool in writing electron configurations.

5.14 RULES FOR ASSIGNING ELECTRON ORBITALS


Electrons fill orbitals in a way to minimize the energy of the atom. The
electrons in an atom therefore fill the principal energy levels in order of
increasing energy (the electrons are getting farther from the nucleus). The
order of levels filled looks like this:
1s, 2s, 2p, 3s, 3p, 4s, 3d, 4p, 5s, 4d, 5p, 6s, 4f, 5d, 6p, 7s,
5f, 6d, and 7p
One way to remember this pattern, probably the
easiest, is to refer to the periodic table and remember
where each orbital block falls to logically deduce this
pattern. Another way is to make a table like the one below
and use vertical lines to determine which sub shells
correspond with each other.

5.15 PAULI EXCLUSION PRINCIPLE


The Pauli exclusion principle states that no two electrons can have the
same four quantum numbers. The first three (n, l, and ml) may be the same,
but the fourth quantum number must be different. A single orbital can hold a
maximum of two electrons, which must have opposing spins; otherwise they
would have the same four quantum numbers, which is forbidden. One
electron is spin up (ms = +1/2) and the other would spin down (m s = -1/2).
This tells us that each subshell has double the electrons per orbital. The s
subshell has 1 orbital that can hold up to 2 electrons, the p subshell has 3
orbitals that can hold up to 6 electrons, the d subshell has 5 orbitals that
hold up to 10 electrons, and the f subshell has 7 orbitals with 14 electrons.
Example 1
The first three quantum numbers of an electron are n=1, l=0,
ml=0. Only two electrons can correspond to these, which would be
either ms = -1/2 or ms = +1/2. As we already know from our studies
of quantum numbers and electron orbitals, we can conclude that
these four quantum numbers refer to 1s subshell. If only one of
the ms values are given then we would have 1s 1 (denoting
hydrogen) if both are given we would have 1s 2 (denoting helium).
Visually, this is be represented as:

As shown, the 1s subshell can hold only two electrons and when
filled the electrons have opposite spins.

5.16 HUND'S RULE


When assigning electrons in orbitals, each electron will first fill all the
orbitals with similar energy (also referred to as degenerate) before pairing
with another electron in a half-filled orbital. Atoms at ground states tend to
have as many unpaired electrons as possible. When visualizing this
processes, think about how electrons are exhibiting the same behavior as the
same poles on a magnet would if they came into contact; as the negatively
charged electrons fill orbitals they first try to get as far as possible from each
other before having to pair up.

Example
If we look at the correct electron configuration of Nitrogen (Z = 7),
a very important element in the biology of plants: 1s2 2s2 2p3

We can clearly see that p orbitals are half filled as there are three
electrons and three p orbitals. This is because Hund's Rule states
that the three electrons in the 2p subshell will fill all the empty
orbitals first before filling orbitals with electrons in them. If we

look at the element after Nitrogen in the same period, Oxygen (Z


= 8) its electron configuration is: 1s2 2s2 2p4

Oxygen has one more electron than Nitrogen and as the orbitals
are all half filled the electron must pair up.

5.17 THE AUFBAU PROCESS


Aufbau comes from the German word "aufbauen," meaning "to
build." When writing electron configurations, orbitals are built up from atom
to atom. When writing the electron configuration for an atom, orbitals are
filled in order of increasing atomic number. However, there are some
exceptions to this rule.
Example
Following the pattern across a period from B (Z=5) to Ne (Z=10), the
number of electrons increases and the subshells are filled. This example
focuses on the p subshell, which fills from boron to neon.

B (Z=5) configuration: 1s2 2s2 2p1

C (Z=6) configuration:1s2 2s2 2p2

N (Z=7) configuration:1s2 2s2 2p3

O (Z=8) configuration:1s2 2s2 2p4

F (Z=9) configuration:1s2 2s2 2p5

Ne (Z=10) configuration:1s2 2s2 2p6

Exceptions
Although the Aufbau rule accurately predicts the electron configuration
of most elements, there are notable exceptions among the transition metals
and heavier elements. The reason these exceptions occur is that some

elements are more stable with fewer electrons in some subshells and more
electrons in others. A list of the exceptions to the Aufbau process is given
below.
Table 1: Exceptions to Electron Configuration Trends
Period 4:
Chromium: Z:24 [Ar] 3d54s1

Period 5:
Niobium: Z:41 [Kr] 5s1 4d4

Copper: Z:29 [Ar] 3d104s1

Molybdenum: Z:42 [Kr] 5s1 4d5


Ruthenium: Z:44 [Kr] 5s1 4d7
Rhodium: Z:45 [Kr] 5s1 4d8
Palladium: Z:46 [Kr] 4d10
Silver: Z:47 [Kr] 5s1 4d10

Period 6:

Period 7:

Lanthanum: Z:57 [Xe] 6s2 5d1

Actinium: Z:89 [Rn] 7s2 6d1

Cerium: Z:58 [Xe] 6s2 4f1 5d1

Thorium: Z:90 [Rn] 7s2 6d2

Gadolinium: Z:64 [Xe] 6s2 4f7


5d1

Protactium: Z:91 [Rn] 7s2 5f2 6d1

Platinum: Z:78 [Xe] 6s1 4f14 5d9 Uranium: Z:92 [Rn] 7s2 5f3 6d1
Gold: Z:79 [Xe] 6s1 4f14 5d10

Neptunium: Z:93 [Rn] 7s2 5f4 6d1


Curium: Z:96 [Rn] 7s2 5f7 6d1
Lawrencium: Z:103 [Rn] 7s2 5f14
7p1

5.18 WRITING ELECTRON CONFIGURATIONS


When writing an electron configuration, first write the energy level (the
period), then the subshell to be filled and the superscript, which is the
number of electrons in that subshell. The total number of electrons is the

atomic number, Z. The rules above allow one to write the electron
configurations for all the elements in the periodic table.
Three methods are used to write electron configurations:
1. orbital diagrams
2. spdf notation
3. noble gas notation
Each method has its own purpose and each has its own drawbacks.

5.19 ORBITAL DIAGRAMS


An orbital diagram, like those shown above, is a visual way to
reconstruct the electron configuration by showing each of the separate
orbitals and the spins on the electrons. This is done by first determining the
subshell (s,p,d, or f) then drawing in each electron according to the stated
rules above.
Example: Aluminum
Write the electron configuration for aluminum.
SOLUTION
Aluminum is in the 3rd period and it has an atomic number of Z=13. If we
look at the periodic table we can see that its in the p-block as it is in group
13. Now we shall look at the orbitals it will fill: 1s, 2s, 2p, 3s, 3p. We know
that aluminum completely fills the 1s, 2s, 2p, and 3s orbitals because
mathematically this would be 2+2+6+2=12. The last electron is in the 3p
orbital. Also another way of thinking about it is that as you move from each
orbital block, the subshells become filled as you complete each section of
the orbital in the period. The block that the atom is in (in the case for
aluminum: 3p) is where we will count to get the number of electrons in the
last subshell (for aluminum this would be one electron because its the first
element in the period 3 p-block). This gives the following:

Note that in the orbital diagram, the two opposing spins of the electron can
be visualized. This is why it is sometimes useful to think about electron
configuration in terms of the diagram. However, because it is the most
time consuming method, it is more common to write or see electron
configurations in spdf notation and noble gas notation. Another example is
the electron configuration of iridium:

The electron configuration of iridium is much longer than aluminum.


Although drawing out each orbital may prove to be helpful in determining
unpaired electrons, it is very time consuming and often not as practical as
the spdf notation, especially for atoms with much longer configurations.
Hund's rule is also followed, as each electron fills up each 5d orbital before
being forced to pair with another electron.

5.20 ELECTRON NOTATION USING SPDF

The most common way to describe electron configurations is to write


distributions in the spdf notation. Although the distributions of electrons in
each orbital are not as apparent as in the diagram, the total number of
electrons in each energy level is described by a superscript that follows the
relating energy level. To write the electron configuration of an atom, identify
the energy level of interest and write the number of electrons in the energy
level as its superscript as follows: 1s 2. This is the electron configuration of
helium; it denotes a full s orbital. The periodic table is used as a reference to
accurately write the electron configurations of all atoms.

Example: Yttrium
Start with the straightforward problem of finding the electron configuration
of the element yttrium. As always, refer to the periodic table. The element
yttrium (symbolized Y) is a transition metal, found in the fifth period and in
Group 3. In total it has thirty-nine electrons. Its electron configuration is as
follows:
1s2 2s2 2p6 3s2 3p6 4s2 3d10 4p6 5s2 4d1
This is a much simpler and more efficient way to portray electron
configuration of an atom. A logical way of thinking about it is that all that is
required is to fill orbitals across a period and through orbital blocks. The
number of elements in each block is the same as in the energy level it
corresponds. For example, there are 2 elements in the s-block, and 10
elements in the d-block. Moving across, simply count how many elements
fall in each block. Yttrium is the first element in the fourth period dblock; thus there is one electron in that energy level. To check the answer,
verify that the subscripts add up to the atomic number. In this case,
2+2+6+2+6+2+10+6+2+1= 39 and Z=39, so the answer is correct.
A slightly more complicated example is the electron configuration of
bismuth (symbolized Bi, with Z = 83). The periodic table gives the following
electron configuration:
1s2 2s2 2p6 3s2 3p6 4s2 3d10 4p65s2 4d10 5p6 6s2 4f14 5d10 6p3
The reason why this electron configuration seems more complex is that
the f-block, the Lanthanide series, is involved. Most students who first learn
electron configurations often have trouble with configurations that must pass
through the f-block because they often overlook this break in the table and
skip that energy level. Its important to remember that when passing the 5d
and 6d energy levels that one must pass through the f-block lanthanoid and
actinoid series. Keeping this in mind, this "complex" problem is greatly
simplified.
Another method (but less commonly used) of writing the spdf notation
is the expanded notation format. This is the same concept as before, except
that each individual orbital is represented with a subscript. The p, d, and f
orbitals have different sublevels. The p orbitals are px, py, and pz, and if
represented on the 2p energy with full orbitals would look like: 2p x2 2py2 2pz2.
The expanded notation for neon (Ne, Z=10) is written as follows:
1s2 2s2 2px2 2py2 2pz2
The individual orbitals are represented, but the spins on the electrons
are not; opposite spins are assumed. When representing the configuration of

an atom with half filled orbitals, indicate the two half filled orbitals. The
expanded notation for carbon is written as follows:
1s2 2s2 2px1 2py1
Because this form of the spdf notation is not typically used, it is not as
important to dwell on this detail as it is to understand how to use the general
spdf notation.

5.21 NOBLE GAS NOTATION

This brings up an interesting point about elements and electron


configurations. As the p subshell is filled in the above example about the
Aufbau principle (the trend from boron to neon), it reaches the group
commonly known as the noble gases. The noble gases have the most stable
electron configurations, and are known for being relatively inert. All noble
gases have their subshells filled and can be used them as a shorthand way of
writing electron configurations for subsequent atoms. This method of writing
configurations is called the noble gas notation, in which the noble gas in the
period above the element that is being analyzed is used to denote the
subshells that element has filled and after which the valence electrons
(electrons filling orbitals in the outer most shells) are written. This looks
slightly different from spdf notation, as the reference noble gas must be
indicated.

Example: Vanadium
What is the electronic configuration of vanadium (V, Z=23)?
SOLUTION
Vanadium is the transition metal in the fourth period and the fifth group.
The noble gas preceding it is argon (Ar, Z=18), and knowing that vanadium
has filled those orbitals before it, argon is used as the reference noble gas.
The noble gas in the configuration is denoted E, in brackets: [E]. To find the
valance electrons that follow, subtract the atomic numbers: 23 - 18 = 5.
Instead of 23 electrons to distribute in orbitals, there are 5. Now there
is enough information to write the electron configuration:
Vanadium, V: [Ar] 4s2 3d3

This method streamlines the process of distributing electrons by showing


the valence electrons, which determine the chemical properties of atoms.
In addition, when determining the number of unpaired electrons in an
atom, this method allows quick visualization of the configurations of the
valance electrons. In the example above, there are a full s orbital and three
half filled d orbitals.

6CHEMICAL BONDS
6.1 ENERGY AND BONDING
Lets start by imagining that there are two hydrogen atoms
approaching one another. As they move closer together, there are three
forces that act on the atoms at the same time. These forces are shown in
figure and are described below:

Forces acting on two approaching atoms: (1) repulsion between electrons,


(2) attraction between protons and electrons and (3) repulsion between
protons.

1. repulsive force between the electrons of the atoms, since like charges
repel
2. attractive force between the nucleus of one atom and the electrons of
another
3. repulsive force between the two positively-charged nuclei

In the example of the two hydrogen atoms, where the resultant force
between them is attraction, the energy of the system is zero when the atoms
are far apart (point A), because there is no interaction between the atoms.
When the atoms are closer together, attractive forces dominate and the
atoms are pulled towards each other. As this happens, the potential energy
of the system decreases because energy would now need to be supplied to
the system in order to move the atoms apart. However, as the atoms move
closer together (i.e. left along the horizontal axis of the graph), repulsive
forces start to dominate and this causes the potential energy of the system
to rise again. At some point, the attractive and repulsive effects are
balanced, and the energy of the system is at its minimum (point X). It is at
this point, when the energy is at a minimum, that bonding takes place.
The distance marked P is the bond length, i.e. the distance between the
nuclei of the atoms when they bond. Q represents the bond energy i.e. the
amount of energy that must be added to the system to break the bonds that
have formed. Bond strength means how strongly one atom attracts and is
held to another. The strength of a bond is related to the bond length, the size
of the bonded atoms and the number of bonds between the atoms. In
general, the shorter the bond length, the stronger the bond between the
atoms, and the smaller the atoms involved, the stronger the bond. The
greater the number of bonds between atoms, the greater will be the bond
strength.

6.2 WHAT HAPPENS WHEN ATOMS BOND?

A chemical bond is formed when atoms are held together by attractive


forces. This attraction occurs when electrons are shared between atoms, or
when electrons are exchanged between the atoms that are involved in the
bond. The sharing or exchange of electrons takes place so that the outer
energy levels of the atoms involved are filled and the atoms are more stable.
If an electron is shared, it means that it will spend its time moving in the
electron orbitals around both atoms.
If an electron is exchanged it means that it is transferred from one
atom to another, in other words one atom gains an electron while the other
loses an electron.
A chemical bond is the physical process that causes atoms and
molecules to be attracted to each other, and held together in more stable
chemical compounds.
The type of bond that is formed depends on the elements that are
involved. In this section, we will be looking at three types of chemical
bonding: covalent, ionic and metallic bonding.
You need to remember that it is the valence electrons that are involved in
bonding and those atoms will try to fill their outer energy levels so that they
are more stable.

6.3 TYPES OF CHEMICAL BONDS


6.3.1

Ionic Bonds

Ionic bonding is the complete transfer of valence electron(s) between atoms


and is a type of chemical bond that generates two oppositely charged ions. It
is observed because metals with few electrons in its outer-most orbital. By
losing those electrons, these metals can achieve noble-gas configuration and
satisfy the octet rule. Similarly, nonmetals that have close to 8 electrons in
its valence shell tend to readily accept electrons to achieve its noble gas
configuration.
In ionic bonding, electrons are transferred from one atom to another
resulting in the formation of positive and negative ions. The electrostatic
attractions between the positive and negative ions hold the compound
together. The predicted overall energy of the ionic bonding process, which
includes the ionization energy of the metal and electron affinity of the
nonmetal, is usually positive, indicating that the reaction is endothermic and
unfavorable. However, this reaction is highly favorable because of their

electrostatic attraction. At the most ideal inter-atomic distance, attraction


between these particles releases enough energy to facilitate the reaction.
Most ionic compounds tend to dissociate in polar solvents because they
are often polar. This phenomenon is due to the opposite charges on each
ions.
At a simple level, a lot of importance is attached to the electronic
structures of noble gases like neon or argon which have eight electrons in
their outer energy levels (or two in the case of helium). These noble gas
structures are thought of as being in some way a "desirable" thing for an
atom to have. One may well have been left with the strong impression that
when other atoms react, they try to organize things such that their outer
levels are either completely full or completely empty.
In chemical bonds, atoms can either
share their valence electrons. In the
case where one or more atoms lose
and other atoms gain them in order to produce
noble gas electron configuration, the bond is called an
ionic bond.
Typical of ionic bonds are those
alkali halides such as sodium chloride, NaCl.

transfer or
extreme
electrons
a

in the

6.4 METALLIC BONDS


The properties of metals suggest that their atoms possess strong
bonds, yet the ease of conduction of heat and electricity suggest that
electrons can move freely in all directions in a metal. The general
observations give rise to a picture of "positive ions in a sea of electrons" to
describe metallic bonding.

6.5 HYDROGEN BONDING


Hydrogen bonding differs from other uses of the word "bond" since it is
a force of attraction between a hydrogen atom in one molecule and a small
atom of high electronegativity in another molecule. That is, it is an
intermolecular force, not an intramolecular force as in the common use of
the word bond.
When hydrogen atoms are joined in a polar covalent bondwith a small
atom of high electronegativity such as O, F or N, the partial positive charge
on the hydrogen is highly concentrated because of its small size. If the
hydrogen is close to another oxygen, fluorine or nitrogen in another
molecule, then there is a force of attraction termed a dipole-dipole

interaction. This attraction or "hydrogen bond" can have about 5% to 10% of


the strength of a covalent bond.
Hydrogen bonding has a very important effect on the properties of water and ice. Hydrogen
bonding is also very important in proteins and nucleic acids and therefore in life processes. The
"unzipping" of DNA is a breaking of hydrogen bonds which help hold the two strands of the
double helix together.

6.6 LEWIS STRUCTURES


Lewis Structures are visual representations of the bonds between
atoms and illustrate the lone pairs of electrons in molecules. They can also
be called Lewis dot diagrams and are used as a simple way to show the
configuration of atoms within a molecule.

6.6.1

LEWIS THEORY

Between 1916 and 1919, Gilbert Newton Lewis, Walther Kossel, and Irving
Langmuir came up with a theory to explain chemical bonding. This theory
would be later called Lewis Theory and it is based on the following
principles:
1. Valence electrons, or the electrons in the outermost electron shell,
have an essential role in chemical bonding.
2. Ionic bonds are formed between atoms when electrons are
transferred from one atom to another. Ionic bond is a bond between
nonmetals and metals.
3. Covalent bonds are formed between atoms when pairs of electrons
are shared between atoms. A covalent bond is between two nonmetals.
4. Electrons are transferred/shared so that each atom may reach a more
stable electron configuration i.e. the noble gas configuration which
contains 8 valence electrons. This is called octet rule.

6.6.2

LEWIS SYMBOLS AND LEWIS STRUCTURES

A Lewis Symbol for an element is composed of a chemical symbol


surrounded by dots that are used to represent valence electrons. An example
of a Lewis symbol is shown below with the element Carbon, which has the
electron configuration of 1s22s22p2:

This Lewis symbol shows that carbon has four valence electrons in its
outer orbital and these four electrons play a major role in bonding of carbon
molecules.
Lewis symbols differ slightly for ions. When forming a Lewis symbol for
an ion, the chemical symbol is surrounded by dots that are used to represent
valence electrons, and the whole structure is placed in square brackets with
superscript representing the charge of the ion. An example of a Lewis symbol
for the cation and anion of Carbon is shown below:
Cation of Carbon

Anion of Carbon

6.7 NAMING OF COMPOUNDS


Ionic compounds consist of cations (positive ions) and anions (negative
ions). The nomenclature, or naming, of ionic compounds is based on the
names of the component ions. Here are the principal naming conventions for
ionic compounds , along with examples to show how they are used:

Roman Numerals
A Roman numeral in parentheses, followed by the name of the
element, is used for elements that can form more than one positive
ion. This is usually seen with metals. You can use a chart to see the
possible valences for the elements.
Fe2+ Iron (II)
Fe3+ Iron (III)
Cu+ Copper (I)
Cu2+ Copper (II)

-ous and -ic


Although Roman numerals are used to denote the ionic charge of
cations, it is still common to see and use the endings -ous or -ic.
These endings are added to the Latin name of the element (e.g.,
stannous/stannic for tin) to represent the ions with lesser or greater
charge, respectively. The Roman numeral naming convention has wider
appeal because many ions have more than two valences.
Fe2+ Ferrous
Fe3+ Ferric
Cu+ Cuprous
Cu2+ Cupric

-ide
The -ide ending is added to the name of a monoatomic ion of an
element.
H- Hydride
F- Fluoride
O2- Oxide
S2- Sulfide
N3- Nitride
P3- Phosphide

-ite and -ate


Some polyatomic anions contain oxygen. These anions are called
oxyanions. When an element forms two oxyanions, the one with less
oxygen is given a name ending in -ite and the one with more oxgyen is
given a name that ends in -ate.
NO2- Nitrite
NO3- Nitrate
SO32- Sulfite
SO42- Sulfate

hypo- and perIn the case where there is a series of four oxyanions, the hypo- and
per- prefixes are used in conjunction with the -ite and -ate suffixes.
The hypo- and per- prefixes indicate less oxygen and more oxygen,
respectively.
ClO- Hypochlorite
ClO2- Chlorite
ClO3- Chlorate
ClO4- Perchlorate

bi- and di- hydrogen


Polyatomic anions sometimes gain one or more H+ ions to form anions
of a lower charge. These ions are named by adding the word
hydrogen or dihydrogen in front of the name of the anion. It is still
common to see and use the older naming convention in which the
prefix bi- is used to indicate the addition of a single hydrogen ion.
HCO3- Hydrogen carbonate or bicarbonate
HSO4- Hydrogen sulfate or bisulfate
H2PO4- Dihydrogen phosphate

6.8 INTERMOLECULAR FORCES

Forces binding atoms in a molecule are due to chemical bonding. The


energy required to break a bond is called the bond-energy. For example the
average bond-energy for O-H bonds in water is 463 kJ/mol. On average, 463
kJ is required to break 6.023x1023 O-H bonds, or 926 kJ to convert 1.0 mole of
water into 1.0 mol of O and 2.0 mol of H atoms. A space filling model of
water molecule is shown here.
The forces holding molecules together are generally called intermolecular forces. The
energy required to break molecules apart is much smaller than a typical bond-energy, but
intermolecular forces play important roles in determining the properties of a substances.
Intermolecular forces are particularly important in terms how molecules interact and form
biological organisms or even life. This link gives an excellent introduction to the interactions
between molecules.

6.9 VAN DER WAALS INTERACTIONS


Van der Waals forces are driven by induced electrical interactions
between two or more atoms or molecules that are very close to each other.
Van der Waals interaction is the weakest of all intermolecular attractions
between molecules. However, with a lot of Van der Waals forces interacting
between two objects, the interaction can be very strong.

6.10 DIPOLE-DIPOLE INTERACTION


Dipole-Dipole interactions occur between molecules that have
permanent dipoles; these molecules are also referred to as polar molecules.
The figure below shows the electrostatic interaction between two dipoles.

6.11 INDUCED DIPOLES


An induced dipole moment is a temporary condition during which a
neutral nonpolar atom (i.e. Helium) undergo a separation of charges due to
the environment. When an instantaneous dipole atom approaches a
neighboring atom, it can cause that atom to also produce dipoles. The
neighboring atom is then considered to have an induced dipole moment.

6.12 INDUCED DIPOLE-INDUCED DIPOLE INTERACTION


Induced dipole-induced dipole interactions are also known as
dispersion or London forces (name after the German physicist Fritz
London). They are large networks of intermolecular forces between nonpolar
and non-charged molecules and atoms (i.e. alkanes, noble gases, and
halogens). Molecules that have induced dipoles may also induce neighboring
molecules to have dipole moments, so a large network of induced dipole-

induced dipole interactions may exist. The image below illustrates a network
of induced dipole-induced dipole interactions.

6.1THE SHAPE OF MOLECULES


The shape of a covalent molecule can be predicted using the Valence
Shell Electron Pair Repul-sion (VSEPR) theory. This is a model in chemistry
that tries to predict the shapes of molecules.Very simply, VSEPR theory says
that the electron pairs in a molecule will arrange themselvesaround the
central atom of the molecule so that the repulsion between their negative
charges isas small as possible. In other words, the electron pairs arrange
themselves so that they are as far
apart as they can be. Depending on the number of electron pairs in the
molecule, it will have adifferent shape.
Valence shell electron pair repulsion theory (VSEPR) is a model in
chemistry, which is used to predict the shape of individual molecules, based
upon the extent of their electron-pair repulsion.
VSEPR theory is based on the idea that the geometry of a molecule is
mostly determined byrepulsion among the pairs of electrons around a central
atom. The pairs of electrons maybe bonding or non-bonding (also called lone
pairs). Only valence electrons of the central atom influence the molecular
shape in a meaningful way.

Das könnte Ihnen auch gefallen