Sie sind auf Seite 1von 73

1.

Pure substances
A pure substance is a sample of matter with both definite and
constant
composition with
distinct chemical
properties. A pure
substance can be
either an element
or a compound, but
the composition of
a pure substance
doesn’t vary.

 Elements
An element is composed of a single kind of atom. An atom is the
smallest particle of an element that still has all the properties of
the element.
Here’s an example: Gold is an element. If you slice and slice a
chunk of gold until only one tiny particle is left that can’t be
chopped any more without losing the properties that make
gold gold, then you’ve got an atom.
Allotropes
Non-metals Metals Metalloids
Carbon, Oxygen, Tin, Iron, Cobalt, Boron, Silicon,
Phosphorus, Polonium Arsenic, Germanium,
Selenium, Sulfur Antimony, Tellurium
Allotropes are two or more forms of the same element in the same
physical state (solid, liquid, or gas) that differ from each other in their
physical, and sometimes chemical, properties. The most
notable examples of allotropes are found in groups 14, 15, and 16 of
the periodic table. Gaseous oxygen, for example, exists in three
allotropic forms: monatomic oxygen (O), a diatomic molecule (O2), and
in a triatomic molecule known as ozone (O3). Allotropes are different
structural modifications of an element;[1] the atoms of the element
are bonded together in a different manner.

A striking example of differing physical properties among allotropes is


the case of carbon. Solid carbon exists in two allotropic
forms: diamond and graphite. Diamond is the hardest naturally occurring
substance and has the highest melting point (more than 6,335°F
[3,502°C]) of any element. In contrast, graphite is a very soft material,
the substance from which the "lead" in lead pencils is made.
The allotropes of phosphorus illustrate the variations in chemical
properties that may occur among such forms. White phosphorus,
for example, is a waxy white solid that bursts into flame spontaneously
when exposed to air. It is also highly toxic. On the other hand, a
second allotrope of phosphorus known as red phosphorus is far more
stable, does not react with air, and is essentially nontoxic.
Allotropes differ from each other structurally depending on the number
of atoms in a molecule of the element. There are allotropes of sulfur,
for example, that contain 2, 6, 7, 8, 10, 12, 18, and 20 atoms
per molecule (formulas S2 to S20). Several of these, however, are not

very stable.
 Compounds
A compound is composed of two or more elements in a specific
ratio. For example, water is a compound made up of two elements,
hydrogen (H) and oxygen (O). These elements are combined in a
very specific way — in a ratio of two hydrogen atoms to one
oxygen atom, known as:
Many compounds contain hydrogen and oxygen, but only one has
that special 2 to 1 ratio we call water. The compound water has
physical and chemical properties different from both hydrogen and
oxygen — water’s properties are a unique combination of the two
elements. IMPORTANT: Chemists can’t easily separate the
components of a compound: They have to resort to some
type of chemical reaction.

Mixtures
Mixtures are physical combinations of pure substances that have no
definite or constant composition — the composition of a mixture varies
according to who prepares the mixture.
A mixture refers to the physical combination of two or more substances
in which the identities are retained and are mixed in the form
of solutions, suspensions, and colloids.

Mixtures are the one product of a mechanical blending or mixing of


chemical substances such as elements and compounds, without
chemical bonding or other chemical change, so that each ingredient
substance retains its own chemical properties and makeup.
Although chemists have a difficult time separating compounds into their
specific elements, the different parts of a mixture can be easily
separated by physical means, such as filtration.
For example, suppose you have a mixture of salt and sand, and you want
to purify the sand by removing the salt. You can do this by adding water,
dissolving the salt, and then filtering the mixture. You then end up with
pure sand.
Mixtures can be either homogeneous or heterogeneous:
A homogeneous mixture, sometimes called a solution, is relatively
uniform in composition; every portion of the mixture is like every other
portion. For example, if you dissolve sugar in water and mix it really well,
your mixture is basically the same no matter where you sample it.
A heterogeneous mixture is a mixture whose composition varies from
position to position within the sample. For example, if you put some
sugar in a jar, add some sand, and then give the jar a couple of shakes,
your mixture doesn’t have the same composition throughout the jar.
Because the sand is heavier, there’s probably more sand at the bottom
of the jar and more sugar at the top.
Separation of mixtures

 Chromatography
Chromatography is a separation technique used to separate the different
components in a liquid mixture. It was introduced by a Russian Scientist
Michael Tswett. Chromatography involves the sample being dissolved in
a particular solvent called mobile phase. The mobile phase may be a gas
or liquid. The mobile phase is then passed through another phase called
stationary phase. The stationary phase may be a solid packed in a glass
plate or a piece of chromatography paper.
The various components of the mixture travel at different speeds,
causing them to separate. There are different types of chromatographic
techniques such as column chromatography, TLC, paper
chromatography, and gas chromatography.
Paper chromatography is one of the important chromatographic
methods. Paper chromatography uses paper as the stationary phase and
a liquid solvent as the mobile phase. In paper chromatography, the
sample is placed on a spot on the paper and the paper is carefully
dipped into a solvent. The solvent rises up the paper due to capillary
action and the components of the mixture rise up at different rates and
thus are separated from one another.
Applications:
 To separate colors in a dye.
 To separate pigments from natural colors.
 To separate drugs from blood.

 Centrifugation
Sometimes the solid particles in a liquid are very small and can pass
through a filter paper. For such particles, the filtration technique cannot
be used for separation. Such mixtures are separated by centrifugation.
So, centrifugation is the process of separation of insoluble materials from
a liquid where normal filtration does not work well. The centrifugation is
based on the size, shape, and density of the particles, viscosity of the
medium, and the speed of rotation. The principle is that the denser
particles are forced to the bottom and the lighter particles stay at the
top when spun rapidly.
The apparatus used for centrifugation is called a centrifuge. The
centrifuge consists of a centrifuge tube holder called rotor. The rotor
holds balanced centrifugal tubes of equal amounts of the solid-liquid
mixture. On rapid rotation of the rotor, the centrifuge tubes rotate
horizontally and due to the centrifugal force, the denser insoluble
particles separate from the liquid. When the rotation stops, the solid
particles end up at the bottom of the centrifuge tube with liquid at the
top.
Applications:
 Used in diagnostic laboratories for blood and urine tests.
 Used in dairies and home to separate butter from cream.
 Used in washing machines to squeeze water from wet clothes

 Simple distillation
Simple distillation is a method used for the separation of components of
a mixture containing two miscible liquids that boil without decomposition
and have sufficient difference in their boiling points.
The distillation process involves heating a liquid to its boiling points, and
transferring the vapors into the cold portion of the apparatus, then
condensing the vapors and collecting the condensed liquid in a
container. In this process, when the temperature of a liquid rises, the
vapor pressure of the liquid increases. When the vapor pressure of the
liquid and the atmospheric pressure reach the same level, the liquid
passes into its vapor state. The vapors pass over the heated portion of
the apparatus until they come into contact with the cold surface of the
water-cooled condenser. When the vapor cools, it condenses and passes
down the condenser and is collected into a receiver through the vacuum
adapter.
Applications:
 Separation of acetone and water.

 Distillation of alcohol
Fractional distillation: Fractional distillation is used for the separation
of a mixture of two or more miscible liquids for which the difference in
boiling points is less than 25K. The apparatus for fractional distillation is
similar to that of simple distillation, except that a fractionating column is
fitted in between the distillation flask and the condenser.
A simple fractionating column is a tube packed with glass beads. The
beads provide surface for the vapors to cool and condense repeatedly.
When vapors of a mixture are passed through the fractionating column,
because of the repeated condensation and evaporation, the vapors of
the liquid with the lower boiling point first pass out of the fractionating
column, condense and are collected in the receiver flask. The other
liquid, with a slightly higher boiling point, can be collected in similar
fashion in another receiver flask.
Applications:
 Separation of different fractions from petroleum products.
 Separation of a mixture of methanol and ethanol.

 Filtration
Filtration is any of various mechanical, physical or biological operations
that separate solids from fluids (liquids or gases) by adding a medium
through which only the fluid can pass. The fluid that passes through is
called the filtrate. In physical filters oversize solids in the fluid are
retained and in biological filters particulates are trapped and ingested
and metabolites are retained and removed. However, the separation is
not complete; solids will be contaminated with some fluid and filtrate will
contain fine particles (depending on the pore size, filter thickness and
biological activity).

 Precipitation
It is the creation of a solid from a solution. When the reaction occurs in a
liquid solution, the solid formed is called
the 'precipitate'. The chemical that causes
the solid to form is called the 'precipitant'.
Without sufficient force of gravity
(settling) to bring the solid particles
together, the precipitate remains
in suspension. After sedimentation,
especially when using a centrifuge to
press it into a compact mass, the
precipitate may be referred to as a 'pellet'.
Precipitation can be used as a medium.
The precipitate-free liquid remaining
above the solid is called the 'supernate' or
'supernatant'. Powders derived from precipitation have
also historically been known as 'flowers'. When the solid appears in the
form of cellulose fibers which have been through chemical processing,
the process is often referred to as regeneration.
 Magnetism
Purification
Purification in a chemical context is the physical separation of a
chemical substance of interest from foreign
or contaminating substances. Pure results of a successful purification
process are termed isolate. The following list of chemical purification
methods should not be considered exhaustive.
Affinity purification is used to purify proteins by retaining them on a
column through their affinity to antibodies, enzymes or receptors which
have been immobilised on the column.
Filtration is a mechanical method to
separate solids from liquids or gases by passing the feed stream through
a porous sheet such as a cloth or membrane, which retains the solids
and allows the liquid to pass through.
Centrifugation is a process in which light particles are revolved at high
speed with the help of an electric motor so that the fine particles which
do not settle at bottom would settle down.
Evaporation is used to remove volatile liquids from non-
volatile solutes which cannot be done through filtration due to the small
size of the substances.
Liquid–liquid extraction removes an impurity or recovers a desired
product by dissolving the crude material in a solvent in which other
components of the feed material are soluble.
Crystallization separates a product from a liquid feed stream, often in
extremely pure form, by cooling the feed stream or adding precipitants
which lower the solubility of the desired product so that it forms crystals.
The pure solid crystals are then separated from the remaining liquor by
filtration or centrifugation.
Recrystallization: In analytical and synthetic chemistry work,
purchased reagents of doubtful purity may be recrystallized, e.g.
dissolved in a very pure solvent, and then crystallized, and the crystals
recovered, in order to improve and/or verify their purity.
Adsorption removes a soluble impurity from a feed stream by trapping
it on the surface of a solid material such as activated carbon which
forms strong non-covalent chemical bonds with the
impurity. Chromatography employs adsorption and desorption on a
packed bed of a solid to purify multiple components of a single feed
stream.
Smelting is used to produce metals from raw ore, and involves
adding chemicals to the ore and heating it up to the melting point of the
metal.
Refining is used primarily in the petroleum industry, whereby crude oil
is heated and separated into stages according to the condensation
points of the various elements.
Downstream processing refers to purification of chemicals,
pharmaceuticals and food ingredients produced by fermentation or
synthesized by plant and animal tissues, for example antibiotics, citric
acid, vitamin E, and insulin.
Fractionation refers to a purification strategy in which some relatively
inefficient purification method is repeatedly applied to isolate the
desired substance in progressively greater purity.
Electrolysis refers to the breakdown of substances using an electric
current. This removes impurities in a substance that an electric current is
run through
Sublimation is the process of changing of any substance (usually on
heating) from a solid to a gas (or from gas to a solid) without passing
through liquid phase.
Bioleaching is the extraction of metals from their ores through the use
of living organisms.
2. States of matter

 Solid
In a solid the particles (ions, atoms or molecules) are closely packed
together. The forces between particles are strong so that the particles
cannot move freely but can only vibrate. As a result, a solid has a stable,
definite shape, and a definite volume. Solids can only change their
shape by force, as when broken or cut.
In crystalline solids, the particles (atoms, molecules, or ions) are packed
in a regularly ordered, repeating pattern. There are various
different crystal structures, and the same substance can have more than
one structure (or solid phase). For example, iron has a body-centred
cubic structure at temperatures below 912 °C, and a face-centred
cubic structure between 912 and 1394 °C. Ice has fifteen known crystal
structures, or fifteen solid phases, which exist at various temperatures
and pressures.[2]
Glasses and other non-crystalline, amorphous solids without long-range
order are not thermal equilibrium ground states; therefore they are
described below as nonclassical states of matter.
Solids can be transformed into liquids by melting, and liquids can be
transformed into solids by freezing. Solids can also change directly into
gases through the process of sublimation, and gases can likewise
change directly into solids through deposition.

 Liquid
A liquid is a nearly incompressible fluid that conforms to the shape of its
container but retains a (nearly) constant volume independent of
pressure. The volume is definite if the temperature and pressure are
constant. When a solid is heated above its melting point, it becomes
liquid, given that the pressure is higher than the triple point of the
substance. Intermolecular (or interatomic or interionic) forces are still
important, but the molecules have enough energy to move relative to
each other and the structure is mobile. This means that the shape of a
liquid is not definite but is determined by its container. The volume is
usually greater than that of the corresponding solid, the best known
exception being water, H2O. The highest temperature at which a given
liquid can exist is its critical temperature.[3]

 Gas
The spaces between gas molecules are very big. Gas molecules have
very weak or no bonds at all. The molecules in "gas" can move freely
and fast.
A gas is a compressible fluid. Not only will a gas conform to the shape of
its container but it will also expand to fill the container.
In a gas, the molecules have enough kinetic energy so that the effect of
intermolecular forces is small (or zero for an ideal gas), and the typical
distance between neighboring molecules is much greater than the
molecular size. A gas has no definite shape or volume, but occupies the
entire container in which it is confined. A liquid may be converted to a
gas by heating at constant pressure to the boiling point, or else by
reducing the pressure at constant temperature.
At temperatures below its critical temperature, a gas is also called
a vapor, and can be liquefied by compression alone without cooling. A
vapor can exist in equilibrium with a liquid (or solid), in which case the
gas pressure equals the vapor pressure of the liquid (or solid).
A supercritical fluid (SCF) is a gas whose temperature and pressure are
above the critical temperature and critical pressure respectively. In this
state, the distinction between liquid and gas disappears. A supercritical
fluid has the physical properties of a gas, but its high density confers
solvent properties in some cases, which leads to useful applications. For
example, supercritical carbon dioxide is used to extract caffeine in the

manufacture of decaffeinated coffee.

Particles constituting substances

Atom structure
All atoms are made from three subatomic particles
� Protons, neutron & electrons.
These particles have the following properties:
In the above table I have used a unit of mass called the atomic mass
unit (amu). This unit is much more convenient to use than grams for
describing masses of atoms. It is defined so that both protons and
neutrons have a mass of approximately 1 amu. Its precise definition

Particle Charge Mass (g) Mass (amu)

Proton +1 1.6727 x 10-24 g 1.007316

Neutron 0 1.6750 x 10-24 g 1.008701


Electron -1 9.110 x 10-28 g 0.000549

will be given later.


The important points to keep in mind are as follows:
Protons and neutrons have almost the same mass, while the electron is
approximately 2000 times lighter.
Protons and electrons carry charges of equal magnitude, but opposite
charge. Neutrons carry no charge (they are neutral).
It was once thought that protons, neutrons and electrons were spread
out in a rather uniform fashion to form the atom (see J.J. Thompson’s
plum pudding model of the atom on page 42), but now we know the
actual structure of the atom to be quite different.
What does an atom look like?
Protons and neutrons are held together rather closely in the center of
the atom. Together they make up the nucleus, which accounts for nearly
all of the mass of the atom.
Electrons move rapidly around the nucleus and constitute almost the
entire volume of the atom. Although quantum mechanics are necessary
to explain the motion of an electron about the nucleus, we can say that
the distribution of electrons about an atom is such that the atom has a
spherical shape.

Atoms have sizes on the order of 1-5 � (1 angstrom = 1 � = 1 � 10-


10 m) and masses on the order of 1-300 amu.
To put the mass and dimensions of an atom into perspective consider
the following analogies. If an atom were the size of Ohio stadium, the
nucleus would only be the size of a small marble. However, the mass of
that marble would be ~ 115 million tons.
What holds an atom together?
The negatively charged electron is attracted to the positively charged
nucleus by a Coulombic attraction.
The protons and neutrons are held together in the nucleus by the strong
nuclear force.
How many electrons, protons and neutrons are contained in an atom?
Atoms in their natural state have no charge, that is they are neutral.
Therefore, in a neutral atom the number of protons and electrons are the
same. If this condition is violated the atom has a net charge and is called
an ion.
The number of protons in the nucleus determines the identity of the
atom. For example all carbon atoms contain six protons, all gold atoms
contain 79 protons, all lead atoms contain 82 protons.
Two atoms with the same number of protons, but different numbers of
neutrons are called isotopes.
How does the structure of the atom relate to its properties?
Chemical reactions involve either the transfer or the sharing of electrons
between atoms. Therefore, the chemical reactivity/ properties of an
element is primarily dependent upon the number of electrons in an atom
of that element. Protons also play a significant role because the
tendency for an atom to either lose, gain or share electrons is dependent
upon the charge of the nucleus.
Therefore, we can say that the chemical reactivity of an atom is
dependent upon the number of electrons and protons, and independent
of the number of neutrons.
The mass and radioactive properties of an atom are dependent upon the
number of protons and neutrons in the nucleus.
Note: The number of protons, neutrons and electrons in an atom
completely determine its properties and identity, regardless of how and
where the atom was made. So it is inaccurate to speak of synthetic
atoms and natural atoms. In other words a lead atom is a lead atom, end
of story. It doesn’t matter if was mined from the earth, produced in a
nuclear reactor, or came to earth on an asteroid.
Symbolism
There is a symbolism or shorthand for describing atoms which is
universally used across all scientific disciplines:

Atomic Number (Z) � The # of protons

Mass Number (A) � [The # of protons] + [the # of neutrons]


The number of protons, neutrons and electrons in an atom are uniquely
specified by the following symbol
ASyC
where:

Sy = The elemental symbol (i.e. C, N, Cr) � defines the # of protons


A = The mass number � [# of protons] + [# of neutrons]

C = The net charge � [# of protons] – [# of electrons]


Example

Lets start with a neutral boron 10 atom � 10B


Since the atom is a boron atom the periodic table tells us that there
are 5 protons in the nucleus Z = 5.
The atom is neutral so that the number of electrons must balance the
number of protons, 5 electrons.
The mass number is 10, so that the number of neutrons is A - Z = 10 - 5
= 5 neutrons.
Electronic configuration
In atomic physics and quantum chemistry, the electron configuration is
the distribution of electrons of an atom or molecule (or other physical
structure) in atomic or molecular orbitals. For example, the electron
configuration of the neon atom is 1s2 2s2 2p6.
Electronic configurations describe each electron as moving
independently in an orbital, in an average field created by all other
orbitals. Mathematically, configurations are described by Slater
determinants or configuration state functions.
According to the laws of quantum mechanics, for systems with only one
electron, an energy is associated with each electron configuration and,
upon certain conditions, electrons are able to move from one
configuration to another by the emission or absorption of a quantum of
energy, in the form of a photon.
Knowledge of the electron configuration of different atoms is useful in
understanding the structure of the periodic table of elements. This is
also useful for describing the chemical bonds that hold atoms together.
In bulk materials, this same idea helps explain the peculiar properties
of lasers and semiconductors.
Electron shells and the Bohr model
An early model of the atom was developed in 1913 by the Danish
scientist Niels Bohr (1885–1962). The Bohr model shows the atom as a
central nucleus containing protons and neutrons, with the electrons in
circular electron shells at specific distances from the nucleus, similar to
planets orbiting around the sun. Each electron shell has a different
energy level, with those shells closest to the nucleus being lower in
energy than those farther from the nucleus. By convention, each shell is
assigned a number and the symbol n—for example, the electron shell
closest to the nucleus is called 1n. In order to move between shells, an
electron must absorb or release an amount of energy corresponding
exactly to the difference in energy between the shells. For instance, if an
electron absorbs energy from a photon, it may become excited and
move to a higher-energy shell; conversely, when an excited electron
drops back down to a lower-energy shell, it will release energy, often in
the form of heat.

Bohr model of an atom, showing energy levels as concentric circles


surrounding the nucleus. Energy must be added to move an electron
outward to a higher energy level, and energy is released when an
electron falls down from a higher energy level to a closer-in one.
Atoms, like other things governed by the laws of physics, tend to take on
the lowest-energy, most stable configuration they can. Thus, the
electron shells of an atom are populated from the inside out, with
electrons filling up the low-energy shells closer to the nucleus before
they move into the higher-energy shells further out. The shell closest to
the nucleus, 1n, can hold two electrons, while the next shell, 2n, can
hold eight, and the third shell, 3n, can hold up to eighteen.
The number of electrons in the outermost shell of a particular atom
determines its reactivity, or tendency to form chemical bonds with other
atoms. This outermost shell is known as the valence shell, and the
electrons found in it are called valence electrons. In general, atoms
are most stable, least reactive, when their outermost electron shell is
full. Most of the elements important in biology need eight electrons in
their outermost shell in order to be stable, and this rule of thumb is
known as the octet rule. Some atoms can be stable with an octet even
though their valence shell is the 3n shell, which can hold up to 18
electrons. We will explore the reason for this when we discuss electron
orbitals below.
Examples of some neutral atoms and their electron configurations are
shown below. In this table, you can see that helium has a full valence
shell, with two electrons in its first and only, 1n, shell. Similarly, neon
has a complete outer 2n shell containing eight electrons. These electron
configurations make helium and neon very stable. Although argon does
not technically have a full outer shell, since the 3n shell can hold up to
eighteen electrons, it is stable like neon and helium because it has eight
electrons in the 3n shell and thus satisfies the octet rule. In contrast,
chlorine has only seven electrons in its outermost shell, while sodium
has just one. These patterns do not fill the outermost shell or satisfy the
octet rule, making chlorine and sodium reactive, eager to gain or lose
electrons to reach a more stable configuration.
Bohr diagrams of various elements
Image credit: OpenStax Biology
Electron configurations and the periodic table
Elements are placed in order on the periodic table based on their atomic
number, how many protons they have. In a neutral atom, the number of
electrons will equal the number of protons, so we can easily determine
electron number from atomic number. In addition, the position of an
element in the periodic table—its column, or group, and row, or period—
provides useful information about how those electrons are arranged.
If we consider just the first three rows of the table, which include the
major elements important to life, each row corresponds to the filling of a
different electron shell: helium and hydrogen place their electrons in the
1n shell, while second-row elements like Li start filling the 2n shell, and
third-row elements like Na continue with the 3n shell. Similarly, an
element’s column number gives information about its number of valence
electrons and reactivity. In general, the number of valence electrons is
the same within a column and increases from left to right within a row.
Group 1 elements have just one valence electron and group 18 elements
have eight, except for helium, which has only two electrons total. Thus,
group number is a good predictor of how reactive each element will be:
Helium (\text{He}HeH, e), neon (\text{Ne}NeN, e), and argon
(\text{Ar}ArA, r), as group 18 elements, have outer electron shells that
are full or satisfy the octet rule. This makes them highly stable as single
atoms. Because of their non-reactivity, they are called the inert
gases or noble gases.
Hydrogen (\text{H}HH), lithium (\text{Li}LiL, i), and sodium
(\text{Na}NaN, a), as group 1 elements, have just one electron in their
outermost shells. They are unstable as single atoms, but can become
stable by losing or sharing their one valence electron. If these elements
fully lose an electron—as \text{Li}LiL, i and \text{Na}NaN, a typically do
—they become positively charged ions: \text{Li}^+Li+L, i, start
superscript, plus, end superscript and \text{Na}^+Na+N, a, start
superscript, plus, end superscript.
Fluorine (\text{F}FF) and chlorine (\text{Cl}ClC, l), as group 17
elements, have seven electrons in their outermost shells. They tend to
achieve a stable octet by taking an electron from other atoms, becoming
negatively charged ions: \text{F}^-F−F, start superscript, minus, end
superscript and \text{Cl}^-Cl−C, l, start superscript, minus, end
superscript.
Carbon (\text{C}CC), as a group 14 element, has four electrons in its
outer shell. Carbon typically shares electrons to achieve a complete
valence shell, forming bonds with multiple other atoms.
Thus, the columns of the periodic table reflect the number of electrons
found in each element’s valence shell, which in turn determines how the
element will react.
Subshells and orbitals
The Bohr model is useful to explain the reactivity and chemical bonding
of many elements, but it actually doesn’t give a very accurate
description of how electrons are distributed in space around the nucleus.
Specifically, electrons don’t really circle the nucleus, but rather spend
most of their time in sometimes-complex-shaped regions of space
around the nucleus, known as electron orbitals. We can’t actually
know where an electron is at any given moment in time, but we can
mathematically determine the volume of space in which it is most likely
to be found—say, the volume of space in which it will spend 90% of its
time. This high-probability region makes up an orbital, and each orbital
can hold up to two electrons.
So, how do these
mathematically defined
orbitals fit in with the
electron shells we saw in
the Bohr model? We can
break each electron shell
down into one or more
subshells, which are
simply sets of one or
more orbitals. Subshells
are designated by the
letters sss, ppp, ddd,
and fff, and each letter
indicates a different shape. For instance, sss subshells have a single,
spherical orbital, while ppp subshells contain three dumbbell-shaped
orbitals at right angles to each other. Most of organic chemistry—the
chemistry of carbon-containing compounds, which are central to biology
—involves interactions between electrons in sss and ppp subshells, so
these are the most important subshell types to be familiar with.
However, atoms with many electrons may place some of their electrons
in ddd and fff subshells. Subshells ddd and fff have more complex
shapes and contain five and seven orbitals, respectively.
3D diagram of circular 1s and 2s orbitals and dumbbell-shaped 2p
orbitals. There are three 2p orbitals, and they are at right angles to each
other.
Image credit: modified from OpenStax Biology
The first electron shell, 1n, corresponds to a single 1s1s1, s orbital.
The 1s1s1, s orbital is the closest orbital to the nucleus, and it fills with
electrons first, before any other orbital. Hydrogen has just one electron,
so it has a single spot in the 1s1s1, sorbital occupied. This can be written
out in a shorthand form called an electron configuration as 1s^ 11s1
1, s, start superscript, 1, end superscript, where the superscripted 1
refers to the one electron in the 1s1s1, s orbital. Helium has two
electrons, so it can completely fill the1s1s1, s orbital with its two
electrons. This is written out as 1s^ 21s21, s, start superscript, 2, end
superscript, referring to the two electrons of helium in the 1s1s1,
s orbital. On the periodic table, hydrogen and helium are the only two
elements in the first row, or period, which reflects that they only have
electrons in their first shell. Hydrogen and helium are the only two
elements that have electrons exclusively in the 1s1s1, s orbital in their
neutral, non-charged, state.
The second electron shell, 2n, contains another spherical sss orbital plus
three dumbbell-shaped ppp orbitals, each of which can hold two
electrons. After the 1s1s1, s orbital is filled, the second electron shell
begins to fill, with electrons going first into the 2s2s2, s orbital and then
into the three ppp orbitals. Elements in the second row of the periodic
table place their electrons in the 2n shell as well as the 1n shell. For
instance, lithium (\text{Li}LiL, i) has three electrons: two fill the 1s1s1,
sorbital, and the third is placed in the 2s2s2, s orbital, giving an electron
configuration of 1s^ 21s21, s, start superscript, 2, end superscript 2s^
12s12, s, start superscript, 1, end superscript. Neon (\text{Ne}NeN, e),
on the other hand, has a total of ten electrons: two are in its
innermost 1s1s1, s orbital and eight fill the second shell—two each in
the 2s2s2, s and three ppp orbitals, 1s^ 21s21, s, start superscript, 2,
end superscript 2s^ 22s22, s, start superscript, 2, end
superscript 2p^62p62, p, start superscript, 6, end superscript. Because
its 2n shell is filled, it is energetically stable as a single atom and will
rarely form chemical bonds with other atoms.
The third electron shell, 3n, also contains an sss orbital and
three ppp orbitals, and the third-row elements of the periodic table place
their electrons in these orbitals, much as second-row elements do for the
2n shell. The 3n shell also contains a ddd orbital, but this orbital is
considerably higher in energy than the 3s3s3, s and 3p3p3, p orbitals
and does not begin to fill until the fourth row of the periodic table. This is
why third-row elements, such as argon, can be stable with just eight
valence electrons: their sss and ppp subshells are filled, even though the
entire 3n shell is not.
While electron shells and orbitals are closely related, orbitals provide a
more accurate picture of the electron configuration of an atom. That’s
because orbitals actually specify the shape and position of the regions of
space that electrons occupy.
The periodic law
It was developed independently by Dmitri Mendeleev and Lothar Meyer
in 1869. Mendeleev created the first periodic table and was shortly
followed by Meyer. They both arranged the elements by their mass and
proposed that certain properties periodically reoccur. Meyer formed his
periodic law based on the atomic volume or molar volume, which is the
atomic mass divided by the density in solid form. Mendeleev's table is
noteworthy because it exhibits mostly accurate values for atomic mass
and it also contains blank spaces for unknown elements.
Introduction
In 1804 physicist John Dalton advanced the atomic theory of
matter, helping scientists determine the mass of the known elements.
Around the same time, two chemists Sir Humphry Davy and Michael
Faraday developed electrochemistry which aided in the discovery of
new elements. By 1829, chemist Johann Wolfgang
Doberiner observed that certain elements with similar properties occur
in group of three such as; chlorine, bromine, iodine; calcium, strontium,
and barium; sulfur, selenium, tellurium; iron, cobalt, manganese.
However, at the time of this discovery too few elements had been
discovered and there was confusion between molecular weight and
atomic weights; therefore, chemists never really understood the
significance of Doberiner's triad.
In 1859 two physicists Robert Willhem Bunsen and Gustav Robert
Kirchof discovered spectroscopy which allowed for discovery of many
new elements. This gave scientists the tools to reveal the relationships
between elements. Thus in 1864, chemist John A. R Newland arranged
the elements in increasing of atomic weights. Explaining that a given set
of properties reoccurs every eight place, he named it the law of Octaves.
The Periodic Law
In 1869, Dmitri Mendeleev and Lothar Meyer individually came up
with their own periodic law "when the elements are arranged in order of
increasing atomic mass, certain sets of properties recur
periodically." Meyer based his laws on the atomic volume (the atomic
mass of an element divided by the density of its solid form), this
property is called Molar volume.
Atomic (molar) volume (cm3/mol)= molar mass (g/ mol)ρ (cm3/g)Atomic
(molar) volume (cm3/mol)= molar mass (g/ mol)ρ (cm3/g)
Mendeleev's Periodic Table
Mendeleev's periodic table is an arrangement of the elements that group
similar elements together. He left blank spaces for the undiscovered
elements (atomic masses, element: 44, scandium; 68, gallium; 72,
germanium; & 100, technetium) so that certain elements can be
grouped together. However, Mendeleev had not predicted the noble
gases, so no spots were left for them.
Example
The alkali metals (Mendeleev's group I) have high molar volumes and
they also have low melting points which decrease in the order:

Li (174 oC) > Na (97.8 oC) > K (63.7 oC) > Rb (38.9 oC) > Cs (28.5 oC)
Atomic Number as the Basis for the Periodic Law
Assuming there were errors in atomic masses, Mendeleev placed certain
elements not in order of increasing atomic mass so that they could fit
into the proper groups (similar elements have similar properties) of his
periodic table. An example of this was with argon (atomic mass 39.9),
which was put in front of potassium (atomic mass 39.1). Elements were
placed into groups that expressed similar chemical behavior.
In 1913 Henry G.J. Moseley did researched the X-Ray spectra of the
elements and suggested that the energies of electron orbitals depend on
the nuclear charge and the nuclear charges of atoms in the target, which
is also known as anode, dictate the frequencies of emitted X-Rays.
Moseley was able to tie the X-Ray frequencies to numbers equal to the
nuclear charges, therefore showing the placement of the elements in
Mendeleev's periodic table. The equation he used:
ν=A(Z−b)2ν=A(Z−b)2
with
νν: X-Ray frequency
ZZ: Atomic Number
AA and bb: constants

Note
With Moseley's contribution the Periodic Law can be restated:
Similar properties recur periodically when elements are arranged
according to increasing atomic number."

Atomic numbers, not weights, determine the factor of chemical


properties. As mentioned before, argon weights more than potassium
(39.9 vs. 39.1, respectively), yet argon is in front of potassium. Thus, we
can see that elements are arranged based on their atomic number. The
periodic law is found to help determine many patterns of many different
properties of elements; melting and boiling points, densities, electrical
conductivity, reactivity, acidic, basic, valance, polarity, and solubility.
The table below shows that elements increase from left to right
accordingly to their atomic number. The vertical columns have similar
properties within their group for example Lithium is similar to sodium,
beryllium is similar to magnesium, and so on.

Gr 1 2 1 1 1 1 1 1
ou 3 4 5 6 7 8
p

Ele L B B C N O F N
m i e e
en
t

At 3 4 5 6 7 8 9 1
o 0
mi
c
Nu
m
be
r

At 6 9 1 1 1 1 1 2
o . . 0 2 4 5 8 0
mi 9 0 . . . . . .
c 4 1 8 0 0 9 9 1
Ma 1 1 1 9 9 8
ss

Ele N M A S P S C A
m a g l i l r
en
t

At 1 1 1 1 1 1 1 1
o 1 2 3 4 5 6 7 8
mi
c
Nu
m
be
r

At 2 2 2 2 3 3 3 3
o 2 4 6 0 0 2 5 9
mi . . . . . . . .
c 9 3 9 0 9 0 4 9
Ma 9 1 8 9 7 7 5 5
ss
So, elements in Group 1 (periodic table) have similar chemical
properties, they are called alkali metals. Elements in Group 2 have
similar chemical properties, they are called the alkaline earth metals.
Short form periodic table
The short form periodic table is a table where elements are arranged in 7
rows, periods, with increasing atomic numbers from left to right. There
are 18 vertical columns known as groups. This table is based on
Mendeleev's periodic table and the periodic law.
Long form Periodic Table
In the long form, each period correlates to the building up of electronic
shell; the first two groups (1-2) (s-block) and the last 6 groups (13-18)
(p-block) make up the main-group elements and the groups (3-12) in
between the s and p blocks are called the transition metals. Group 18
elements are called noble gases, and group 17 are called halogens. The
f-block elements, called inner transition metals, which are at the bottom
of the periodic table (periods 8 and 9); the 15 elements after barium
(atomic number 56) are called lanthanides and the 14 elements after
radium (atomic number 88) are called actinides.
Problems
1) The periodic law states that
similar properties recur periodically when elements are arranged
according to increasing atomic number
similar properties recur periodically when elements are arranged
according to increasing atomic weight
similar properties are everywhere on the periodic table
elements in the same period have same characteristics
2) Which element is most similar to Sodium
Potassium
Aluminum
Oxygen
Calcium
3) According to the periodic law, would argon be in front of
potassium or after? Explain why.
4) Which element is most similar to Calcium?
Carbon
Oxygen
Strontium
Iodine
5) Who were the two chemists that came up with the periodic
law?
John Dalton and Michael Faraday
Dmitri Mendeleev and Lothar Meyer
Michael Faraday and Lothar Meyer
John Dalton and Dmitri Mendeleev
Answers
A
A
Argon would in front of potassium because the periodic law states that
the periodic table increases from left to right based on atomic number
not atomic weights
C
B

The basic phase diagram


What is a phase?
At its simplest, a phase can be just another term for solid,
liquid or gas. If you have some ice floating in water, you
have a solid phase present and a liquid phase. If there is
air above the mixture, then that is another phase.
But the term can be used more generally than this. For
example, oil floating on water also consists of two phases -
in this case, two liquid phases. If the oil and water are
contained in a bucket, then the solid bucket is yet another
phase. In fact, there might be more than one solid phase if
the handle is attached separately to the bucket rather than
moulded as a part of the bucket.
You can recognise the presence of the different phases
because there is an obvious boundary between them - a
boundary between the solid ice and the liquid water, for
example, or the boundary between the two liquids.

Phase diagrams
A phase diagram lets you work out exactly what phases
are present at any given temperature and pressure. In the
cases we'll be looking at on this page, the phases will
simply be the solid, liquid or vapour (gas) states of a pure
substance.
This is the phase diagram for a typical pure substance.

These diagrams (including this one) are nearly always


drawn highly distorted in order to see what is going on
more easily. There are usually two major distortions. We'll
discuss these when they become relevant.
If you look at the diagram, you will see that there are three
lines, three areas marked "solid", "liquid" and "vapour",
and two special points marked "C" and "T".
The three areas
These are easy! Suppose you have a pure substance at
three different sets of conditions of temperature and
pressure corresponding to 1, 2 and 3 in the next diagram.

Under the set of conditions at 1 in the diagram, the


substance would be a solid because it falls into that area of
the phase diagram. At 2, it would be a liquid; and at 3, it
would be a vapour (a gas).

Note: I'm using the terms vapour and gas as if


they were interchangeable. There are subtle
differences between them that I'm not ready to
explain for a while yet. Be patient!

Moving from solid to liquid by changing the temperature:


Suppose you had a solid and increased the temperature
while keeping the pressure constant - as shown in the next
diagram. As the temperature increases to the point where
it crosses the line, the solid will turn to liquid. In other
words, it melts.
If you repeated this at a higher fixed pressure, the melting
temperature would be higher because the line between the
solid and liquid areas slopes slightly forward.

Note: This is one of the cases where we distort


these diagrams to make them easier to discuss.
This line is much more vertical in practice than we
normally draw it. There would be very little
change in melting point at a higher pressure. The
diagram would be very difficult to follow if we
didn't exaggerate it a bit.

So what actually is this line separating the solid and liquid


areas of the diagram?
It simply shows the effect of pressure on melting point.
Anywhere on this line, there is an equilibrium between
solid and liquid.
You can apply Le Chatelier's Principle to this equilibrium
just as if it was a chemical equilibrium. If you increase the
pressure, the equilibrium will move in such a way as to
counter the change you have just made.

If it converted from liquid to solid, the pressure would tend


to decrease again because the solid takes up slightly less
space for most substances.
That means that increasing the pressure on the equilibrium
mixture of solid and liquid at its original melting point will
convert the mixture back into the solid again. In other
words, it will no longer melt at this temperature.
To make it melt at this higher pressure, you will have to
increase the temperature a bit. Raising the pressure raises
the melting point of most solids. That's why the melting
point line slopes forward for most substances.
Moving from solid to liquid by changing the pressure:
You can also play around with this by looking at what
happens if you decrease the pressure on a solid at
constant temperature.

Note: You have got to be a bit careful about this,


because exactly what happens if you decrease
the pressure depends on exactly what your
starting conditions are. We'll talk some more
about this when we look at the line separating the
solid region from the vapour region.

Moving from liquid to vapour:


In the same sort of way, you can do this either by changing
the temperature or the pressure.

The liquid will change to a vapour - it boils - when it


crosses the boundary line between the two areas. If it is
temperature that you are varying, you can easily read off
the boiling temperature from the phase diagram. In the
diagram above, it is the temperature where the red arrow
crosses the boundary line.
So, again, what is the significance of this line separating
the two areas?
Anywhere along this line, there will be an equilibrium
between the liquid and the vapour. The line is most easily
seen as the effect of pressure on the boiling point of the
liquid.
As the pressure increases, so the boiling point increases.

Note: I don't want to make any very big deal over


this, but this line is actually exactly the same as
the graph for the effect of temperature on the
saturated vapour pressure of the liquid. Saturated
vapour pressure is dealt with on a separate page.
A liquid will boil when its saturated vapour
pressure is equal to the external pressure.
Suppose you measured the saturated vapour
pressure of a liquid at 50°C, and it turned out to
be 75 kPa. You could plot that as one point on a
vapour pressure curve, and then go on to
measure other saturated vapour pressures at
different temperatures and plot those as well.
Now, suppose that you had the liquid exposed to
a total external pressure of 75 kPa, and gradually
increased the temperature. The liquid would boil
when its saturated vapour pressure became equal
to the external pressure - in this case at 50°C. If
you have the complete vapour pressure curve,
you could equally well find the boiling point
corresponding to any other external pressure.
That means that the plot of saturated vapour
pressure against temperature is exactly the same
as the curve relating boiling point and external
pressure - they are just two ways of looking at the
same thing.
If all you are interested in doing is interpreting
one of these phase diagrams, you probably don't
have to worry too much about this.

The critical point


You will have noticed that this liquid-vapour equilibrium
curve has a top limit that I have labelled as C in the phase
diagram.
This is known as the critical point. The temperature and
pressure corresponding to this are known as the critical
temperature and critical pressure.
If you increase the pressure on a gas (vapour) at a
temperature lower than the critical temperature, you will
eventually cross the liquid-vapour equilibrium line and the
vapour will condense to give a liquid.

This works fine as long as the gas is below the critical


temperature. What, though, if your temperature
was above the critical temperature? There wouldn't be any
line to cross!
That is because, above the critical temperature, it is
impossible to condense a gas into a liquid just by
increasing the pressure. All you get is a highly compressed
gas. The particles have too much energy for the
intermolecular attractions to hold them together as a
liquid.
The critical temperature obviously varies from substance
to substance and depends on the strength of the
attractions between the particles. The stronger the
intermolecular attractions, the higher the critical
temperature.

Note: This is now a good point for a quick


comment about the use of the words "gas" and
"vapour". To a large extent you just use the term
which feels right. You don't usually talk about
"ethanol gas", although you would say "ethanol
vapour". Equally, you wouldn't talk about oxygen
as being a vapour - you always call it a gas.
There are various guide-lines that you can use if
you want to. For example, if the substance is
commonly a liquid at or around room
temperature, you tend to call what comes away
from it a vapour. A slightly wider use would be to
call it a vapour if the substance is below its
critical point, and a gas if it is above it. Certainly
it would be unusual to call anything a vapour if it
was above its critical point at room temperature -
oxygen or nitrogen or hydrogen, for example.
These would all be described as gases.
This is absolutely NOT something that is at all
worth getting worked up about!

Moving from solid to vapour:


There's just one more line to look at on the phase diagram.
This is the line in the bottom left-hand corner between the
solid and vapour areas.
That line represents solid-vapour equilibrium. If the
conditions of temperature and pressure fell exactly on that
line, there would be solid and vapour in equilibrium with
each other - the solid would be subliming. (Sublimation is
the change directly from solid to vapour or vice versa
without going through the liquid phase.)
Once again, you can cross that line by either increasing
the temperature of the solid, or decreasing the pressure.
The diagram shows the effect of increasing the
temperature of a solid at a (probably very low) constant
pressure. The pressure obviously has to be low enough
that a liquid can't form - in other words, it has to happen
below the point labelled as T.

You could read the sublimation temperature off the


diagram. It will be the temperature at which the line is
crossed.

The triple point


Point T on the diagram is called the triple point.
If you think about the three lines which meet at that point,
they represent conditions of:
solid-liquid equilibrium
liquid-vapour equilibrium
solid-vapour equilibrium
Where all three lines meet, you must have a unique
combination of temperature and pressure where all three
phases are in equilibrium together. That's why it is called
a triple point.
If you controlled the conditions of temperature and
pressure in order to land on this point, you would see an
equilibrium which involved the solid melting and subliming,
and the liquid in contact with it boiling to produce a vapour
- and all the reverse changes happening as well.
If you held the temperature and pressure at those values,
and kept the system closed so that nothing escaped, that's
how it would stay. A strange set of affairs!

Normal melting and boiling points


The normal melting and boiling points are those when the
pressure is 1 atmosphere. These can be found from the
phase diagram by drawing a line across at 1 atmosphere
pressure.

The phase diagram for water


There is only one difference between this and the phase
diagram that we've looked at up to now. The solid-liquid
equilibrium line (the melting point line) slopes backwards
rather than forwards.
In the case of water, the melting point gets lower at higher
pressures. Why?

If you have this equilibrium and increase the pressure on


it, according to Le Chatelier's Principle the equilibrium will
move to reduce the pressure again. That means that it will
move to the side with the smaller volume. Liquid water is
produced.
To make the liquid water freeze again at this higher
pressure, you will have to reduce the temperature. Higher
pressures mean lower melting (freezing) points.
Now lets put some numbers on the diagram to show the
exact positions of the critical point and triple point for
water.
Notice that the triple point for water occurs at a very low
pressure. Notice also that the critical temperature is
374°C. It would be impossible to convert water from a gas
to a liquid by compressing it above this temperature.
The normal melting and boiling points of water are found in
exactly the same way as we have already discussed - by
seeing where the 1 atmosphere pressure line crosses the
solid-liquid and then the liquid-vapour equilibrium lines.

Note: Further up the page I mentioned two ways


in which these diagrams are distorted to make
them easier to follow. I have already pointed out
that the solid-liquid equilibrium line should really
be much more vertical. This last diagram
illustrates the other major distortion - which is to
the scales of both pressure and temperature.
Look, for example, at the gaps between the
various quoted pressure figures and then imagine
that you had to plot those on a bit of graph
paper! The temperature scale is equally
haphazard.

Just one final example of using this diagram (because it


appeals to me). Imagine lowering the pressure on liquid
water along the line in the diagram below.

The phase diagram shows that the water would first freeze
to form ice as it crossed into the solid area. When the
pressure fell low enough, the ice would then sublime to
give water vapour. In other words, the change is from
liquid to solid to vapour. I find that satisfyingly bizarre!

The phase diagram for carbon dioxide

The only thing special about this phase diagram is the


position of the triple point which is well above atmospheric
pressure. It is impossible to get any liquid carbon dioxide
at pressures less than 5.11 atmospheres.
That means that at 1 atmosphere pressure, carbon dioxide
will sublime at a temperature of -78°C.
This is the reason that solid carbon dioxide is often known
as "dry ice". You can't get liquid carbon dioxide under
normal conditions - only the solid or the vapour.
Substances and chemical bonds
Ionic bonding
Ionic bonding is a type of chemical bond that involves the electrostatic
attraction between oppositely charged ions, and is the primary
interaction occurring in ionic compounds. The ions are atoms that have
gained one or more electrons (known as anions, which are negatively
charged) and atoms that have lost one or more electrons (known
as cations, which are positively charged). This transfer of electrons is
known as electrovalence in contrast to covalence. In the simplest case,
the cation is a metal atom and the anion is a nonmetal atom, but these
ions can be of a more complex nature, e.g. molecular ions like NH4+or
SO42−. In simpler words, an ionic bond is the transfer of electrons from
a metal to a non-metal in order for both atoms to obtain a full valence
shell.
It is important to recognize that clean ionic bonding – in which one atom
or molecule completely share an electron from another – cannot exist: all
ionic compounds have some degree of covalent bonding, or electron
sharing. Thus, the term "ionic bonding" is given when the ionic character
is greater than the covalent character—that is, a bond in which a
large electronegativity difference exists between the two atoms, causing
the bonding to be more polar (ionic) than in covalent bonding where
electrons are shared more equally. Bonds with partially ionic and
partially covalent character are called polar covalent bonds.
Ionic compounds conduct electricity when molten or in solution, typically
as a solid. Ionic compounds generally have a high melting point,
depending on the charge of the ions they consist of. The higher the
charges the stronger the cohesive forces and the higher the melting
point. They also tend to be soluble in water. Here, the opposite trend
roughly holds: the weaker the cohesive forces, the greater the solubility.
Ionic bonding can result from a redox reaction when atoms of an
element (usually metal), whose ionization energy is low, give some of
their electrons to achieve a stable electron configuration. In doing so,
cations are formed. The atom of another element (usually nonmetal),
whose electron affinity is positive, then accepts the electron(s), again to
attain a stable electron configuration, and after accepting electron(s) the
atom becomes an anion. Typically, the stable electron configuration is
one of the noble gases for elements in the s-block and the p-block, and
particular stable electron configurations for d-block and f-
block elements. The electrostatic attraction between the anions and
cations leads to the formation of a solid with a crystallographic lattice in
which the ions are stacked in an alternating fashion. In such a lattice, it
is usually not possible to distinguish discrete molecular units, so that the
compounds formed are not molecular in nature. However, the ions
themselves can be complex and form molecular ions like the acetate
anion or the ammonium cation.
For example, common table salt is sodium chloride. When sodium (Na)
and chlorine (Cl) are combined, the sodium atoms each lose an electron,
forming cations (Na+), and the chlorine atoms each gain an electron to
form anions (Cl−). These ions are then attracted to each other in a 1:1
ratio to form sodium chloride (NaCl).
Na + Cl → Na+ + Cl− → NaCl
However, to maintain charge neutrality, strict ratios between anions and
cations are observed so that ionic compounds, in general, obey the rules
of stoichiometry despite not being molecular compounds. For
compounds that are transitional to the alloys and possess mixed ionic
and metallic bonding, this may not be the case anymore. Many sulfides,
e.g., do form non-stoichiometric compounds.
Many ionic compounds are referred to as salts as they can also be
formed by the neutralization reaction of an Arrhenius base like NaOH
with an Arrhenius acid like HCl
NaOH + HCl → NaCl + H2O
The salt NaCl is then said to consist of the acid rest Cl− and the base
rest Na+.

Representation of ionic bonding between lithium and fluorine to


form lithium fluoride. Lithium has a low ionization energy and readily
gives up its lone valence electron to a fluorine atom, which has a
positive electron affinity and accepts the electron that was donated by
the lithium atom. The end-result is that lithium
is isoelectronic with helium and fluorine is isoelectronic with neon.
Electrostatic interaction occurs between the two resulting ions, but
typically aggregation is not limited to two of them. Instead, aggregation
into a whole lattice held together by ionic bonding is the result.
The removal of electrons from the cation is endothermic, raising the
system's overall energy. There may also be energy changes associated
with breaking of existing bonds or the addition of more than one electron
to form anions. However, the action of the anion's accepting the cation's
valence electrons and the subsequent attraction of the ions to each
other releases (lattice) energy and, thus, lowers the overall energy of the
system.
Ionic bonding will occur only if the overall energy change for the reaction
is favorable. In general, the reaction is exothermic, but, e.g., the
formation of mercuric oxide (HgO) is endothermic. The charge of the
resulting ions is a major factor in the strength of ionic bonding, e.g. a
salt C+A− is held together by electrostatic forces roughly four times
weaker than C2+A2− according to Coulombs law, where C and A
represent a generic cation and anion respectively. Of course the sizes of
the ions and the particular packing of the lattice are ignored in this
simple argument.
Ionic crystals
In chemistry, an ionic compound is a chemical compound composed
of ions held together by electrostatic forces termed ionic bonding. The
compound is neutral overall, but consists of positively charged ions
called cations and negatively charged ions called anions. These can
be simple ions such as the sodium (Na+) and chloride (Cl−) in sodium
chloride, or polyatomic species such as the ammonium (NH+
4) and carbonate (CO2−
3) ions in ammonium carbonate. Individual ions within an ionic
compound usually have multiple nearest neighbours, so are not
considered to be part of molecules, but instead part of a continuous
three-dimensional network, usually in a crystalline structure.
Ionic compounds containing hydrogen ions (H+) are classified as acids,
and those containing basic ions hydroxide (OH−) or oxide (O2−) are
classified as bases. Ionic compounds without these ions are also known
as salts and can be formed by acid–base reactions. Ionic compounds can
also be produced from their constituent ions by evaporation of
their solvent, precipitation, freezing, a solid-state reaction, or
the electron transfer reaction of reactive metals with reactive non-
metals, such as halogen gases.
Ionic compounds typically have high melting and boiling points, and
are hard and brittle. As solids they are almost always electrically
insulating, but when melted or dissolved they become highly conductive,
because the ions are mobilized.
Acidity/basicity
Ionic compounds containing hydrogen ions (H+) are classified as acids,
and those containing electropositive cations[56] and basic anions
ions hydroxide (OH−) or oxide (O2−) are classified as bases. Other ionic
compounds are known as salts and can be formed by acid–base
reactions.[57] If the compound is the result of a reaction between
a strong acid and a weak base, the result is an acidic salt. If it is the
result of a reaction between a strong base and a weak acid, the result is
a basic salt. If it is the result of a reaction between a strong acid and a
strong base, the result is a neutral salt. Weak acids reacted with weak
bases can produce ionic compounds with both the conjugate base ion
and conjugate acid ion, such as ammonium acetate.
Some ions are classed as amphoteric, being able to react with either an
acid or a base.[58] This is also true of some compounds with ionic
character, typically oxides or hydroxides of less-electropositive metals
(so the compound also has significant covalent character), such as zinc
oxide, aluminium hydroxide, aluminium oxide and lead(II) oxide.
Melting and boiling points
Electrostatic forces between particles are strongest when the charges
are high, and the distance between the nuclei of the ions is small. In
such cases, the compounds generally have very high melting and boiling
points and a low vapour pressure.[60] Trends in melting points can be
even better explained when the structure and ionic size ratio is taken
into account.[61] Above their melting point ionic solids melt and
become molten salts (although some ionic compounds such
as aluminium chloride and iron(III) chloride show molecule-like structures
in the liquid phase).[62] Inorganic compounds with simple ions typically
have small ions, and thus have high melting points, so are solids at room
temperature. Some substances with larger ions, however, have a
melting point below or near room temperature (often defined as up to
100 °C), and are termed ionic liquids.[63]Ions in ionic liquids often have
uneven charge distributions, or bulky substituents like hydrocarbon
chains, which also play a role in determining the strength of the
interactions and propensity to melt.[64]
Even when the local structure and bonding of an ionic solid is disrupted
sufficiently to melt it, there are still strong long-range electrostatic forces
of attraction holding the liquid together and preventing ions boiling to
form a gas phase.[65] This means that even room temperature ionic
liquids have low vapour pressures, and require substantially higher
temperatures to boil.[65] Boiling points exhibit similar trends to melting
points in terms of the size of ions and strength of other interactions.
[65] When vapourized, the ions are still not freed of one another. For
example, in the vapour phase sodium chloride exists as diatomic
"molecules".
Brittleness
Most ionic compounds are very brittle. Once they reach the limit of their
strength, they cannot deform mealleably, because the strict alignment
of positive and negative ions must be maintained. Instead the material
undergoes fracture via cleavage.[67] As the temperature is elevated
(usually close to the melting point) a ductile–brittle transition occurs,
and plastic flow becomes possible by the motion of dislocations.
Compressibility
The compressibility of an ionic compound is strongly determined by its
structure, and in particular the coordination number. For example,
halides with the caesium chloride structure (coordination number 8) are
less compressible than those with the sodium chloride structure
(coordination number 6), and less again than those with a coordination
number of 4.
Solubility
When ionic compounds dissolve, the individual ions dissociate and
are solvated by the solvent and dispersed throughout the resulting
solution.[70] Because the ions are released into solution when dissolved,
and can conduct charge, soluble ionic compounds are the most common
class of strong electrolytes, and their solutions have a high electrical
conductivity.[71]
The aqueous solubility of a variety of ionic compounds as a function of
temperature. Some compounds exhibiting unusual solubility behaviour
have been included.
The solubility is highest in polar solvents (such as water) or ionic liquids,
but tends to be low in nonpolar solvents (such as petrol/gasoline).
[72] This is principally because the resulting ion–dipole interactions are
significantly stronger than ion-induced dipole interactions, so the heat of
solution is higher. When the oppositely charged ions in the solid ionic
lattice are surrounded by the opposite pole of a polar molecule, the solid
ions are pulled out of the lattice and into the liquid. If
the solvation energy exceeds the lattice energy, the negative
net enthalpy change of solution provides a thermodynamic drive to
remove ions from their positions in the crystal and dissolve in the liquid.
In addition, the entropy change of solution is usually positive for most
solid solutes like ionic compounds, which means that their solubility
increases when the temperature increases.[73] There are some unusual
ionic compounds such as cerium(III) sulfate, where this entropy change
is negative, due to extra order induced in the water upon solution, and
the solubility decreases with temperature.
Electrical conductivity
Although ionic compounds contain charged atoms or clusters, these
materials do not typically conduct electricity to any significant extent
when the substance is solid. In order to conduct, the charged particles
must be mobile rather than stationary in a crystal lattice. This is
achieved to some degree at high temperatures when the defect
concentration increases the ionic mobility and solid state ionic
conductivity is observed. When the ionic compounds are dissolved in a
liquid or are melted into a liquid, they can conduct electricity because
the ions become completely mobile.[74] This conductivity gain upon
dissolving or melting is sometimes used as a defining characteristic of
ionic compounds.
In some unusual ionic compounds: fast ion conductors, and ionic glasses,
[52] one or more of the ionic components has a significant mobility,
allowing conductivity even while the material as a whole remains solid.
[76] This is often highly temperature dependant, and may be the result
of either a phase change or a high defect concentration.[76] These
materials are used in all solid-state supercapacitors, batteries, and fuel
cells, and in various kinds of chemical sensors.
POOR CONDUCTORS OF HEAT AND ELECTRICITY
Colour
The colour of an ionic compound is often different to the colour of an
aqueous solution containing the constituent ions,[ or the hydrated form
of the same compound.
The anions in compounds with bonds with the most ionic character tend
to be colourless (with an absorption band in the ultraviolet part of the
spectrum). In compounds with less ionic character, their colour deepens
through yellow, orange, red and black (as the absorption band shifts to
longer wavelengths into the visible spectrum).
The absorption band of simple cations shift toward shorter wavelength
when they are involved in more covalent interactions.This occurs
during hydration of metal ions, so colourless anhydrous ionic compounds
with an anion absorbing in the infrared can become colourful in solution.
Ionization energy
The ionization energy (IE) is qualitatively defined as the amount of
energy required to remove the most loosely bound electron, the valence
electron, of an isolated gaseous atom to form a cation. It is
quantitatively expressed in symbols as
X + energy → X+ + e−
where X is any atom or molecule capable of being ionized, X+ is that
atom or molecule with an electron removed, and e− is the removed
electron. This is an endothermic process.
Generally, the closer the electrons are to the nucleus of the atom, the
higher the atom's ionization energy.
The units for ionization energy are different in physics and chemistry. In
physics, the unit is the amount of energy required to remove a single
electron from a single atom or molecule: expressed as an electron volt.
In chemistry, the units are the amount of energy it takes for all of the
atoms in a mole of substance to lose one electron each: molar ionization
energy or enthalpy, expressed as kilojoules per mole (kJ/mol)
or kilocalories per mole (kcal/mol).[1]
Comparison of IEs of atoms in the periodic table reveals two patterns:
IEs generally increase as one moves from left to right within a
given period.
IEs generally decrease as one moves down a given group.
Electron affinity
In chemistry and atomic physics, the electron affinity of
an atom or molecule is defined as the amount of
energy released or spent when an electron is added to a neutral atom or
molecule in the gaseous state to form a negative ion.[1]
X + e− → X− + energy
In solid state physics, the electron affinity for a surface is defined
somewhat differently (see below).
This property is measured for atoms and molecules in the gaseous state
only, since in a solid or liquid state their energy levels would be changed
by contact with other atoms or molecules. A list of the electron affinities
was used by Robert S. Mulliken to develop an electronegativity scale for
atoms, equal to the average of the electron affinity and ionization
potential. Other theoretical concepts that use electron affinity include
electronic chemical potential and chemical hardness. Another example,
a molecule or atom that has a more positive value of electron affinity
than another is often called an electron acceptor and the less positive
an electron donor. Together they may undergo charge-transfer reactions.
lthough Eea varies greatly across the periodic table, some patterns
emerge. Generally, nonmetals have more positive Eea than metals.
Atoms whose anions are more stable than neutral atoms have a
greater Eea. Chlorine most strongly attracts extra
electrons; mercury most weakly attracts an extra electron. The electron
affinities of the noble gases have not been conclusively measured, so
they may or may not have slightly negative values.
Eea generally increases across a period (row) in the periodic table. This
is caused by the filling of the valence shell of the atom; a Group 17 atom
releases more energy than a Group 1 atom on gaining an electron
because it obtains a filled valence shell and therefore is more stable.
A trend of decreasing Eea going down the groups in the periodic table
might be expected. The additional electron will be entering an orbital
farther away from the nucleus. Since this electron is farther from the
nucleus it is less attracted to the nucleus and would release less energy
when added. However, a clear counterexample to this trend can be
found in Group 2, and inspecting the entire periodic table, it turns out
that the proposed trend only applies to Group 1 atoms.
Thus, electron affinity follows the left-right trend of electronegativity but
not the up-down trend.
Metallic bonds
Metallic bonding arises from the electrostatic attractive force
between conduction electrons (in the form of an electron cloud of
delocalized electrons) and positively charged metal ions. It may be
described as the sharing of free electrons among a lattice of positively
charged ions (cations). Metallic bonding accounts for many physical
properties of metals, such as strength, ductility, thermal and electrical
resistivity and conductivity, opacity, and luster.
Metallic bonding is not the only type of chemical bonding a metal can
exhibit, even as a pure substance. For example,
elemental gallium consists of covalently-bound pairs of atoms in both
liquid and solid state—these pairs form a crystal lattice with metallic
bonding between them. Another example of a metal–metal covalent
bond is mercurous ion (Hg2+
2).

Metallic crystals
Metallic Properties
In a metal, atoms readily lose electrons to form positive ions (cations).
These ions are surrounded by delocalized electrons, which are
responsible for conductivity. The solid produced is held together by
electrostatic interactions between the ions and the electron cloud. These
interactions are called metallic bonds. Metallic bonding accounts for
many physical properties of metals, such as strength, malleability,
ductility, thermal and electrical conductivity, opacity, and luster.
Metallic Bonding
Loosely bound and mobile electrons surround the positive nuclei of metal
atoms.

Understood as the sharing of


"free" electrons among a lattice of
positively charged ions (cations),
metallic bonding is sometimes
compared to the bonding of
molten salts; however, this
simplistic view holds true for very
few metals. In a quantum-
mechanical view, the conducting
electrons spread
their density equally over all
atoms that function as neutral
(non-charged) entities.
Atoms in metals are arranged like closely-packed spheres, and two
packing patterns are particularly common: body-centered cubic, wherein
each metal is surrounded by eight equivalent metals, and face-centered
cubic, in which the metals are surrounded by six neighboring atoms.
Several metals adopt both structures, depending on the temperature.
Metals in general have high electrical conductivity, high thermal
conductivity, and high density. They typically are deformable (malleable)
under stress, without cleaving. Some metals (the alkali
and alkaline earth metals) have low density, low hardness, and
low melting points. In terms of optical properties, metals are opaque,
shiny, and lustrous.
Melting Point and Strength
The strength of a metal derives from the electrostatic attraction between
the lattice of positive ions and the "sea" of valence electrons in which
they are immersed. The larger the nuclear charge (atomic number) of
the atomic nucleus, and the smaller the atom's size, the greater this
attraction. In general, the transition metals with their valence-level d
electrons are stronger and have higher melting points:
Fe, 1539°C
Re, 3180 °C
Os, 2727 °C
W, 3380°C.
The majority of metals have higher densities than the majority of
nonmetals. Nonetheless, there is wide variation in the densities of
metals. Lithium (Li) is the least dense solid element, and osmium (Os) is
the densest. The metals of groups IA and IIA are referred to as the light
metals because they are exceptions to this generalization. The high
density of most metals is due to the tightly packed crystal lattice of the
metallic structure.
Electrical Conductivity: Why Are Metals Good Conductors?
In order for a substance to conduct electricity, it must contain charged
particles (charge carriers) that are sufficiently mobile to move in
response to an applied electric field. In the case of ionic compounds in
water solutions, the ions themselves serve this function. The same thing
holds true of ionic compounds when melted. Ionic solids contain the
same charge carriers, but because they are fixed in place, these solids
are insulators.
In metals, the charge carriers are the electrons, and because they move
freely through the lattice, metals are highly conductive. The very low
mass and inertia of the electrons allows them to conduct high-frequency
alternating currents, something that electrolytic solutions cannot do.
Electrical conductivity, as well as the electrons' contribution to the heat
capacity and heat conductivity of metals, can be calculated from
the free electron model, which does not take the detailed structure of
the ion lattice into account.
Mechanical properties
Mechanical properties of metals include malleability and ductility,
meaning the capacity for plastic deformation. Reversible elastic
deformation in metals can be described by Hooke's Law for restoring
forces, in which the stress is linearly proportional to the strain. Applied
heat, or forces larger than the elastic limit, may cause an irreversible
deformation of the object, known as plastic deformation or plasticity.
Metallic solids are known and valued for these qualities, which derive
from the non-directional nature of the attractions between the atomic
nuclei and the sea of electrons. The bonding within ionic or covalent
solids may be stronger, but it is also directional, making these solids
brittle and subject to fracture when struck with a hammer, for example.
A metal, by contrast, is more likely to be simply deformed or dented.
Although metals are black due to their ability to absorb
all wavelengths equally, gold (Au) has a distinctive color. According to
the theory of special relativity, increased mass of inner-shell electrons
that have very high momentum causes orbitals to contract. Because
outer electrons are less affected, blue-light absorption is increased,
resulting in enhanced reflection of yellow and red light.
Malleability
Malleability is a substance's ability to deform under pressure
(compressive stress). If malleable, a material may be flattened into thin
sheets by hammering or rolling. Malleable materials can be flattened
into metal leaf.
Malleability is a physical property of matter, usually metals. The property
usually applies to the family groups 1 to 12 on the modern periodic
table of elements. It is the ability of a solid to bend or be hammered into
other shapes without breaking. Examples of malleable metals
are gold, iron, aluminum, copper, silver, and lead.
Gold and silver are highly malleable. When a piece of hot iron is
hammered it takes the shape of a sheet. The property is not seen in non-
metals. Non-malleable metals may break apart when struck by a
hammer. Malleable metals usually bend and twist in various shapes.
Malleability and Ductility: The sea of electrons surrounding the protons
act like a cushion, and so when the metal is hammered on, for instance,
the over all composition of the structure of the metal is not harmed or
changed. The protons may be rearranged but the sea of electrons with
adjust to the new formation of protons and keep the metal intact.
Covalent bond
A covalent bond, also called a molecular bond, is a chemical bond that
involves the sharing of electron pairs between atoms. These electron
pairs are known as shared pairs or bonding pairs, and the stable balance
of attractive and repulsive forces between atoms, when they
share electrons, is known as covalent bonding. For many molecules, the
sharing of electrons allows each atom to attain the equivalent of a full
outer shell, corresponding to a stable electronic configuration.
Covalent bonding includes many kinds of interactions, including σ-
bonding, π-bonding, metal-to-metal bonding, agostic interactions, bent
bonds, and three-center two-electron bonds. The term covalent
bond dates from 1939.[4] The prefix co- means jointly, associated in
action, partnered to a lesser degree, etc.; thus a "co-valent bond", in
essence, means that the atoms share "valence", such as is discussed
in valence bond theory.
In the molecule H 2, the hydrogen atoms share the two electrons via
covalent bonding. Covalency is greatest between atoms of
similar electronegativities. Thus, covalent bonding does not necessarily
require that the two atoms be of the same elements, only that they be of
comparable electronegativity. Covalent bonding that entails sharing of
electrons over more than two atoms is said to be delocalized.

Covalent bonds are also affected by the electronegativity of the


connected atoms which determines the chemical polarity of the bond.
Two atoms with equal electronegativity will make nonpolar covalent
bonds such as H–H. An unequal relationship creates a polar covalent
bond such as with H−Cl. However polarity also requires geometric
asymmetry, or else dipoles may cancel out resulting in a non-polar
molecule.
Coordinate covalent bond
A coordinate covalent bond, also known as a dative bond or coordinate
bond is a kind of 2-center, 2-electron covalent bond in which the two
electrons derive from the same atom. The bonding of metal ions
to ligands involves this kind of interaction.
Coordinate covalent bonding is pervasive. In all metal aquo
complexes [M(H2O)n]x+, the bonding between water and the metal
cation is described as a coordinate covalent bond. Metal-ligand
interactions in most organometallic compounds and most coordination
compounds are described similarly.
The term dipolar bond is used in organic chemistry for compounds such
as amine oxides for which the electronic structure can be described in
terms of the basic amine donating two electrons to an oxygen atom.
R
3N → O
The arrow → indicates that both electrons in the bond originate from the
amine moiety. In a standard covalent bond each atom contributes one
electron. Therefore, an alternative description is that the amine gives
away one electron to the oxygen atom, which is then used, with the
remaining unpaired electron on the nitrogen atom, to form a standard
covalent bond. The process of transferring the electron from nitrogen to
oxygen creates formal charges, so the electronic structure may also be
depicted as
R
3N+
O−

Hexaamminecobalt(III) chloride
This electronic structure has an electric dipole, hence the name di-polar
bond. In reality, the atoms carry fractional charges; the
more electronegative atom of the two involved in the bond will usually
carry a fractional negative charge. One exception to this is carbon
monoxide. In this case, the carbon atom carries the fractional negative
charge despite its being less electronegative than oxygen.
An example of a dative covalent bond is provided by the interaction
between a molecule of ammonia, a Lewis base with a lone pair of
electrons on the nitrogen atom, and boron trifluoride, a Lewis acid by
virtue of the boron atom having an incomplete octet of electrons. In
forming the adduct, the boron atom attains an octet configuration.
The electronic structure of a coordination complex can be described in
terms of the set of ligands each donating a pair of electrons to a metal
centre. For example, in Hexaamminecobalt(III) chloride,
each ammonia ligand donates its lone pair of electrons to the cobalt(III)
ion. In this case, the bonds formed are described as coordinate bonds.
In all cases the bond is a covalent bond. The prefix dipolar, dative or
coordinate merely serves to indicate the origin of the electrons used in
creating the bond.
Covalent crystals
Covalent Network Solids
A covalent bond is a chemical bond that involves the sharing of pairs of
electrons between atoms. This sharing results in a stable balance of
attractive and repulsive forces between those atoms. Covalent solids are
a class of extended-lattice compounds in which each atom is covalently
bonded to its nearest neighbors. This means that the entire crystal is, in
effect, one giant molecule. The extraordinarily strong binding forces that
join all adjacent atoms account for the extreme hardness of these solids.
They cannot be broken or abraded without breaking a large number of
covalent chemical bonds. Similarly, a covalent solid cannot "melt" in the
usual sense, since the entire crystal is one giant molecule. When heated
to very high temperatures, these solids usually decompose into
their elements.
Another property of covalent network solids is poor electrical
conductivity, since there are no delocalized electrons. When molten,
unlike ionic compounds, the substance is still unable to conduct
electricity, since the macromolecule consists of uncharged atoms rather
than ions. (This is also contrary to most forms of metallic bonds. )
A Case Study: Allotropes of Carbon
Graphite is an allotrope of carbon. In this allotrope, each atom of carbon
forms three covalent bonds, leaving one electron in each
outer orbital delocalized, creating multiple "free electrons" within each
plane of carbon. This grants graphite electrical conductivity. Its melting
point is high, due to the large amount of energy required to rearrange
the covalent bonds. It is also quite hard because of the strong covalent
bonding throughout the lattice. However, because of the planar bonding
arrangements between the carbon atoms, the layers in graphite can be
easily displaced, so the substance is malleable. This explains the use of
graphite in pencils, where the layers of carbon are "shedded" on paper
(pencil "lead" is typically a mixture of graphite and clay, and was
invented for this use in 1795). Graphite is generally insoluble in
any solvent due to the difficulty of solvating a very large molecule.

Diamond and Graphite: Two Allotropes of Carbon


These two allotropes of carbon are covalent network solids which differ
in the bonding geometry of the carbon atoms. In diamond, the bonding
occurs in the tetrahedral geometry, while in graphite the carbons bond
with each other in the trigonal planar arrangement. This difference
accounts for the drastically different appearance and properties of these
two forms of carbon.

Diamond is also an allotrope of carbon. Diamond is the hardest material


known, defining the upper end of the 1-10 scale known as Moh's
hardness scale. Diamond cannot be melted; above 1700 °C it is
converted to graphite, the more stable form of carbon. The diamond unit
cell is face-centered cubic and contains eight carbon atoms.
Other Examples
Boron nitride (BN) is similar to carbon because it exists as a diamond-like
cubic polymorph as well as in a hexagonal form similar to graphite.
Hexagonal boron nitride
Hexagonal boron nitride, a two-dimensional material, is similar in
structure to graphite.

Cubic boron nitride is the second-hardest material after diamond, and it


is used in industrial abrasives and cutting tools.

Difference between covalent crystalls and moleculas crystalls


Crystalline solids contain atoms or molecules in a lattice display.
Covalent crystals, also known as network solids, and molecular crystals
represent two types of crystalline solids. Each solid exhibits different
properties but there is only one difference in their structure. That one
difference accounts for the different properties of the crystalline solids.
COVALENT BONDING
Covalent crystals exhibit covalent bonding; the principle that each atom
on the lattice is covalently bonded to every other atom. Covalent
bonding means the atoms have a strong attraction toward one another
and are held in place by that attraction. Network solids means the atoms
form a network with each atom connected to four other atoms. This
bonding in effect creates one large molecule that is tightly packed
together. This characteristic defines covalent crystals and makes them
structurally different from molecular crystals.
MOLECULAR BONDING
Molecular crystals contain either atoms or molecules, depending upon
the type of crystal, at each lattice site. They do not have covalent
bonding; the attraction is weak between the atoms or molecules. No
chemical bonds exist as in covalent crystals; electrostatic forces
between the atoms or molecules hold the molecular crystal together.
This difference causes molecular crystals to be loosely held together and
easily pulled apart.
EXAMPLES
Examples of covalent crystals include diamonds, quartz and silicon
carbide. All of these covalent crystals contain atoms that are tightly
packed and difficult to separate. Their structure varies widely from the
atoms in molecular crystals such as water and carbon dioxide which are
easily separated.
MELTING POINT
The differences in structure between covalent crystals and molecular
crystals cause the melting points of each type of crystal to differ.
Covalent crystals have high melting points while molecular crystals have
low melting points.

Electronegativity
Electronegativity, symbol χ, is a chemical property that describes the
tendency of an atom to attract electrons (or electron density) towards
itself.[1] An atom's electronegativity is affected by both its atomic
number and the distance at which its valence electrons reside from the
charged nucleus. The higher the associated electronegativity number,
the more an element or compound attracts electrons towards it.
Caesium is the least electronegative element in the periodic table
(=0.79), while fluorine is most electronegative (=3.98). (Francium and
caesium were originally both assigned 0.7; caesium's value was later
refined to 0.79, but no experimental data allows a similar refinement for
francium. However, francium's ionization energy is known to be slightly
higher than caesium's, in accordance with the relativistic stabilization of
the 7s orbital, and this in turn implies that francium is in fact more
electronegative than caesium.
In general, electronegativity increases on passing from left to right
along a period, and decreases on descending a group. Hence, fluorine is
the most electronegative of the elements (not counting noble gases),
whereas caesium is the least electronegative, at least of those elements
for which substantial data is available.This would lead one to believe
that caesium fluoride is the compound whose bonding features the most
ionic character.
Intermolecular forces
In physical chemistry, the van der Waals forces, named after Dutch
scientist Johannes Diderik van der Waals, are distance dependent
interactions between atoms. Unlike hydrogen or covalent bonds, these
attractions are not a result of any chemical electronic bond and so this
force is more susceptible to being perturbed. Being the weakest of the
weak chemical forces, with a strength between 0.4 and 4kJ/mol they
may still support an integral structural load when multitudes of such
interactions are present. Such a force results from a transient shift in
electron density. Specifically, as the electrons are in orbit of the protons
and neutrons within an atom the electron density may tend to shift more
greatly on a side. Thus, this generates a transient charge to which a
nearby atom can be either attracted or repelled. When the interatomic
distance of two atoms is greater than 0.6nm the force is not strong
enough to be observed. In the same vein, when the interatomic distance
is below 0.4nm the force becomes repulsive. If no other forces are
present, the point at which the force becomes repulsive rather than
attractive as two atoms near one another is called the Van der Waals
contact distance. This results from the electron clouds of two atoms
unfavorably coming into contact.[1] It can be shown that van der Waals
forces are of the same origin as the Casimir effect, arising
from quantum interactions with the zero-point field.[2] The resulting van
der Waals forces can be attractive or repulsive.[3]
The term includes:
force between permanent dipoles (Keesom force)
force between a permanent dipole and a corresponding induced dipole
(Debye force)
force between instantaneously induced dipoles (London dispersion
force).
It is also sometimes used loosely as a synonym for the totality
of intermolecular forces.[4] Van der Waals forces are relatively weak
compared to covalent bonds, but play a fundamental role in fields as
diverse as supramolecular chemistry, structural biology, polymer
science, nanotechnology, surface science, and condensed matter
physics. Van der Waals forces define many properties of organic
compounds and molecular solids, including their solubility in polar and
non-polar media.
In low molecular weight alcohols, the hydrogen-bonding properties of
their polar hydroxyl group dominate other weaker van der Waals
interactions. In higher molecular weight alcohols, the properties of the
nonpolar hydrocarbon chain(s) dominate and define the solubility. Van
der Waals forces quickly vanish at longer distances between interacting
molecules.
Hydrogen bond
A hydrogen bond is the electrostatic attraction between two polar groups
that occurs when a hydrogen (H) atom covalently bound to a
highly electronegative atom such as nitrogen (N), oxygen (O),
or fluorine (F) experiences the electrostatic field of another highly
electronegative atom nearby. It’s the strongest of the intermolecular
forces.
Hydrogen bonds can occur between molecules (intermolecular) or within
different parts of a single molecule (intramolecular).[4] Depending on
geometry and environment, the hydrogen bond free energy content is
between 1 and 5 kcal/mol. This makes it stronger than a van der Waals
interaction, but weaker than covalent or ionic bonds. This type of bond
can occur in inorganic molecules such as water and in organic
molecules like DNA and proteins.
Intermolecular hydrogen bonding is responsible for the high boiling point
of water (100 °C) compared to the other group 16 hydrides that have
much weaker hydrogen bonds.[5] Intramolecular hydrogen bonding is
partly responsible for the secondary and tertiary structures
of proteins and nucleic acids. It also plays an important role in the
structure of polymers, both synthetic and natural.
In 2011, an IUPAC Task Group recommended a modern evidence-based
definition of hydrogen bonding, which was published in
the IUPAC journal Pure and Applied Chemistry. This definition specifies:
The hydrogen bond is an attractive interaction between a hydrogen
atom from a molecule or a molecular fragment X–H in which X is more
electronegative than H, and an atom or a group of atoms in the same or
a different molecule, in which there is evidence of bond formation.
The most ubiquitous and perhaps simplest example of a hydrogen bond
is found between water molecules. In a discrete water molecule, there
are two hydrogen atoms and one oxygen atom. Two molecules
of water can form a hydrogen bond between them; the simplest case,
when only two molecules are present, is called the water dimer and is
often used as a model system. When more molecules are present, as is
the case with liquid water, more bonds are possible because the oxygen
of one water molecule has two lone pairs of electrons, each of which can
form a hydrogen bond with a hydrogen on another water molecule. This
can repeat such that every water molecule is H-bonded with up to four
other molecules, as shown in the figure (two through its two lone pairs,
and two through its two hydrogen atoms). Hydrogen bonding strongly
affects the crystal structure of ice, helping to create an open hexagonal
lattice. The density of ice is less than the density of water at the same
temperature; thus, the solid phase of water floats on the liquid, unlike
most other substances.
Liquid water's high boiling point is due to the high number of hydrogen
bonds each molecule can form, relative to its low molecular mass. Owing
to the difficulty of breaking these bonds, water has a very high boiling
point, melting point, and viscosity compared to otherwise similar liquids
not conjoined by hydrogen bonds.
Chemical bonds and properties of substances
Comparison of Properties of Ionic and Covalent Compounds
Because of the nature of ionic and covalent bonds, the materials
produced by those bonds tend to have quite different macroscopic
properties. The atoms of covalent materials are bound tightly to each
other in stable molecules, but those molecules are generally not very
strongly attracted to other molecules in the material. On the other hand,
the atoms (ions) in ionic materials show strong attractions to other ions
in their vicinity. This generally leads to low melting points for covalent
solids, and high melting points for ionic solids. For example, the
molecule carbon tetrachloride is a non-polar covalent molecule, CCl4. It's
melting point is -23°C. By contrast, the ionic solid NaCl has a melting
point of 800°C.
Ionic Compounds Covalent Compounds

Crystalline solids (made of ions) Gases, liquids, or solids (made of


molecules)
High melting and boiling points
Low melting and boiling points
Conduct electricity when melted
Poor electrical conductors in all
phases
Many soluble in water but not in
nonpolar liquid Many soluble in nonpolar liquids but
not in water
You can anticipate some things about bonds from the positions of the
constituents in the periodic table. Elements from opposite ends of the
periodic table will generally form ionic bonds. They will have large
differences in electronegativity and will usually form positive and
negative ions. The elements with the largest electronegativities are in
the upper right of the periodic table, and the elements with the smallest
electronegativities are on the bottom left. If these extremes are
combined, such as in RbF, the dissociation energy is large. As can be
seen from the illustration below, hydrogen is the exception to that rule,
forming covalent bonds.
Elements which are close together in electronegativity tend to form
covalent bonds and can exist as stable free molecules. Carbon dioxide is
a common example.

General Properties of Covalent Compounds that Exhibit


Hydrogen Bonding
• Lower Melting Points than ionic compounds, but higher than nonpolar
or polar covalent compounds – Intermolecular Forces aren’t as strong as
interionic forces, but Hydrogen bonding is stronger than other
intermolecular forces • Cannot conduct electricity as solid, liquid, or
when dissolved – Still no ions • Will dissolve in liquids with similar
molecular polarity.
TIP
Classifying compounds as ionic or covalent
Compounds containing two elements (so called binary compounds) can
either have ionic or covalent bonding. If a compound is made from a
metal and a non-metal, its bonding will be ionic. If a compound is made
from two non-metals, its bonding will be covalent.
Amount of substance
Atomic mass
Relative atomic mass (symbol: Ar) is a dimensionless (number
only) physical quantity. In its modern definition, it is the ratio of the
average mass of atoms of an element in a given sample to one unified
atomic mass unit. The unified atomic mass unit, symbol u, is defined
being  1⁄12 of the mass of a carbon-12 atom.[2][3] The mass of atoms
can vary (between atoms of the same element), due to the presence of
various isotopes of that element. Since both values in the ratio are
expressed in the same unit (u), the resulting value is dimensionless;
hence the value is relative.

Molecular weight
Molecular mass or molecular weight is the mass of a molecule. It is
calculated as the sum of the atomic weights of each
constituent element multiplied by the number of atoms of that element
in the molecular formula.
Both atomic and molecular masses are usually obtained relative to the
mass of the isotope 12C (carbon 12), which by definition[1] is equal to
12. For example, the molecular weight of methane, whose molecular
formula is CH4, is calculated as follows:

Atomic Total
mass mass

C 12.011 12.011

H 1.00794 4.03176

CH
16.043
4

Formula weight: The mass in grams of a mole of a compound. Equals to


the sum of the masses of the atoms in the compound.
Molarity
There are two types of percent
concentration: percent by mass and percent by volume.
PERCENT BY MASS
Percent by mass (m/m) is the mass of solute divided by the total mass of
the solution, multiplied by 100 %.
Percent by mass = mass of solutetotal mass of solution × 100 %
Example
What is the percent by mass of a solution that contains 26.5 g of glucose
in 500 g of solution?
Solution
Percent by mass =
mass of glucosetotal mass of solution×100%=26.5g500g × 100 % =
5.30 %
PERCENT BY VOLUME
Percent by volume (v/v) is the volume of solute divided by the total
volume of the solution, multiplied by 100 %.
Percent by volume = volume of solutetotal volume of solution × 100 %
Example
How would you prepare 250 mL of 70 % (v/v) of rubbing alcohol
Solution
70 % = volume of rubbing alcoholtotal volume of solution×100% × 100
%
So
Volume of rubbing alcohol = volume of solution × 70 %100 % = 250 mL
× 70100
= 175 mL
You would add enough water to 175 mL of rubbing alcohol to make a
total of 250 mL of solution.
Chemical formulas
molecular formula: a chemical formula that gives the total number of
atoms of each element in each molecule of a substance

Ions

https://www.youtube.com/watch?v=bPoxAdcYIHU

Formulas for Ionic Compounds

The atoms in molecules bond to one another through "sharing"of


electrons. Ionic compounds on the other had have atoms or molecules
that bond to one another through their mutual attraction between
positive and negative charges:

Positively charged atoms or molecules are called "cations" and


negatively charged atoms or molecules are called "anions". Thus cations
and anions attract one another. Conversely, cations repel one another as
do anions.
Electrostatic attraction is indiscriminate. That is a cation can attract
more than one anion and visa versa. The result is that cation-anions
attractions form a large array that we call an ionic compound or "salt".
The bonds holding these ions together are called "ionic" bonds.
However, this array has a very specific composition completely dictated
by the charges on the cations and anions.
Composition of ionic compounds
The composition of ionic compounds is determined by the requirement
that the compounds must be electrically neutral. That is that the charges
of the cations and anions must balance or 'cancel" out one another. For
example consider sodium cations (Na+) and Chlorine anions (Cl-).
Sodium has a positive 1 charge and chloride has a negative 1 charge.
Thus one sodium cation cancels one chloride anion resulting in the
formula Na1Cl1 or NaCl. This formula is called the "formula unit" since it
represents only one unit of the vast NaCl array or lattice. Below are other
examples of simple ionic compounds made up of single atom
(monatomic) ions:
Polyatomic ions and compounds
The ionic compounds above were made up of monatomic ions. Molecules
can also be ions - polyatomic ions. Most polyatomic ions are anions with
one notable exception - the ammoniun cation (NH4+). The composition
of salts with polyatomic ions is determined by the same rule as with
monatomic ions. A few example are given below:

Note that when there is two or more polyatomic ions in the formula the
ion is enclosed with parenthesis "()".
Lewis structures
Writing Lewis Structures by Trial and Error
The Lewis structure of a compound can be generated by trial and error.
We start by writing symbols that contain the correct number of valence
electrons for the atoms in the molecule. We then combine electrons to
form covalent bonds until we come up with a Lewis structure in which all
of the elements (with the exception of the hydrogen atoms) have an
octet of valence electrons.
Example: Let's apply the trial and error approach to generating the Lewis
structure of carbon dioxide, CO2. We start by determining the number of
valence electrons on each atom from the electron configurations of the
elements. Carbon has four valence electrons, and oxygen has six.
C: [He] 2s2 2p2
O: [He] 2s2 2p4
We can symbolize this information as shown at the top of the figure
below. We now combine one electron from each atom to form covalent
bonds between the atoms. When this is done, each oxygen atom has a
total of seven valence electrons and the carbon atom has a total of six
valence electrons. Because none of these atoms have an octet of
valence electrons, we combine another electron on each atom to form
two more bonds. The result is a Lewis structure in which each atom has
an octet of valence electrons.

A Step-By-Step Approach To Writing Lewis Structures


The trial-and-error method for writing Lewis structures can be time
consuming. For all but the simplest molecules, the following step-by-step
process is faster.
Step 1: Determine the total number of valence electrons.
Step 2: Write the skeleton structure of the molecule.
Step 3: Use two valence electrons to form each bond in the skeleton
structure.
Step 4: Try to satisfy the octets of the atoms by distributing the
remaining valence electrons as nonbonding electrons.
The first step in this process involves calculating the number of valence
electrons in the molecule or ion. For a neutral molecule this is nothing
more than the sum of the valence electrons on each atom. If the
molecule carries an electric charge, we add one electron for each
negative charge or subtract an electron for each positive charge.
Example: Let's determine the number of valence electrons in the
chlorate (ClO3-) ion.
A chlorine atom (Group VIIA) has seven valence electrons and each
oxygen atom (Group VIA) has six valence electrons. Because the
chlorate ion has a charge of -1, this ion contains one more electron than
a neutral ClO3 molecule. Thus, the ClO3- ion has a total of 26 valence
electrons.
ClO3-: 7 + 3(6) + 1 = 26
The second step in this process involves deciding which atoms in the
molecule are connected by covalent bonds. The formula of the
compound often provides a hint as to the skeleton structure. The formula
for the chlorate ion, for example, suggests the following skeleton
structure.

The third step assumes that the skeleton structure of the molecule is
held together by covalent bonds. The valence electrons are therefore
divided into two categories: bonding electrons and nonbonding
electrons. Because it takes two electrons to form a covalent bond, we
can calculate the number of nonbonding electrons in the molecule by
subtracting two electrons from the total number of valence electrons for
each bond in the skeleton structure.
There are three covalent bonds in the most reasonable skeleton
structure for the chlorate ion. As a result, six of the 26 valence electrons
must be used as bonding electrons. This leaves 20 nonbonding electrons
in the valence shell.

26 valence electrons

- 6 bonding electrons

�������������������

20 nonbonding electrons
The nonbonding valence electrons are now used to satisfy the octets of
the atoms in the molecule. Each oxygen atom in the ClO3- ion already
has two electrons the electrons in the Cl-O covalent bond. Because
each oxygen atom needs six nonbonding electrons to satisfy its octet, it
takes 18 nonbonding electrons to satisfy the three oxygen atoms. This
leaves one pair of nonbonding electrons, which can be used to fill the
octet of the central atom.

Drawing Skeleton Structures


The most difficult part of the four-step process in the previous section is
writing the skeleton structure of the molecule. As a general rule, the less
electronegative element is at the center of the molecule.
Example: The formulas of thionyl chloride (SOCl2) and sulfuryl chloride
(SO2Cl2) can be translated into the following skeleton structures.

It is also useful to recognize that the formulas for complex molecules are
often written in a way that hints at the skeleton structure of the
molecule.
Example: Dimethyl ether is often written as CH3OCH3, which translates
into the following skeleton structure.

Finally, it is useful to recognize that many compounds that are acids


contain O-H bonds.
Example: The formula of acetic acid is often written as CH3CO2H,
because this molecule contains the following skeleton structure.

Molecules that Contain Too Many or Not Enough Electrons


Too Few Electrons
Occasionally we encounter a molecule that doesn't seem to have
enough valence electrons. If we can't get a satisfactory Lewis structure
by sharing a single pair of electrons, it may be possible to achieve this
goal by sharing two or even three pairs of electrons.
Example: Consider formaldehyde (H2CO) which contains 12 valence
electrons.
H2CO: 2(1) + 4 + 6 = 12
The formula of this molecule suggests the following skeleton structure.

There are three covalent bonds in this skeleton structure, which means
that six valence electrons must be used as bonding electrons. This
leaves six nonbonding electrons. It is impossible, however, to satisfy the
octets of the atoms in this molecule with only six nonbonding electrons.
When the nonbonding electrons are used to satisfy the octet of the
oxygen atom, the carbon atom has a total of only six valence electrons.
We therefore assume that the carbon and oxygen atoms share two pairs
of electrons. There are now four bonds in the skeleton structure, which
leaves only four nonbonding electrons. This is enough, however, to
satisfy the octets of the carbon and oxygen atoms.

Every once in a while, we encounter a molecule for which it is impossible


to write a satisfactory Lewis structure.
Example: Consider boron trifluoride (BF3) which contains 24 valence
electrons.
BF3: 3 + 3(7) = 24
There are three covalent bonds in the most reasonable skeleton
structure for the molecule. Because it takes six electrons to form the
skeleton structure, there are 18 nonbonding valence electrons. Each
fluorine atom needs six nonbonding electrons to satisfy its octet. Thus,
all of the nonbonding electrons are consumed by the three fluorine
atoms. As a result, we run out of electrons while the boron atom has only
six valence electrons.
The elements that form strong double or triple bonds are C, N, O, P, and
S. Because neither boron nor fluorine falls in this category, we have to
stop with what appears to be an unsatisfactory Lewis structure.
Too Many Electrons
It is also possible to encounter a molecule that seems to have too many
valence electrons. When that happens, we expand the valence shell of
the central atom.
Example: Consider the Lewis structure for sulfur tetrafluoride (SF4)
which contains 34 valence electrons.
SF4: 6 + 4(7) = 34
There are four covalent bonds in the skeleton structure for SF4. Because
this requires using eight valence electrons to form the covalent bonds
that hold the molecule together, there are 26 nonbonding valence
electrons.
Each fluorine atom needs six nonbonding electrons to satisfy its octet.
Because there are four of these atoms, so we need 24 nonbonding
electrons for this purpose. But there are 26 nonbonding electrons in this
molecule. We have already satisfied the octets for all five atoms, and we
still have one more pair of valence electrons. We therefore expand the
valence shell of the sulfur atom to hold more than eight electrons.

This raises an interesting question: How does the sulfur atom in SF4 hold
10 electrons in its valence shell? The electron configuration for a neutral
sulfur atom seems to suggest that it takes eight electrons to fill the
3s and 3p orbitals in the valence shell of this atom. But let's look, once
again, at the selection rules for atomic orbitals. According to these rules,
the n = 3 shell of orbitals contains 3s, 3p, and 3dorbitals. Because the
3d orbitals on a neutral sulfur atom are all empty, one of these orbitals
can be used to hold the extra pair of electrons on the sulfur atom in SF4.
S: [Ne] 3s2 3p4 3d0
Resonance Hybrids
Two Lewis structures can be written for sulfur dioxide.
The only difference between these Lewis structures is the identity of the
oxygen atom to which the double bond is formed. As a result, they must
be equally satisfactory representations of the molecule.
Interestingly enough, neither of these structures is correct. The two
Lewis structures suggest that one of the sulfur-oxygen bonds is stronger
than the other. There is no difference between the length of the two
bonds in SO2, however, which suggests that the two sulfur-oxygen
bonds are equally strong.
When we can write more than one satisfactory Lewis structure, the
molecule is an average, or resonance hybrid, of these structures. The
meaning of the term resonance can be best understood by an analogy. In
music, the notes in a chord are often said to resonate they mix to
give something that is more than the sum of its parts. In a similar sense,
the two Lewis structures for the SO2 molecule are in resonance. They
mix to give a hybrid that is more than the sum of its components. The
fact that SO2 is a resonance hybrid of two Lewis structures is indicated
by writing a double-headed arrow between these Lewis structures, as
shown in the figure above.
Molecular Formula: The actual formula for a molecule.
Problem:
A compound is 75.46% carbon, 4.43% hydrogen, and 20.10% oxygen by
mass. It has a molecular weight of 318.31 g/mol. What is the molecular
formula for this compound?
Strategy:
Find the empirical formula.
Get the mass of each element by assuming a certain overall mass for
the sample (100 g is a good mass to assume when working with
percentages).
(.7546) (100 g) = 75.46 g C
(.0443) (100 g) = 4.43 g H
(.2010) (100 g) = 20.10 g O
Convert the mass of each element to moles.
(75.46 g C) (1 mol/ 12.00 g C) = 6.289 mol C
(4.43 g H) (1 mol/ 1.008 g H) = 4.39 mol H
(20.10 g O) (1 mol/ 16.00 g O) = 1.256 mol O
Find the ratio of the moles of each element.
(1.256 mol O)/ (1.256) = 1 mol O
(6.289 mol C)/ (1.256) = 5.007 mol C
(4.39 mol H)/ (1.256) = 3.50 mol H
Use the mole ratio to write the empirical fomula.
Multiplying the mole ratios by two to get whole number, the empirical
formula becomes:
C10H7O2
Find the mass of the empirical unit.
10(12.00) + 7(1.008) + 2(16.00) = 159.06 g/mol
Figure out how many empirical units are in a molecular unit.
(318.31 g/mol) / (159.06 g/mol) = 2.001 empirical units per molecular
unit
Write the molecular formula.

Since there are two empirical units in a molecular unit, the


molecular formula is:
C20H14O4

Das könnte Ihnen auch gefallen