Sunteți pe pagina 1din 58

CHAPTER 1: INTERMOLECULAR FORCES AND LIQUIDS AND SOLIDS

Kinetic Molecular Theory

Solids have strong forces of attraction. The Kinetic energy is not sufficient to overcome
the attractive forces. There are consequences of the orderly arrangements of solids. The particles
are not free to move and their movement is limited to vibration. Since they are held in fixed
positions, they occupy a small volume and thus, have small density. Also, since they are not free
to move, diffusion is slow and can only be compressed very slightly.

Liquid particles have sufficient kinetic energy to overcome their attractive forces. Thus,
the particles can move at short distances and collide with each other. But liquid particles do not
move away independent of each other. Liquids, therefore, have fixed volume but do not have fixed
shape. Diffusion of liquids is more rapid than those of solids because of the closely packed particles
in solids. Since the particles are quite close, compressibility is also very slight but greater than that
of solids.

Intermolecular Forces

Water is the only substance we routinely encounter as a solid, a liquid, and a gas. At low
temperatures, it is a solid in which the individual molecules are locked into a rigid structure. As
we raise the temperature, the average kinetic energy of the molecules increases, which increases
the rate at which these molecules move.

There are three ways in which a water molecule move: (1) vibration, (2) rotation, and (3)
translation. Water molecules vibrate when H--O bonds are stretched or bent. Rotation involves the
motion of a molecule around its center of gravity. Translation literally means to change from one
place to another. It therefore describes the motion of molecules through space.
The attractive forces of solids, liquids, and gases differ from one another as discussed by
the Kinetic Molecular Theory of Matter. Among the three states, solids have the strongest forces.
The forces of attraction for gases are very weak and therefore considered negligible. What is the
nature of these forces in solids and liquids? The particles of solids and liquids are significantly
attracted to each other and therefore are close to each other. Because of the proximity of the
particles, the particles occupy a significant volume of the substance. Also, the particles are not in
random motion. The particles interact with one another as they move, creating intermolecular
forces.

To understand the effect of this motion, we need to differentiate


between intramolecular and intermolecular bonds. The covalent bonds between the hydrogen and
oxygen atoms in a water molecule are called intramolecular bonds. (The prefix intra- comes
from the Latin stem meaning "within or inside." Thus, intramural sports match teams from the
same institution.) The bonds between the neighboring water molecules in ice are
called intermolecular bonds, from the Latin stem meaning "between." (This far more common
prefix is used in words such as interface, intercollegiate, and international.)

The intramolecular bonds that hold the atoms in H2O molecules together are almost 25
times as strong as the intermolecular bonds between water molecules. (It takes 464 kJ/mol to break
the H--O bonds within a water molecule and only 19 kJ/mol to break the bonds between water
molecules.)

All three modes of motion disrupt the bonds between water molecules. As the system
becomes warmer, the thermal energy of the water molecules eventually becomes too large to allow
these molecules to be locked into the rigid structure of ice. At this point, the solid melts to form a
liquid in which intermolecular bonds are constantly broken and reformed as the molecules move
through the liquid. Eventually, the thermal energy of the water molecules becomes so large that
they move too rapidly to form intermolecular bonds and the liquid boils to form a gas in which
each particle moves more or less randomly through space.

The difference between solids and liquids, or liquids and gases, is therefore based on a
competition between the strength of intermolecular bonds and the thermal energy of the system.
At a given temperature, substances that contain strong intermolecular bonds are more likely to be
solids. For a given intermolecular bond strength, the higher the temperature, the more likely the
substance will be a gas.

Dipole-Dipole Forces
Recall that covalent compounds can be classified as nonpolar or polar. A nonpolar
molecule results when the electronegativity difference between two atoms is less than 0.4 and
the molecule is symmetrical in shape. On the other hand, a polar molecule is formed when
there is an uneven sharing of electrons between two atoms. In this case, the electronegativity
difference is greater than 0.4 and the molecule is unsymmetrical. The more electronegative
atom becomes partially negative and the less electronegativity becomes partially positive.
The word partial means that the charge is somewhere between 0 and 1. This results in the
formation of a permanent dipole. The partial positive charge on one end of the molecule
becomes attracted to the partial negative end of another molecule. Dipole-dipole forces,
therefore, are formed between neighboring molecules with permanent dipoles. The dipole-
dipole forces are strong because of the attraction of opposite charges that are permanent
within the entire substance.

Many molecules contain bonds that fall between the extremes of ionic and covalent bonds.
The difference between the electronegativities of the atoms in these molecules is large enough that
the electrons aren't shared equally, and yet small enough that the electrons aren't drawn exclusively
to one of the atoms to form positive and negative ions. The bonds in these molecules are said to
be polar, because they have positive and negative ends, or poles, and the molecules are often said
to have a dipole moment.

HCl molecules, for example, have a dipole moment because the hydrogen atom has a slight
positive charge and the chlorine atom has a slight negative charge. Because of the force of
attraction between oppositely charged particles, there is a small dipole-dipole force of attraction
between adjacent HCl molecules.

The dipole-dipole interaction in HCl is relatively weak; only 3.3 kJ/mol. (The covalent
bonds between the hydrogen and chlorine atoms in HCl are 130 times as strong.) The force of
attraction between HCl molecules is so small that hydrogen chloride boils at -85.0oC.

Ion-dipole Forces

An ion-dipole force is an attractive force that results from the electrostatic attraction between
an ion and a neutral molecule that has a dipole.
• Most commonly found in solutions. Especially important for solutions of ionic compounds
in polar liquids.
• A positive ion (cation) attracts the partially negative end of a neutral polar molecule.
• A negative ion (anion) attracts the partially positive end of a neutral polar molecule.

• Ion-dipole attractions become stronger as either the charge on the ion increases, or as the
magnitude of the dipole of the polar molecule increases.
• Ion-dipole and ion-induced dipole forces operate much like dipole-dipole and induced
dipole-dipole interactions. However, ion-dipole forces involve ions instead of solely polar
molecules. Ion-dipole forces are stronger than dipole interactions because the charge of
any ion is much greater than the charge of a dipole; the strength of the ion-dipole force is
proportionate to ion charge. Ion-dipole bonding is also stronger than hydrogen bonding.
An ion-dipole force consists of an ion and a polar molecule aligning so that the positive
and negative charges are next to one another, allowing for maximum attraction.
• Ion-dipole forces are generated between polar water molecules and a sodium ion. The
oxygen atom in the water molecule has a slight negative charge and is attracted to the
positive sodium ion. These intermolecular ion-dipole forces are much weaker than covalent
or ionic bonds.

London Dispersion Forces

Although charges are usually distributed evenly between atoms in non-polar molecules,
spontaneous dipoles can still occur. When this occurs, non-polar molecules form weak attractions
with other non-polar molecules. These London dispersion forces are often found in the halogens
(e.g., F2 and I2), the noble gases (e.g., Ne and Ar), and in other non-polar molecules, such as carbon
dioxide and methane. London dispersion forces are part of the van der Waals forces, or weak
intermolecular attractions.

Van der Waals forces help explain how nitrogen can be liquefied. Nitrogen gas (N2) is
diatomic and non-polar because both nitrogen atoms have the same degree of electronegativity. If
there are no dipoles, what would make the nitrogen atoms stick together to form a liquid? London
dispersion forces allow otherwise non-polar molecules to have attractive forces. However, they
are by far the weakest forces that hold molecules together.

Liquid nitrogen: Without London dispersion forces, diatomic nitrogen would not remain liquid.

Hydrogen Bonding

A hydrogen bond is a strong intermolecular force created by the relative positivity of


hydrogen atoms.

Forming a Hydrogen Bond

A hydrogen bond is the electromagnetic attraction created between a partially positively


charged hydrogen atom attached to a highly electronegative atom and another nearby
electronegative atom. A hydrogen bond is a type of dipole-dipole interaction; it is not a true
chemical bond. These attractions can occur between molecules (intermolecularly) or within
different parts of a single molecule (intramolecularly).
Hydrogen Bond Donor

A hydrogen atom attached to a relatively electronegative atom is a hydrogen bond donor.


This electronegative atom is usually fluorine, oxygen, or nitrogen. The electronegative atom
attracts the electron cloud from around the hydrogen nucleus and, by decentralizing the cloud,
leaves the hydrogen atom with a positive partial charge. Because of the small size of hydrogen
relative to other atoms and molecules, the resulting charge, though only partial, is stronger. In the
molecule ethanol, there is one hydrogen atom bonded to an oxygen atom, which is very
electronegative. This hydrogen atom is a hydrogen bond donor.

Hydrogen Bond Acceptor

A hydrogen bond results when this strong partial positive charge attracts a lone pair of
electrons on another atom, which becomes the hydrogen bond acceptor. An electronegative atom
such as fluorine, oxygen, or nitrogen is a hydrogen bond acceptor, regardless of whether it is
bonded to a hydrogen atom or not. Greater electronegativity of the hydrogen bond acceptor will
create a stronger hydrogen bond. The diethyl ether molecule contains an oxygen atom that is not
bonded to a hydrogen atom, making it a hydrogen bond acceptor.

A hydrogen attached to carbon can also participate in hydrogen bonding when the carbon
atom is bound to electronegative atoms, as is the case in chloroform (CHCl3). As in a molecule
where a hydrogen is attached to nitrogen, oxygen, or fluorine, the electronegative atom attracts the
electron cloud from around the hydrogen nucleus and, by decentralizing the cloud, leaves the
hydrogen atom with a positive partial charge.

Applications for Hydrogen Bonds

Hydrogen bonds occur in inorganic molecules, such as water, and organic molecules, such
as DNA and proteins. The two complementary strands of DNA are held together by hydrogen
bonds between complementary nucleotides (A&T, C&G). Hydrogen bonding in water contributes
to its unique properties, including its high boiling point (100 °C) and surface tension.

In biology, intramolecular hydrogen bonding is partly responsible for the secondary,


tertiary, and quaternary structures of proteins and nucleic acids. The hydrogen bonds help the
proteins and nucleic acids form and maintain specific shapes.

Some Properties of Liquids


Surface tension, capillary action, and viscosity are unique properties of liquids that depend
on the nature of intermolecular interactions. Surface tension is the energy required to increase the
surface area of a liquid. Surfactants are molecules that reduce the surface tension of polar liquids
like water. Capillary action is the phenomenon in which liquids rise up into a narrow tube called
a capillary. The viscosity of a liquid is its resistance to flow.
• two properties of liquids: viscosity and surface tension

1. Viscosity
• viscosity – resistance of a liquid to flow
• the greater the viscosity the more slowly the liquid flows
• measured by timing how long it takes a certain amount of liquid to flow through a thin tube
under gravitational forces
• can also be measured by how long it takes steel spheres to fall through the liquid
• viscosity related to ease with which individual molecules of liquid can move with respect
to one another
• depends on attractive forces between molecules, and whether structural features exist to
cause molecules to be entangled
• viscosity decreases with increasing temperature
2. Surface Tension
• surface tension – energy required to increase the surface area of a liquid by a unit amount
• surface tension of water at 20° C is 7.29 x 10-2 J/m2
• 7.29 x 10-2 J/m2 must be supplied to increase surface area of a given amount of water by 1
m2
• cohesive forces – intermolecular forces that bind similar molecules
• adhesive forces – intermolecular forces that bind a substance to a surface
• capillary action – rise of liquids up very narrow tubes

Phase Changes
Fusion, vaporization, and sublimation are endothermic processes, whereas freezing,
condensation, and deposition are exothermic processes. Changes of state are examples of phase
changes, or phase transitions. All phase changes are accompanied by changes in the energy of a
system. Changes from a more-ordered state to a less-ordered state (such as a liquid to a gas) are
endothermic. Changes from a less-ordered state to a more-ordered state (such as a liquid to a
solid) are always exothermic.
1. Energy Changes Accompanying Phase Changes
o phase changes require energy
o phase changes to less ordered state requires energy
o melting process of solid called fusion
o heat of fusion – enthalpy change of melting a solid
o D Hfus water = 6.01 kJ/mol
o heat of vaporization – heat needed for vaporization of liquid
o D Hvap water = 40.67 kJ/mol
o melting, vaporization, and sublimation are endothermic
o freezing, condensation, and deposition are exothermic
2. Heating Curves
o heating curve – graph of temperature of system versus the amount of heat added
o used to calculate enthalpy changes
o supercooled water – when water if cooled to a temperature below 0° C
3. Critical Temperature and Pressure
o critical temperature – highest temperature at which a substance can exist as a liquid
o critical pressure – pressure required to bring about liquefaction at critical temperature
o the greater the intermolecular attractive forces, the more readily gases liquefy ® higher
critical temperature
o cannot liquefy a gas by applying pressure if gas is above critical temperature

Vapor Pressure
Because the molecules of a liquid are in constant motion and possess a wide range of
kinetic energies, at any moment some fraction of them has enough energy to escape from the
surface of the liquid to enter the gas or vapor phase. This process, called vaporization or
evaporation, generates a vapor pressure above the liquid. Molecules in the gas phase can collide
with the liquid surface and reenter the liquid via condensation. Eventually, a steady state or
dynamic equilibrium is reached.
vapor pressure – measures tendency of a liquid to evaporate
1. Explaining Vapor Pressure on the Molecular Level
o dynamic equilibrium – condition when two opposing processes are occurring
simultaneously at equal rates
o vapor pressure of a liquid is the pressure exerted by its vapor when the liquid and vapor
states are in dynamic equilibrium
2. Volatility, Vapor Pressure, and Temperature
o volatile – liquids that evaporate readily
o vapor pressure increases with increasing temperature
3. Vapor Pressure and Boiling Point
o liquids boil when its vapor pressure equals the external pressure acting on the surface of
the liquid
o temperature of boiling increase with increasing external pressure
o normal boiling point – boiling point of a liquid at 1 atm
o higher pressures cause water to boil at higher temperatures

Structure of Solids
A crystalline solid can be represented by its unit cell, which is the smallest identical unit
that when stacked together produces the characteristic three-dimensional structure. Solids are
characterized by an extended three-dimensional arrangement of atoms, ions, or molecules in
which the components are generally locked into their positions. The components can be arranged
in a regular repeating three-dimensional array. The smallest repeating unit of a crystal lattice is
the unit cell.
• crystalline solid – solid whose atoms, ion, or molecules are ordered in well-defined
arrangements
o flat surfaces or faces that make definite angles
o regular shapes
• amorphous solid – solid whose particles have no orderly structure
o lack well-defined faces and shapes
o mixtures of molecules that do not stack together well
o large, complicated molecules
o intermolecular forces vary in strength
o does not melt at a specific temperature but soften over a temperature range
1. Unit Cell

o unit cell – repeating unit of a solid
o crystal lattice – three-dimensional array of points, each representing an identical
environment within the crystal
o three types of cubic unit cell: primitive cubic, body-centered cubic, and face-centered cubic
o primitive cubic – lattice points at corners only
o body-centered cubic – lattice points at corners and center
o face-centered cubic – lattice points at center of each face and at each corner
2. The Crystal structure of Sodium Chloride

o total cation-to-anion ratio of a unit cell must be the same as that for entire crystal
3. Close Packing of Spheres

o structures of crystalline solids are those that bring particles in closest contact to maximize
the attractive forces
o most particles that make up solids are spherical
o two forms of close packing: cubic close packing and hexagonal close packing
o hexagonal close packing – spheres of the third layer that are placed in line with those of
the first layer
o coordination number – number of particles immediately surrounding a particle in the
crystal structure
o both forms of close packing have coordination number of 12
Bonding of Solids
The major types of solids are ionic, molecular, covalent, and metallic. Ionic solids consist
of positively and negatively charged ions held together by electrostatic forces; the strength of the
bonding is reflected in the lattice energy. Ionic solids tend to have high melting points and are
rather hard. Molecular solids are held together by relatively weak forces, such as dipole–dipole
interactions, hydrogen bonds, and London dispersion forces. Metallic solids have unusual
properties.

1. Molecular Solids
• molecular solids – atoms or molecules held together by intermolecular forces
• soft, low melting points
• gases or liquids at room temperature from molecular solids at low temperature
• properties depends on strengths of forces and ability of molecules to pack efficiently in
three dimensions
• intermolecular forces that depend on close contact are not as effective
2. Covalent-Network Solids
• covalent-network solids – atoms held together in large networks or chains by covalent
bonds
• hard, high melting points
3. Ionic Solids
• ionic solids – ions held together by ionic bonds
• strength depends on charges of ions
• structure of ionic solids depends on charges and relative sizes of ions
4. Metallic Solids
• metallic solids – metal atoms
• usually have hexagonal close-packed, cubic close-packed, or body-centered-cubic
structures
• each atom has 8 or 12 adjacent atoms
• bonding due to valence electrons that are delocalized throughout entire solid
• strength of bonding increases as number of electrons available for bonding increases
• mobility of electrons make metallic solids good conductors of heat and electricity
CHAPTER 2: PHYSICAL PROPERTIES OF SOLUTIONS

In the classification of matter, you learned that matter that can be classified as pure
substance and mixtures. Mixtures can either be homogeneous or heterogeneous. In heterogeneous
mixtures, the components cannot be identified from one another. Homogeneous mixtures are also
called solutions. (Homogeneous mixtures have uniform composition, they are made up of two or
more substances and can be separated by physical means like evaporation, filtration, and
distillation).

Solutions
A solution is a homogeneous mixture made up of atoms, ions, or molecules. It has two
phases, namely, the solute (dissolved medium) and the solvent (the dissolving medium). In a
mixture of instant coffee powder and hot water, the coffee powder is the solute and the solvent is
water. The substance present in smaller amount is the solute. In a mixture of two liquids where
water is one of the substances, the water is always the solvent and the solution is called an aqueous
solution.
There are three types of a solution based on the final state of the solution as shown in the
table. Some examples are also given for each type of solution as well as the state of the solvents
and solutes.
A solutions may also be classified based on the amount of solute present. A dilute solution
contains a great amount of solvent compared to the solute, whereas a concentrated solution is one
that has a greater amount of solute compared to the solvent.

Concentration of Solutions
Concentration refers to the amount of solute present in a given amount of solvent. A
solution is concentrated if it contains greater amount of solute compared to the solvent while a
solution is dilute if it contains a greater amount of solvent compared to the solute. The
concentration unit used was grams solute per 100g of solvent. There are other methods of
expressing the concentration of solutions that may be used and these are discussed in this section.
I. Percent by mass. This expresses the mass of solute per 100 g of solution. Mass of solution
is equal to the mass of solute plus the mass of solvent. A solution that contains 30% by mass
of sugar means that the solution contains 30 g of sugar and 70 g of water. It also means that
there are 30 g of sugar per 100 g of solution. The formula for percent by mass is:
𝑀𝑎𝑠𝑠 𝑠𝑜𝑙𝑢𝑡𝑒
% 𝑏𝑦 𝑚𝑎𝑠𝑠 𝑠𝑜𝑙𝑢𝑡𝑒 = 𝑥 100
𝑀𝑎𝑠𝑠 𝑜𝑓 𝑠𝑜𝑙𝑢𝑡𝑖𝑜𝑛

𝑀𝑎𝑠𝑠 𝑠𝑜𝑙𝑢𝑡𝑖𝑜𝑛 = 𝑚𝑎𝑠𝑠 𝑠𝑜𝑙𝑢𝑡𝑒 + 𝑚𝑎𝑠𝑠 𝑜𝑓 𝑠𝑜𝑙𝑣𝑒𝑛𝑡

One hundred grams of air consists of 78 g of nitrogen, 21 g of oxygen, 0.04 g CO 2, and 0.06 g of
other gases. This method is also frequently used in the field of allied health.
If the solution involves a solute and a solvent that are both liquids, then percent by volume
is used instead of by mass with the following formula:

𝑉𝑜𝑙𝑢𝑚𝑒 𝑠𝑜𝑙𝑢𝑡𝑒
% 𝑏𝑦 𝑣𝑜𝑙𝑢𝑚𝑒 𝑠𝑜𝑙𝑢𝑡𝑒 = 𝑥 100
𝑉𝑜𝑙𝑢𝑚𝑒 𝑜𝑓 𝑠𝑜𝑙𝑢𝑡𝑖𝑜𝑛
𝑉𝑜𝑙𝑢𝑚𝑒 𝑠𝑜𝑙𝑢𝑡𝑖𝑜𝑛 = 𝑣𝑜𝑙𝑢𝑚𝑒 𝑠𝑜𝑙𝑢𝑡𝑒 + 𝑣𝑜𝑙𝑢𝑚𝑒 𝑜𝑓 𝑠𝑜𝑙𝑣𝑒𝑛𝑡

II. Mole Fraction


Mole fraction is the ratio of the number of moles of one component to the total number of
moles in solution. It is represented by a capital letter X.
𝑀𝑜𝑙𝑒𝑠 𝑠𝑜𝑙𝑢𝑡𝑒
𝑀𝑜𝑙𝑒 𝐹𝑟𝑎𝑐𝑡𝑖𝑜𝑛 (𝑋) 𝑆𝑜𝑙𝑢𝑡𝑒 (𝐴) = 𝑀𝑜𝑙𝑒 𝐹𝑟𝑎𝑐𝑡𝑖𝑜𝑛 (𝑋) 𝑆𝑜𝑙𝑣𝑒𝑛𝑡 (𝐵)
𝑀𝑜𝑙𝑒𝑠 𝑠𝑜𝑙𝑢𝑡𝑖𝑜𝑛
𝑀𝑜𝑙𝑒𝑠 𝑠𝑜𝑙𝑣𝑒𝑛𝑡
=
𝑀𝑜𝑙𝑒𝑠 𝑠𝑜𝑙𝑢𝑡𝑖𝑜𝑛

𝑁𝑢𝑚𝑏𝑒𝑟 𝑜𝑓 𝑚𝑜𝑙𝑒𝑠 𝑠𝑜𝑙𝑢𝑡𝑖𝑜𝑛 = 𝑛𝑢𝑚𝑏𝑒𝑟 𝑜𝑓 𝑚𝑜𝑙𝑒𝑠 𝑠𝑜𝑙𝑢𝑡𝑒 + 𝑛𝑢𝑚𝑏𝑒𝑟 𝑜𝑓 𝑚𝑜𝑙𝑒𝑠 𝑠𝑜𝑙𝑣𝑒𝑛𝑡


𝑀𝑜𝑙𝑒 𝐹𝑟𝑎𝑐𝑡𝑖𝑜𝑛 𝑜𝑓 𝑠𝑜𝑙𝑢𝑡𝑒 + 𝑀𝑜𝑙𝑒 𝐹𝑟𝑎𝑐𝑡𝑖𝑜𝑛 𝑜𝑓 𝑠𝑜𝑙𝑣𝑒𝑛𝑡 = 1
𝑋𝐴 + 𝑋𝐵 =1

Note that mole Fraction is not affected by temperature.


III. Molarity
Concentration of solution may also be expressed in terms of molarity. Molarity is the
ratio of the number of moles of solute per liter of solution which is mathematically
expressed as
𝑛𝑢𝑚𝑏𝑒𝑟 𝑜𝑓 𝑚𝑜𝑙𝑒𝑠 𝑠𝑜𝑙𝑢𝑡𝑒
𝑀𝑜𝑙𝑎𝑟𝑖𝑡𝑦 =
𝑙𝑖𝑡𝑒𝑟 𝑜𝑓 𝑠𝑜𝑙𝑢𝑡𝑖𝑜𝑛
A 3M (3 molar) NaOH means that there are three moles NaOH in one liter of solution or
three moles of NaOH/liter solution. Dimensional analysis can be used to solve problems on
molarity.
IV. Molality
This is defined as the number of moles of solute per kilogram of solvent. It can also be
mathematically expressed as
𝑛𝑢𝑚𝑏𝑒𝑟 𝑜𝑓 𝑚𝑜𝑙𝑒𝑠 𝑠𝑜𝑙𝑢𝑡𝑒
𝑀𝑜𝑙𝑎𝑙𝑖𝑡𝑦 =
𝑘𝑖𝑙𝑜𝑔𝑟𝑎𝑚 𝑠𝑜𝑙𝑣𝑒𝑛𝑡

A 3m solution also means 3 moles solute/kilogram solvent.

Solution stoichiometry
Stoichiometry is solving problems using a balanced chemical equation. The number of
moles of a substance can be related to its molar mass and number of molecules. It can also be
related to the volume at Standard Temperature and Pressure (STP). Stoichiometry may also include
solutions in through molarity. The relationship among mole, mass, molecule, and volume is given
below.

Mass Mass
𝑚 𝑚
𝑛= 𝑛=
𝑀 𝑀

𝑛 𝑛
𝐶= Balanced 𝐶=
Concentration 𝑉 Reactants Products 𝑉 Concentration
Mole Equation Mole

Volume Volume

The diagram shows that given quantities, whether mass, concentration, or volume must first
be converted to reactant moles. The number of moles of reactants calculated should be
related to the number of moles of the products based on the balanced chemical equation.
And depending on the unknown, this number of moles of products could then be converted
to mass, concentration, or volume.
Another application of stoichiometric problems in titration. There are some instances when
there is a need to determine the concentrations of unknown solutions like the concentrations of
ions in a sample of river water. If there are concerns about the purity of water, the concentrations
of minerals like Pb+2, Hg+2, and Cd+2 are determined.
The determination is done through a stoichiometric procedure known as titration. Titration
is a process where the concentration of an unknown solution is determined by reacting it with a
solution with known concentration. The amount of the standard solution to be added is determined
by the end point which is indicated by a color change. The end point indicates that enough of the
standard solution has reacted with the unknown solution.
The standard solution should be well prepared and must come from a stock solution.
Preparation uses the dilution method. Only solutions of lower concentration can be prepared from
stock solutions or solutions of higher concentration. Dilution adds water to the solution of known
volume and concentration. The number of moles of solute does not change. Only the volume of
the solution changes. The relationship is expressed mathematically below.

𝑚𝑜𝑙𝑒𝑠 𝑠𝑜𝑙𝑢𝑡𝑒 𝑠𝑡𝑜𝑐𝑘 𝑠𝑜𝑙𝑢𝑡𝑖𝑜𝑛 = 𝑚𝑜𝑙𝑒𝑠 𝑠𝑜𝑙𝑢𝑡𝑒 𝑑𝑖𝑙𝑢𝑡𝑒𝑑 𝑠𝑜𝑙𝑢𝑡𝑖𝑜𝑛


𝑉1 𝐶1 = 𝑉2 𝐶2

Where C1 is the initial concentration, V1 is the initial volume, C2 is the final concentration and V2
is the final volume.

𝐶2 𝑖𝑠 𝑙𝑒𝑠𝑠 𝑡ℎ𝑎𝑛 𝐶1 𝑉2 𝑖𝑠 𝑔𝑟𝑒𝑎𝑡𝑒𝑟 𝑡ℎ𝑎𝑛 𝑉1 𝑉2 = 𝑉1 + 𝑉𝐻2 𝑂

Solubility
Solubility refers to the amount of solute that can dissolve in a given amount of solvent at
room temperature under given conditions. For example, if 1.0g of sugar is placed in 100 g of water
at 35oC, all of the sugar dissolves. A small portion of sugar is added and the sugar still dissolves.
This means that the 100 g water can still accommodate sugar. The solution is called an unsaturated
solution. If sugar is continuously added, there comes a point when the sugar no longer dissolves.
This means that the 100 g of water already contains the maximum amount of sugar it can hold at
room temperature. This indicates that the solution is already saturated. If the maximum amount of
sugar that can dissolve in 100g of water at 35oC is 70 g, then the solubility of sugar is 70 g/100g
water at 35oC. However, if the sugar solution is heated to 55oC, the solution can dissolve an
additional amount of sugar and the solution becomes supersaturated. Therefore, solubility refers
to the amount of solute that can dissolve in a given amount of solvent at a specified temperature
to produce a saturated solution. It is expressed in grams of solute per 100 g of solvent at a specified
temperature and pressure.

Unsaturated Saturated Supersaturated

Definition The minimum The maximum amount The maximum


amount of solute of solute that can amount of solute
present in a given dissolve in a given present in a given
amount of solvent at amount of solvent at amount of solvent at
room temperature. room temperature. an elevated
temperature.

Description When solute is When solute is added, The solute crystallizes


added, the solute the solute no longer when the solution is
dissolves. dissolves. heated.

Table 1. This table compares an unsaturated, saturated, and a supersaturated solution

When a solute dissolved in a solvent, energy is involved. If heat is absorbed when a solute
dissolves in a solvent, the final temperature of the solution is increased and the reaction is called
an endothermic reaction. A reaction is exothermic if heat is given off or released when a solute
dissolves in a given amount of solvent. The temperature of the initial state is higher than the final
state of the solution.
Solutes may be classified as soluble, slightly soluble, or insoluble depending on the amount
of solute that dissolves in a given amount of solvent at room temperature.
A. Insoluble – a solution is insoluble if less than or equal to 0.1 g of solute dissolves in 100 g
of solvent.
B. Slightly soluble – the amount of solute that dissolves in 100 g of solvent is greater than 0.1
g but less than or equal to 10.0 g.
C. Soluble – if the amount of solute that dissolves in 100 g of solvent a greater than or equal
to 0.01 g.
Factors Affecting Solubility
There are several factors that influence solubility and these include the following:
A. Nature of solute and solvent
As previously discussed, “like dissolves like”. Ionic compounds break up into their
component ions in water. The positive ion of the ionic compound becomes surrounded by
the partial negative end of the water (oxide ions) and the negative end of the ionic
compound is surrounded by the partial positive end (hydrogen) of water. Therefore, ionic
solute dissolves in ionic solvent ionic or polar solute dissolves in polar solvents, and
nonpolar solutes dissolve in nonpolar solvent. Another interaction of the solute and the
solvent is the formation of hydrogen bond as shown previously formed between ethyl
alcohol and water.
Potassium permanganate is an ionic compound and will break up into potassium and
permanganate ions in water. The potassium becomes surrounded by the oxide ion of water
and the permanganate ions are surrounded by the hydrogen ions of water.
Solute Solvent Solubility/Miscibility

Polar Polar Soluble/miscible


Nonpolar Nonpolar Soluble/miscible
Nonpolar Polar Insoluble/immiscible
Ionic Polar Soluble/miscible

Table 2. Different pairings can also be tabulated to predict whether two substances are
soluble/miscible or insoluble/immiscible.

B. Effect of Temperature
1. Solubility of Gas in Liquid
The effect of temperature is different for the solubility of solid in liquid from that of the
solubility of gases in liquids. A cold can of soda tastes quite differently after letting it stand
opened for a few hours and then drinking it up after. At high temperature, the solubility of
gas in liquids decreases because the gas molecules move faster and tend to escape. Since
the carbon dioxide escapes, the taste of the soda becomes flat.
In summary, the solubility of gas in liquid increases with a decrease in temperature and,
decreases with an increase in temperature.
2. Solubility of Solid in Liquid
In general, the solubility of solids increases with an increase in temperature for
endothermic reactions while it decreases for exothermic reactions. For example, in a
hypothetical reaction given below:
A + B + heat AB

Heat is needed to dissolve A in B. For this reason, if temperature increases, more A will
dissolve in B and if temperature decreases, a lower amount of A will dissolve in B. For an
exothermic reaction illustrated below:
AB A + B + heat

Heat is given off or the temperature is lowered to drive the dissolution of A in B. For
this reason, any increase in temperature decreases solubility and a decrease in temperature
increases solubility.
3. Effect of Pressure
Solubility of gas in liquids. The solubility of solids in liquids are not affected by
pressure. However, a change in pressure influences the solubility of gas in liquids. When
pressure is increased, the molecules or ions come closer to each other and there are greater
chances for interaction between the solute and the solvent. Thus, solubility of gas in liquids
increases with an increase in pressure. This is known as Henry’s law named in honor of
William Henry (1774-1836) who conducted experiments on the solubility of gases in
liquids. Bottled sodas are produced under reduced pressure. When the bottle is opened, the
carbon dioxide escapes, which is indicated by bubbling, and the partial pressure of CO2
above the solution decreases. As the pressure drops, the solubility of CO 2 in the solution
decreases. If the bottle is left opened for some time, all of the carbon dioxide will
completely escape and the taste of the beverage becomes flat.
4. Surface Area
Interaction between the solute and the solvent occurs at the surface area. Therefore, the
greater the surface area the greater the interaction of the solute and the solvent, and thus
solubility increases. The surface area can be increased by grinding the solid into finer particles.
5. Stirring
As solute is added to the solvent, solute particles tend to concentrate in a section of the
mixture and the dissolving process slows down. Stirring will disperse the solute into the
sections of the solvent, increasing the dissolution process.
CHAPTER 3 : THERMOCHEMISTRY

Energy Changes in Chemical Reactions: Endothermic and Exothermic Processes


Changes in matter are usually accompanied by absorption or evolution of energy. The
energy change observed is actually a consequence of breaking and formation of bonds. To separate
atoms, energy is needed to overpower the forces that hold them together. In short, energy is
absorbed to break bonds. Conversely, when bonds are formed, energy is evolved. Such energy is
most commonly observed as heat. If heat is absorbed during the process, the reaction is
endothermic. If heat is evolved, the change is exothermic.
Endothermic and Exothermic Process
In an endothermic process, heat flows into the system from the surroundings. Melting of
ice, for example, is an endothermic process. Hence, since I am a part of the surroundings, if I touch
the ice, it would feel cold because the heat from my fingers is passed on to the ice. The ice also
absorbs heat from the air around it causing the temperature of the surroundings to drop. If the ice
is placed in a glass of water, the temperature of water and the glass will drop since the heat from
water and the glass will flow to the ice. Heat is generally transferred from a hotter to a colder one.
For some reason, our fingers may get burned upon touching a hot object since the heat from the
object is transmitted to the fingers in contact with it.
Conversely, in an exothermic process, heat flows out of the system into the surroundings.
Fuels like LPG (liquefied petroleum gas) or wood, give off heat when they are burned. Upon
burning, the system releases heat to the surroundings which might be the air around the stove or
the food that is being cooked. As a consequence of heat transfer, there is an increase in the
temperature of the surroundings. Similarly, mixing concentrated acid and water is exothermic.
Because of the heat produced the surroundings become hot. Thus, water is never added to acid due
to localized heating which may cause spattering or violent reaction. Hence, it is safer to add acid
to water because dilution would lessen the heat effect.
First Law of Thermodynamics
The term thermodynamics is derived from the Greek word “therme” which means heat and
“dynamo” which means power. Thermodynamics deals with the study of energy and its
transformation. This field of study will help in understanding the relationship between heat and
work in all living and nonliving things. A living cell, for instance, evolves heat, does work, and
utilizes energy obtained from burning of food. Burning of gasoline inside an internal combustion
engine provides the push that propel the vehicle, aside from giving out heat. A moving machine
that does work also generates heat. Thermodynamics will provide guidelines that will enable us to
understand the energetics and directions of reactions.
According to the law of conservation of energy can neither be created nor destroyed. In
short, energy is conserved. This is also known as the first law of thermodynamics. The law further
states that energy may be transformed from one form to another and that any energy lost by the
system must be gained by surroundings, and vice versa. Energy can also be transferred back and
forth between the system and the surroundings in the form of heat and work.
The energy contained within the system, referred to as internal energy, E, is simply the sum of the
kinetic and potential energy of all the components of the system. The change in internal energy,
∆𝐸, of the system is a state function, thus, its value is independent of the manner in which the state
of the system is attained. The value is the same regardless of the pathway and depends only on the
initial and final states of the system. It does not depend on how and where the change is carried
out. Thus, if a system is transformed from state A, to a different state B, the value of ∆𝐸 is
expressed as
∆𝐸 = 𝐸𝐵 − 𝐸𝐴
[3.1]
If the transformation is carried out by exchange of heat and performance of work between
a system and its surroundings, the total change in energy of a system, ∆𝐸, is equal to the sum of
the heat absorbed or evolved by the system, q, and the work performed by or done on the system,
w, as expressed by the following equation:
∆𝐸 = 𝑞 + 𝑤 [3.2]

Unlike ∆𝐸, q and w, are not state functions and, therefore, path dependent. Hence, the
amount of heat and work formed during a change in state of a system depend on the way the change
is carried out. However, even though the individual values of q and w change, their sum is a state
function. For example, if the path from initial to final state increases the value of q, the value of w
decreases by the same amount and vice versa.
In applying the equation [3.2] that expresses the first law of thermodynamics, the sign of q
and w must always be taken into account.
• q and w are positive (+) when heat or work enters the system from the surroundings, i.e.,
heat is absorbed by the system and work is done on the system.
• q and w are negative (-) when heat or work transfers from the system to the surroundings,
i.e., heat is evolved and work is done by the system.
• If both q and w are (+), sign of ∆𝐸 is also (+); internal energy increases
• If both q and w are (-), sign of ∆𝐸 is also (-); internal energy decreases
• If q is (+) and w is (-), or, if q is (-) and w is (+), sign of ∆𝐸 depends on the magnitudes of
q and w
Heat added to the system Heat evolved by the system
(+q) SYSTEM (-q)
Work done on the system Work done by the system
(+w) (-w)
SURROUNDINGS
Enthalpy of Chemical Reactions
The term enthalpy, from the Greek word, enthalpein, meaning “to warm” refers to the
energy transferred under constant pressure. It is represented by the symbol, H. It is often referred
to, as heat content. Like E (internal energy), enthalpy, H, is also a state function, and therefore, not
path dependent. Enthalpy cannot be measured, but, it is possible to measure the change in enthalpy
or heat content, ∆𝐻.
For reactions that occur in an open container, at constant pressure, ∆𝐻 is equal to the
difference in enthalpy between the final and initial states of the system as given by the equation:
∆𝐻 = 𝐻𝑓𝑖𝑛𝑎𝑙 − 𝐻𝑖𝑛𝑖𝑡𝑖𝑎𝑙 [3.3]

Enthalpy change, ∆𝐻, that accompanies chemical reaction is called heat of reaction. This
is the net energy change resulting from the breaking and the making of bonds. It represents heat
absorbed or evolved when the reactants are converted into products, at constant pressure.
∆𝐻 = 𝐻𝑝𝑟𝑜𝑑𝑢𝑐𝑡𝑠 − 𝐻𝑟𝑒𝑎𝑐𝑡𝑎𝑛𝑡𝑠 [3.4]

∆𝐻 = 𝑞 (𝑎𝑡 𝑐𝑜𝑛𝑠𝑡𝑎𝑛𝑡 𝑝𝑟𝑒𝑠𝑠𝑢𝑟𝑒) [3.5]

The sign of ∆𝐻 indicates the direction of heat transfer. A positive value of ∆𝐻, indicates
that the system absorbs heat from the surroundings and is, therefore, endothermic. A negative
value indicates that heat is released by the system, hence, the process is exothermic.
Endothermic: 𝑞 = ∆𝐻 > 0 ; 𝐻𝑝𝑟𝑜𝑑𝑢𝑐𝑡𝑠 > 𝐻𝑟𝑒𝑎𝑐𝑡𝑎𝑛𝑡𝑠 [3.6
Exothermic: 𝑞 = ∆𝐻 < 0 ; 𝐻𝑝𝑟𝑜𝑑𝑢𝑐𝑡𝑠 < 𝐻𝑟𝑒𝑎𝑐𝑡𝑎𝑛𝑡𝑠 [3.7]
]

Surroundings Surroundings

Heat SYSTEM Heat SYSTEM


∆𝐻 > 0 ∆𝐻 > 0

Endothermic Exothermic
Thermochemical Equations
The enthalpy of relation between reactants and products is usually shown in a
thermochemical equation. A thermochemical equation is a balanced equation in which the value
of ∆𝐻, with the appropriate sign, is usually written at the right side. Since both the physical state
and amount of substances involved in the process affect the magnitude of ∆𝐻, a thermochemical
equation also shows whether the substance is a solid, liquid, gas, or aqueous, while the coefficients
represents the number of moles of reactants and products. Furthermore, revising the reaction would
lead to a ∆𝐻 value that has the same magnitude but opposite in sign as shown in the following
reaction. Let us consider the combustion of carbon to form carbon dioxide.

𝐶(𝑠) + 𝑂2 (𝑔)

∆𝐻𝑓𝑜𝑟𝑤𝑎𝑟𝑑 = −393.5 𝑘𝐽 ∆𝐻𝑟𝑒𝑣𝑒𝑟𝑠𝑒 = + 393.5 𝑘𝐽

𝐶𝑂2 (𝑔)

Reversing the reaction changes the sign but not the value of ∆𝐻.

A typical example of a thermochemical equation that is endothermic is the conversion of


methyl alcohol to carbon monoxide and hydrogen. Note that ∆𝐻 is positive or ∆𝐻 > 0.
𝐶𝐻3 𝑂𝐻(𝑔) → 𝐶𝑂(𝑔) + 2𝐻2 (𝑔) ∆𝐻 = +90.7 𝑘𝐽 (𝑒𝑛𝑑𝑜𝑡ℎ𝑒𝑟𝑚𝑖𝑐)
On the other hand, an exothermic process, such as formation of hydrofluoric acid from
hydrogen and fluorine, is represented by following equations.
𝐻2 (𝑔) + 𝐹2 (𝑔) → 2𝐻𝐹(𝑔) ∆𝐻 = − 557 𝑘𝐽 (𝑒𝑥𝑜𝑡ℎ𝑒𝑟𝑚𝑖𝑐)
Enthalpy also indicates the amount of heat stored as potential energy. This implies that in
an exothermic reaction, the products will have a lower energy than the reactants. The heat content
of the system is lower after an exothermic process, since heat is released by the system to the
surroundings.
To show that enthalpy is an extensive property, meaning, the value of ∆𝐻 is dependent on
the amount of reactants consumed, let us consider the combustion of acetone, C 3H6O (molar mass
= 58g/mol).
𝐶3 𝐻6 𝑂(𝑙) + 4𝑂2 (𝑔) → 3𝐶𝑂2 (𝑔) + 3𝐻2 𝑂(𝑙) ∆𝐻 = − 1790 𝑘𝐽

The thermochemical equation shows that burning 1 mole of acetone in the presence of 4
mole of oxygen gas releases 1790 kJ of heat. This information can be considered as a
stoichiometric relation:
1 mole C3H6O =1790 kJ
2 moles C3H6O = 2(1790) kJ
Since 1 mole C3H6O has a mass of 58g, the following stoichiometric relation can also be used:
Since 1 mole C3H6O = 58 g C3H6O
Therefore, 58 g C3H6O =1790 kJ
CHAPTER 4: CHEMICAL KINETICS

The rate of any event is measured by the amount of change in a gall events given unit of
time. For example, a car may travel from Manila to Baguio at 80.0km/hour. An elephant may walk
2 meters/minute. A student may read a book at a rate of 1 page per minute or a factory worker may
finish packaging a product at a rate of 20 pieces/30minutes. Almost all events can be measured in
terms of rate. Rate then, gives a measure of change per unit time. The reactants disappear and new
substance, products, are formed. The change from reactants to products take place at different
length of time depending on the reaction. It may be slow, moderately fast, or a very fast reaction.
Rates of reaction are measured by a change in the concentration of reactants and products per unit
time. During the course of reaction, reactants are continuously being changed to products. So, at
any given time, the concentrations of the reactants and products are continuously being change.
When the reaction comes to an end, all of the reactants had been converted into products.
The rate of reaction can then be measured by a change in the concentration of the reactants
and products. As reactants disappear, products appear. If we start with 100 particles of substance
A and no products have been formed. After 10 seconds, only 80 particles of A have been left and
20 particles of the new substance B appear. After 20 second, 60 particles of A are present and 40
particles of B appear. The reaction continues until all the particles of a A have disappeared and
have been converted to the new substance, B.

Time A B
(seconds) (no. of particles) (no. of particles)
0 100 0
10 80 20
20 60 40
30 40 60
40 60 40
50 80 20
60 100 0

The reaction rate is the change in the concentration of the reactants and products per unit
time. It cannot be calculated using a chemical equation whish is what is done in stoichiometry.
The rate of reaction is determined experimentally. For example, for a hypothetical reaction where
A is converted to B, initially, only A is present. After 10 minutes, 0.74 of A is present. The 0.26
has been converted to B. After another 10 minutes, the amount of B present is observed to be 0.46
whereas A further decreased to 0.54. After another 10 minutes 0.40 of A and 0.60 of B are present.
The decrease in the amount of A and an increase in the concentration of B is observed. After 60
minutes, 0.16 A and 0.84 B are present.
Factors affecting Reaction rate
A reaction rate sometimes proceeds slowly, moderately slowly, or quite fast. For example,
food is placed inside the refrigerator to prevent its spoilage hot water is used to dissolve greasy,
more detergent is used to make clothes cleaner, and skin care ingredients that could retard aging
have been explored and many of products have been developed and marketed. The rate of reaction
is influenced by several factors, namely, (1) nature of reactants, (2) temperature, (3) concentration
(4) surface area, and (5) catalysts. This section will discuss each of the different factors.
1. Nature of Reactants
Different substance react differently. There are those that react faster and more
violently than others. For example, sodium metal reacts violently with water producing
hydrogen gas, but gold does not produce any reaction with water. Hydrogen reacts violently
with fluorine, but the reaction of hydrogen and iodine occurs extremely slowly. The rate of
reaction depends on the reactants. Since bonds are broken and rearranged during chemical
reaction, the rate of reaction depends on the complexity of the bonds to be broken. The
more complex the bonds to be destroyed, the slower the rate of reaction. On the other hand,
if bonds are simple, the reactions will occur faster. The reaction between hydrogen and
oxygen to form water occurs explosively and very rapidly because the bonds of hydrogen
and oxygen are simple. When the reactants contain many covalent bonds which are to be
broken, then the reaction would be slow. The formation of polyester is a slow reaction
because the polymer is produced by alternating combination of terephthalic acid and
ethylene glycol. The state of reactant also affects the rate of reaction. The reaction between
two gases would occur faster than the reaction between two liquids. The reaction between
two liquids would occur faster than the reaction between two solids. This is because the
molecules in gases are at far distances from one another according to the Kinetic Molecule
Theory. A reaction that has the same state of matter for all the reactants is called a
homogeneous reaction. A reaction in which the reactants are in different states is called a
heterogeneous reaction.

2. Surface Area
You have learned that the greater the number of effective collision, the faster the
reaction rate is. This factor will, therefore, affect the reaction involving solids. The reaction
of magnesium and hydrochloric acid, for example, will be faster if the surface area of
magnesium is increased because an increased surface area increases the possibility of
greater number of collision.

𝑀𝑔 (𝑠) + 𝐻𝐶𝑙 (𝑎𝑞) 𝑀𝑔𝐶𝑙2 (𝑎𝑞) + 𝐻2 (𝑔)

In this reaction, the hydrogen ions in HCI collides with the outer part of the surface
area of one piece of magnesium ribbon. If the piece of the magnesium ribbon were cut into
three pieces, there will be a greater area to which the hydrogen ions can collide with
magnesium. If the magnesium ribbon were divided into six pieces, then there will even be
a greater area through which the hydrogen ions can collide with the metal.

3. Temperature
We know for a fact that food is placed in the refrigerator to avoid rapid spoilage.
At lower temperatures, the reaction rate is lower. Plants grow faster in warm weather than
in the cold. This shows that temperature greatly affects reaction rate.
At higher temperatures, the molecules move faster. The faster the movement of the
reacting molecules, the greater the chances for molecular collision. Based on the collision
theory, the greater the number of molecular collision, the faster the reaction rate. The
effective collision depends on the kinetic energy of the colliding molecules. The colliding
molecules must have sufficient energy to overcome the activation energy barrier. If the
temperature is increased, the kinetic energy of each of the molecule increase. The increase
in temperature increases the fraction of the molecules that collide with one another with
sufficient energy to form the activated complex. The rule of thumb by chemists is that for
every 10oC increase in temperature, the reaction rate approximately doubles or triple. The
effect of this increase can be determined experimentally.

4. Catalyst
Sugar can be fermented and turned into ethyl alcohol in the presence of yeast. The
substance in yeast that converts the sugar to alcohol is an enzyme. Enzymes are present in
the body and are responsible for reactions that extract energy from food we eat. They
increase the speed of reactions essential to life. Without these enzymes, reactions would
occur extremely slowly and life may not be possible. An enzyme is a catalyst. A catalyst
is a substance that speeds up a reaction and does not appear in the overall chemical
equation. In fact, a catalyst does not have permanent effect and can be regenerated at the
end of the reaction.

5. Concentration
The reaction rate was earlier defined in this chapter as the change in the concentration
of reactants or products per unit of time. For this reason, concentration has great effect on
the reaction rate. For the reaction

𝐴2 + 𝐵2 A2B2 2AB

If the concentration of A2 or B2 is increased, there will be a greater number of particles


for a given space. Since the distances between the particles will be decreased because of
an increased number of particles, there will be greater chances of collision among the
particles. The increased frequency of in collision, increases the reaction rate. The
concentration of reactant is directly proportional to the reaction rate. If the concentration
of reactant is doubled, then the reaction rate is also doubled. If the concentration of reactant
increases by a factor of 4, the reaction rate also increases by a factor of 4.

The Rate Law


The mathematical relationship of reaction rate with reactant concentrations is known as
the rate law. This relationship may rely more heavily on the concentration of one particular
reactant, and the resulting rate law may include some, all, or none of the reactant species involved
in the reaction.

For the following hypothetical reaction

aA+bB→cC

the rate law can be expressed as:

Rate = k[A]y[B]z

The proportionality constant, k, is known as the rate constant and is specific for the reaction
shown at a particular temperature. The rate constant changes with temperature, and its units depend
on the sum of the concentration term exponents in the rate law. The exponents (y and z) must be
experimentally determined and do not necessarily correspond to the coefficients in the balanced
chemical equation.

Reaction Order

The sum of the concentration term exponents in a rate law equation is known as its reaction
order. We can also refer to the relationship for each reactant in terms of its exponent as an order.

For the following reaction between nitrogen dioxide and carbon monoxide:

NO2(g) + CO(g) → NO(g) + CO2(g)

The rate law is experimentally determined to be: rate = k[NO2]2

Therefore, we would say that the overall reaction order for this reaction is second-order (the sum
of all exponents in the rate law is 2), but zero-order for [CO] and second-order for [NO2].

The reaction order is most often a whole number such as 0, 1, or 2; however, there are instances
where the reaction order may be a fraction or even a negative value.

Earlier it was mentioned that the units of the rate constant depend on the order of the reaction.
Let’s quickly examine why this occurs. A simplified rate law can be expressed generically in the
following way:
Rate = k[reactant]y

Units of rate = (units of rate constant)(units of concentration) y

Units of rate constant = =

Therefore, the units of the rate constant should be:

Reaction Order Units of rate


constant

Zero-order M s-1

First-order s-1

Second-order L mol-1 s-1

Determining Rate Laws from Initial Rates

The rate law can be determined experimentally using the method of initial rates, where the
instantaneous reaction rate is measured immediately on mixing the reactants. The process is
repeated over several runs or trials, varying the concentration one reactant at a time. These runs
can then be compared to elucidate how changing the concentration of each reactant affects the
initial rate.

Example
The initial rate of reaction for the reaction E + F → G was measured at three different
initial concentrations of reactants as shown in the table.
1. Determine the rate law of the reaction.
2. Determine the rate constant.

Trial Initial Rate [E] (mole L-1) [F] (mole L-1)


(mole L-1 s-1)
1 2.73 x 10-5 0.100 0.100

2 5.47 x 10-5 0.200 0.100

3 2.71 x 10-5 0.100 0.200

Solution
1. Comparing trials 1 and 2, [E] is doubled, while [F] and the rate constant are held
constant. This comparison will allow us to determine the order of reactant E:
=

2.00 = 2.00y
y=1
Therefore, the reaction is first order with respect to [E].
Comparing trials 1 and 3, [F] is doubled, while [E] and the rate constant are held constant.
This comparison will allow us to determine the order of reactant F:
=

0.993 = 2.00z
z=0
Therefore, the reaction is zero order with respect to [F].
The rate law can now be written as:
Rate = k[E]1
2. Using the rate law we have just determined, substitute in the initial concentration
values and initial rate for any trial and solve for the rate constant:

Rate = k[E]1
Using Trial 1:
k

k = 2.73 x 10-4 s-1

Collision Theory

Collision theory explains why different reactions occur at different rates, and suggests ways to
change the rate of a reaction. Collision theory states that for a chemical reaction to occur, the reacting
particles must collide with one another. The rate of the reaction depends on the frequency of collisions. The
theory also tells us that reacting particles often collide without reacting. For collisions to be successful,
reacting particles must (1) collide with (2) sufficient energy, and (3) with the proper orientation.

Molecules must collide before they can react


This rule is fundamental to any analysis of an ordinary reaction mechanism. It explains
why termolecular processes are so uncommon. The kinetic theory of gases states that for every
1000 binary collisions, there will be only one event in which three molecules simultaneously come
together. Four-way collisions are so improbable that this process has never been demonstrated
in an elementary reaction.

Consider a simple bimolecular step:


A + B → Products (1)
If the two molecules A and B are to react, they must approach closely enough to disrupt
some of their existing bonds and to permit the creation of any new ones that are needed in the
products. Such an encounter is called a collision.
The frequency of collisions between A and B in a gas is proportional to the concentration
of each; if [A] is doubled, the frequency of A-B collisions will double, and doubling [B] will have
the same effect. If all collisions lead to products, then the rate of a bimolecular process is first-
order in A and in B, or second-order overall:
rate=k[A][B] (2)

Not all Collisions are Equal


For a gas at room temperature and normal atmospheric pressure, there are about
1033 collisions in each cubic centimeter of space every second. If every collision between two
reactant molecules yielded products, all reactions would be complete in a fraction of a second.
For example, when two billiard balls collide, they simply bounce off of each other. This is the
most likely outcome if the reaction between A and B requires a significant disruption or
rearrangement of the bonds between their atoms. In order to effectively initiate a
reaction, collisions must be sufficiently energetic (or have sufficient kinetic energy) to bring about
this bond disruption. This is further discussed below.

There is often one additional requirement. In many reactions, especially those involving
more complex molecules, the reacting species must be oriented in a manner that is appropriate
for the particular process. For example, in the gas-phase reaction of dinitrogen oxide with nitric
oxide, the oxygen end of N2O must hit the nitrogen end of NO; altering the orientation of either
molecule prevents the reaction. Owing to the extensive randomization of molecular motions in a
gas or liquid, there are always enough correctly-oriented molecules for some of the molecules to
react. However, the more critical this orientational requirement is, the fewer collisions will be
effective.
Energetic collisions between molecules cause interatomic bonds to stretch and bend,
temporarily weakening them so that they become more susceptible to cleavage. Distortion of the
bonds can expose their associated electron clouds to interactions with other reactants that might
lead to the formation of new bonds.
Figure 2: Anatomy of a collision

Chemical bonds have some of the properties of mechanical springs: their potential
energies depend on the extent to which they are stretched or compressed. Each atom-to-atom
bond can be described by a potential energy diagram that shows how its energy changes with its
length. When the bond absorbs energy (either from heating or through a collision), it is elevated
to a higher quantized vibrational state (indicated by the horizontal lines) that weakens the bond
as its length oscillates between the extended limits corresponding to the curve.
A particular collision will typically excite a number of bonds in this way. Within about 10 –
seconds, this excitation is distributed among the other bonds in the molecule in complex and
13

unpredictable ways that can concentrate the added energy at a particularly vulnerable point. The
affected bond can stretch and bend farther, making it more susceptible to cleavage. Even if the
bond does not break by pure stretching, it can become distorted or twisted so as to expose nearby
electron clouds to interactions with other reactants that might encourage a reaction.
CHAPTER 5: CHEMICAL THERMODYNAMICS

There are two types of processes (or reactions): spontaneous and non-spontaneous.
Spontaneous changes, also called natural processes, proceed when left to themselves, and in the
absence of any attempt to drive them in reverse. The sign convention of changes in free energy
follows the general convention for thermodynamic measurements. This means a release of free
energy from the system corresponds to a negative change in free energy, but to a positive change
for the surroundings. Examples include:

• a smell diffusing in a room


• ice melting in lukewarm water
• salt dissolving in water
• iron rusting.

The laws of thermodynamics govern the direction of a spontaneous process, ensuring that if
a sufficiently large number of individual interactions (like atoms colliding) are involved, then the
direction will always be in the direction of increased entropy.

Spontaneous Processes

Spontaneity does not imply that the reaction proceeds with great speed. For example, the
decay of diamonds into graphite is a spontaneous process that occurs very slowly, taking millions
of years. The rate of a reaction is independent of its spontaneity, and instead depends on the
chemical kinetics of the reaction. Every reactant in a spontaneous process has a tendency to form
the corresponding product. This tendency is related to stability.

Nonspontaneous Processes

An endergonic reaction (also called a nonspontaneous reaction or an unfavorable reaction)


is a chemical reaction in which the standard change in free energy is positive, and energy is
absorbed. The total amount of energy is a loss (it takes more energy to start the reaction than what
is gotten out of it) so the total energy is a negative net result. Endergonic reactions can also be
pushed by coupling them to another reaction, which is strongly exergonic, through a shared
intermediate. Saul Steinberg from The New Yorker illustrates a nonspontaneous process here.

Entropy

Observations of natural processes led a surprising number of chemists of the late 19th
century (including Berthelot and Thomsen) to conclude that all spontaneous reactions must be
exothermic since:
• Objects roll downhill spontaneously (i.e., energy is "lost" from the system)
• Objects do not roll uphill spontaneously (i.e., energy does not suddenly appear from
nowhere)

If this were true all we would need to predict whether a reaction is spontaneous is the
change of enthalpy. If ΔHΔH were negative, the process should be able to occur by itself. If ΔH
were positive, the reaction could not occur by itself.
Indeed, almost all exothermic reactions are spontaneous at standard thermodynamic conditions (1
atmosphere pressure) and 25o C. However a number of common processes which are both
endothermic and spontaneous are known. The most obvious are simple phase changes, like ice
melting at room temperature. Also, many solids dissolve in water and simultaneously absorb heat.

So the energy is now dispersed among the molecules of liquid water which have access to
all kinds of molecular motion states that were not available in the solid. At the same time, the
ordered structure of the solid ice has given way to a much less organized flowing liquid:

But the actual change has occurred in the energy dispersal. This subtle property that matter
possesses in terms of the way energy is dispersed in it is known as entropy. Entropy is sometimes
erroneously referred to as "randomness" or even "disorder" but these descriptions do not fit the
state of energy as well as they seem to describe some of the often obvious results.

Just as reactions which form stronger bonds tend to occur spontaneously, energy is
constantly being dispersed or "spread out" in any process which either happens on its own
or which we make happen.

It is the Second Law of Thermodynamics which gives us the criterion we are seeking to
decide whether a reaction will be spontaneous or not (well, almost...):

In a spontaneous process the entropy of the universe increases.

The "universe" is a pretty big place. Recall the definitions we used when we introduced
chemical thermodynamics. The system is that part of the universe on which we focus our attention-
-generally chemicals. The surroundings are everything else. Taken together they constitute the
universe.
So the Second Law could be written this way for a spontaneous process:
ΔSsys + ΔSsurr > 0
Entropy trends and physical properties
(values in )
1. Entropy increases with mass
• F2(g) = 203 J/mol·K
• Cl2(g) = 224 J/mol·K
• Br2(g) = 245 J/mol·K
• I2(g) = 261 J/mol·K\

2. Entropy increases with melting, vaporization or sublimation


• I2(s)I2(s) = 117 J/mol·K vs. I2(ℓ) = 261 J/mol·K and
• H2O(ℓ)H2O (ℓ) = 70 J/mol·K vs. H2O(g) = 189 J/mol·K

3. Entropy increases when solids or liquids dissolve in water


• CH3OH(ℓ) = 127 J/mol·K vs. CH3OH(aq) = 132 J/mol·K and
• NaCl(s) = 72J/mol·K vs. Na+(aq) + Cl-(aq) = 115 J/mol·K

4. Entropy decreases when a gas is dissolved in water


• HCl(g)HCl(g) = 187 vs. H+(aq) + Cl-(aq) = 55

5. Entropy is lower in hard, brittle substances than in malleable solids like metals
• Diamond (C) = 2.4J/mol·K vs. Pb = 65 J/mol·K

6. Entropy increases with chemical complexity


• NaCl = 72 J/mol·K vs. MgCl2 = 90 J/mol·K vs. AlCl3 = 167 J/mol·K
Of course, the main issue here is how entropy changes during a process. This can be determined
by calculation from standard entropy values (\(S^o\)) in the same way that enthalpy changes are
calculated:

Σ 𝑆 𝑜 𝑝𝑟𝑜𝑑𝑢𝑐𝑡𝑠 − Σ𝑆 𝑜 𝑟𝑒𝑎𝑐𝑡𝑎𝑛𝑡𝑠 = ∆𝑆 𝑜 𝑟𝑥𝑛

The Second Law of Thermodynamics

The second law of thermodynamics states that for any spontaneous process, the overall ΔS
must be greater than or equal to zero; yet, spontaneous chemical reactions can result in a negative
change in entropy. This does not contradict the second law, however, since such a reaction must
have a sufficiently large negative change in enthalpy (heat energy). The increase in temperature of
the reaction surroundings results in a sufficiently large increase in entropy, such that the overall
change in entropy is positive. That is, the ΔS of the surroundings increases enough because of the
exothermicity of the reaction so that it overcompensates for the negative ΔS of the system. Since
the overall ΔS = ΔSsurroundings + ΔSsystem, the overall change in entropy is still positive.

Gibbs Free Energy

What happens when one of the potential driving forces behind a chemical reaction is
favorable and the other is not? We can answer this question by defining a new quantity known as
the Gibbs free energy (G) of the system, which reflects the balance between these forces.

The Gibbs free energy of a system at any moment in time is defined as the enthalpy of the
system minus the product of the temperature times the entropy of the system.

G = H - TS
The Gibbs free energy of the system is a state function because it is defined in terms of
thermodynamic properties that are state functions. The change in the Gibbs free energy of the
system that occurs during a reaction is therefore equal to the change in the enthalpy of the system
minus the change in the product of the temperature times the entropy of the system.

G = H - (TS)
If the reaction is run at constant temperature, this equation can be written as follows.

G= H-T S
The change in the free energy of a system that occurs during a reaction can be measured
under any set of conditions. If the data are collected under standard-state conditions, the result is
the standard-state free energy of reaction ( Go).

Go = Ho - T So

The beauty of the equation defining the free energy of a system is its ability to determine
the relative importance of the enthalpy and entropy terms as driving forces behind a particular
reaction. The change in the free energy of the system that occurs during a reaction measures the
balance between the two driving forces that determine whether a reaction is spontaneous. As we
have seen, the enthalpy and entropy terms have different sign conventions.
Favorable Unfavorable

Ho < 0 Ho > 0

So > 0 So < 0

The entropy term is therefore subtracted from the enthalpy term when calculating Go for
a reaction.

Because of the way the free energy of the system is defined, Go is negative for any
reaction for which Ho is negative and So is positive. Go is therefore negative for any reaction
that is favored by both the enthalpy and entropy terms. We can therefore conclude that any reaction
for which Go is negative should be favorable, or spontaneous.

Favorable, or spontaneous reactions: Go < 0

Conversely, Go is positive for any reaction for which Ho is positive and So is negative.
Any reaction for which Go is positive is therefore unfavorable.

Unfavorable, or non-spontaneous reactions: Go > 0

Reactions are classified as either exothermic ( H < 0) or endothermic ( H > 0) on the


basis of whether they give off or absorb heat. Reactions can also be classified as exergonic ( G <
0) or endergonic ( G > 0) on the basis of whether the free energy of the system decreases or
increases during the reaction.

When a reaction is favored by both enthalpy ( Ho < 0) and entropy ( So > 0), there is no
need to calculate the value of Go to decide whether the reaction should proceed. The same can
be said for reactions favored by neither enthalpy ( Ho > 0) nor entropy ( So< 0). Free energy
calculations become important for reactions favored by only one of these factors.

The Effect of Temperature on the Free Energy of a Reaction

The balance between the contributions from the enthalpy and entropy terms to the free
energy of a reaction depends on the temperature at which the reaction is run. The equation used to
define free energy suggests that the entropy term will become more important as the temperature
increases.
Go = Ho - T So
Since the entropy term is unfavorable, the reaction should become less favorable as the
temperature increases.

Standard-State Free Energies of Reaction

Go for a reaction can be calculated from tabulated standard-state free energy data. Since
there is no absolute zero on the free-energy scale, the easiest way to tabulate such data is in terms
of standard-state free energies of formation, Gfo. As might be expected, the standard-state free
energy of formation of a substance is the difference between the free energy of the substance and
the free energies of its elements in their thermodynamically most stable states at 1 atm, all
measurements being made under standard-state conditions.

Interpreting Standard-State Free Energy of Reaction Data

We are now ready to ask the obvious question: What does the value of Go tell us about
the following reaction?

N2(g) + 3 H2(g) 2 NH3(g) Go = -32.96 kJ

By definition, the value of Go for a reaction measures the difference between the free
energies of the reactants and products when all components of the reaction are present at standard-
state conditions.

Go therefore describes this reaction only when all three components are present at 1 atm
pressure.

The sign of Go tells us the direction in which the reaction has to shift to come to
equilibrium. The fact that Go is negative for this reaction at 25oC means that a system under
standard-state conditions at this temperature would have to shift to the right, converting some of
the reactants into products, before it can reach equilibrium. The magnitude of Go for a reaction
tells us how far the standard state is from equilibrium. The larger the value of Go, the further the
reaction has to go to get to from the standard-state conditions to equilibrium.

Assume, for example, that we start with the following reaction under standard-state
conditions, as shown in the figure below.

N2(g) + 3 H2(g) 2 NH3(g)


The value of G at that moment in time will be equal to the standard-state free energy for
this reaction, Go.

When Qp = 1: G= Go

` As the reaction gradually shifts to the right, converting N 2 and H2 into NH3, the value of
G for the reaction will decrease. If we could find some way to harness the tendency of this reaction
to come to equilibrium, we could get the reaction to do work. The free energy of a reaction at any
moment in time is therefore said to be a measure of the energy available to do work.

The Relationship Between Free Energy and Equilibrium Constants

When a reaction leaves the standard state because of a change in the ratio of the
concentrations of the products to the reactants, we have to describe the system in terms of non-
standard-state free energies of reaction. The difference between Go and G for a reaction is
important. There is only one value of Go for a reaction at a given temperature, but there are an
infinite number of possible values of G.

The figure below shows the relationship between G for the following reaction and the
logarithm to the base e of the reaction quotient for the reaction between N2 and H2 to form NH3.

N2(g) + 3 H2(g) 2 NH3(g)


Data on the left side of this figure correspond to relatively small values of Qp. They
therefore describe systems in which there is far more reactant than product. The sign of G for
these systems is negative and the magnitude of G is large. The system is therefore relatively far
from equilibrium and the reaction must shift to the right to reach equilibrium.

Data on the far right side of this figure describe systems in which there is more product
than reactant. The sign of G is now positive and the magnitude of G is moderately large. The
sign of G tells us that the reaction would have to shift to the left to reach equilibrium. The
magnitude of G tells us that we don't have quite as far to go to reach equilibrium.

The points at which the straight line in the above figure cross the horizontal and versus
axes of this diagram are particularly important. The straight line crosses the vertical axis when the
reaction quotient for the system is equal to 1. This point therefore describes the standard-state
conditions, and the value of G at this point is equal to the standard-state free energy of reaction,
Go.

When Qp = 1: G= Go

The point at which the straight line crosses the horizontal axis describes a system for
which G is equal to zero. Because there is no driving force behind the reaction, the system must
be at equilibrium.

When Qp = Kp: G= 0

The relationship between the free energy of reaction at any moment in time ( G) and the
standard-state free energy of reaction ( Go) is described by the following equation.

G = Go + RT ln Q
In this equation, R is the ideal gas constant in units of J/mol-K, T is the temperature in
kelvin, ln represents a logarithm to the base e, and Q is the reaction quotient at that moment in
time.

As we have seen, the driving force behind a chemical reaction is zero ( G = 0) when the
reaction is at equilibrium (Q = K).

0 = Go + RT ln K
We can therefore solve this equation for the relationship between Go and K.

Go = - RT ln K
This equation allows us to calculate the equilibrium constant for any reaction from the
standard-state free energy of reaction, or vice versa.

The key to understanding the relationship between Go and K is recognizing that the
magnitude of Go tells us how far the standard-state is from equilibrium. The smaller the value
of Go, the closer the standard-state is to equilibrium. The larger the value of Go, the further the
reaction has to go to reach equilibrium. The relationship between Go and the equilibrium constant
for a chemical reaction is illustrated by the data in the table below.

Values of Go and K for Common Reactions at 25oC

Reaction Go(kJ) K

2 SO3(g) 2 SO2(g) + O2(g) 141.7 1.4 x 10-25

H2O(l) H+(aq) + OH-(aq) 79.9 1.0 x 10-14

AgCl(s) + H2O Ag+(aq) + Cl-(aq) 55.6 1.8 x 10-10

HOAc(aq) + H2O H+(aq) + OAc-(aq) 27.1 1.8 x 10-5

N2(g) + 3 H2(g) 2 NH3(g) -32.9 5.8 x 105

HCl(aq) + H2O H+(aq) + Cl-(aq) -34.2 1 x 106

Cu2+(aq) + 4 NH3(aq) Cu(NH3)42+(aq) -76.0 2.1 x 1013

Zn(s) + Cu2+(aq) Zn2+(aq) + Cu(s) -211.8 1.4 x 1037


The equilibrium constant for a reaction can be expressed in two ways: Kc and Kp. We can
write equilibrium constant expressions in terms of the partial pressures of the reactants and
products, or in terms of their concentrations in units of moles per liter.

For gas-phase reactions the equilibrium constant obtained from Go is based on the partial
pressures of the gases (Kp). For reactions in solution, the equilibrium constant that comes from the
calculation is based on concentrations (Kc).

The Temperature Dependence of Equilibrium Constants

Equilibrium constants are not strictly constant because they change with temperature. We
are now ready to understand why.

The standard-state free energy of reaction is a measure of how far the standard-state is from
equilibrium.

Go = - RT ln K
But the magnitude of Go depends on the temperature of the reaction.

Go = Ho - T So
As a result, the equilibrium constant must depend on the temperature of the reaction.

A good example of this phenomenon is the reaction in which NO 2dimerizes to form N2O4.

2 NO2(g) N2O4(g)
This reaction is favored by enthalpy because it forms a new bond, which makes the system
more stable. The reaction is not favored by entropy because it leads to a decrease in the disorder
of the system.

NO2 is a brown gas and N2O4 is colorless. We can therefore monitor the extent to which
NO2 dimerizes to form N2O4 by examining the intensity of the brown color in a sealed tube of this
gas. What should happen to the equilibrium between NO2 and N2O4 as the temperature is lowered?

For the sake of argument, let's assume that there is no significant change in either Ho or
So as the system is cooled. The contribution to the free energy of the reaction from the enthalpy
term is therefore constant, but the contribution from the entropy term becomes smaller as the
temperature is lowered.

Go = Ho - T So
As the tube is cooled, and the entropy term becomes less important, the net effect is a shift
in the equilibrium toward the right. The figure below shows what happens to the intensity of the
brown color when a sealed tube containing NO2 gas is immersed in liquid nitrogen. There is a
drastic decrease in the amount of NO2 in the tube as it is cooled to -196oC.

The Relationship Between Free Energy and Cell Potentials

The value of G for a reaction at any moment in time tells us two things. The sign of
G tells us in what direction the reaction has to shift to reach equilibrium. The magnitude of
G tells us how far the reaction is from equilibrium at that moment.

The potential of an electrochemical cell is a measure of how far an oxidation-reduction


reaction is from equilibrium. The Nernst equation describes the relationship between the cell
potential at any moment in time and the standard-state cell potential.

Let's rearrange this equation as follows.

nFE = nFEo - RT ln Q
We can now compare it with the equation used to describe the relationship between the
free energy of reaction at any moment in time and the standard-state free energy of reaction.

G = Go + RT ln Q

These equations are similar because the Nernst equation is a special case of the more
general free energy relationship. We can convert one of these equations to the other by taking
advantage of the following relationships between the free energy of a reaction and the cell potential
of the reaction when it is run as an electrochemical cell.

G = -nFE
CHAPTER 6: CHEMICAL EQUILIBRIA

Chemical equilibrium, a condition in the course of a reversible chemical reaction in which no net
change in the amounts of reactants and products occurs. A reversible chemical reaction is one in
which the products, as soon as they are formed, react to produce the original reactants.
At equilibrium, the two opposing reactions go on at equal rates, or velocities, hence there is no net
change in the amounts of substances involved. At this point the reaction may be considered to be
completed; i.e., for some specified reaction condition, the maximum conversion of reactants to
products has been attained.

Reversible Reactions

A reversible reaction is a chemical reaction where the reactants form products that, in turn, react
together to give the reactants back. Reversible reactions will reach an equilibrium point where the
concentrations of the reactants and products will no longer change.

A reversible reaction is denoted by a double arrow pointing both directions in a chemical equation.
For example, a two reagent, two product equation would be written as

A+B⇆C+D

Example of a Reversible Reaction

Weak acids and bases may undergo reversible reactions. For example, carbonic acid and water
react this way:

H2CO3 (l) + H2O(l) ⇌ HCO−3 (aq) + H3O+(aq)

Another example of a reversible reaction is:

N2O4 ⇆ 2 NO2

Two chemical reactions occur simultaneously:

N2O4 → 2 NO2

2 NO2 → N2O4

Reversible reactions do not necessarily occur at the same rate in both directions, but they do lead
to an equilibrium condition. If dynamic equilibrium occurs, the product of one reaction is forming
at the same rate as it is used up for the reverse reaction. Equilibrium constants are calculated or
provided to help determine how much reactant and product is formed.
The equilibrium of a reversible reaction depends on the initial concentrations of the reactants and
products and the equilibrium constant, K.

How a Reversible Reaction Works

Most reactions encountered in chemistry are irreversible reactions (or reversible, but with very
little product converting back into reactant). For example, if you burn a piece of wood using the
combustion reaction, you never see the ash spontaneously make new wood, do you? Yet, some
reactions do reverse. How does this work?

The answer has to do with the energy output of each reaction and that required for it to occur. In a
reversible reaction, reacting molecules in a closed system collide with each other and use the
energy to break chemical bonds and form new products. Enough energy is present in the system
for the same process to occur with the products. Bonds are broken and new ones formed, that
happen to result in the initial reactants.

Writing Equilibrium Constant Expressions Involving Solids and Liquids

The equilibrium constant expression is the ratio of the concentrations of a reaction at equilibrium.
Each equilibrium constant expression has a constant value known as K, the equilibrium constant.
When dealing with partial pressures, Kp is used, whereas when dealing with concentrations
(molarity), Kc is employed as the equilibrium constant. Reactions containing pure solids and
liquids results in heterogeneous reactions in which the concentrations of the solids and liquids are
not considered when writing out the equilibrium constant expressions.

The equilibrium constant expression is the ratio of the concentrations of the products over the
reactants
Because K has no units, we relate the concentrations for K as activities. For the following reaction
at equilibrium:
A(aq)+4B(aq)⇌2C(aq)+3D(aq) (1)
The equilibrium constant expression would be:
K=(aC)2(aD)3(aA)(aB)4 (2)
Kc=[C]2[D]3[A][B]4 (3)

Notice how each concentration of product or reactant is raised to the power of its coefficient. For
example, the concentration of D is raised to the power of 3 since it is 3D in the balanced reaction
(eq. 1).
If the reaction involved partial pressures instead of concentrations, then the equilibrium constant,
now Kp, would follow the same formula except with the pressures of the gases written in. In that
case, Kp is only applied to reactions involving gases. Recall the equation relating Kp and Kc:
Kp=Kc(RT)Δn(4)
where
• R is the Ideal Gas Constant (0.0821 L atm mol-1 K-1),
• T is the temperature in Kelvins, and
• Δn is the Sum of Coefficients of gaseous Products - Sum of Coefficients of gaseous
Reactants.

Predicting the direction of a reaction

Often you will know the concentrations of reactants and products for a particular reaction and
want to know whether the system is at equilibrium. If it is not, it is useful to predict how those
concentrations will change as the reaction approaches equilibrium. A useful tool for making such
predictions is the reaction quotient, Q. Q has the same mathematical form as the equilibrium-
constant expression, but Q is a ratio of the actual concentrations (not a ratio of equilibrium
concentrations).
For example, suppose you are interested in the reaction
2SO2(g)+O2(g)⇌2SO3(g)

Kc = [SO3]2[SO2]2[O2] = 245 mol/L (at 1000 K)


and you have added 0.10 mol of each gas to a container with volume 10.0 L. Is the system at
equilibrium? If not, will the concentration of SO3 be greater than or less than 0.010 mol/L when
equilibrium is reached? You can answer these questions by calculating Q and comparing it with Kc.
There are three possibilities:
• If Q = Kc then the actual concentrations of products (and of reactants) are equal to the
equilibrium concentrations and the system is at equilibrium.
• If Q < Kc then the actual concentrations of products are less than the equilibrium
concentrations; the forward reaction will occur and more products will be formed.
• If Q > Kc then the actual concentrations of products are greater than the equilibrium
concentrations; the reverse reaction will occur and more reactants will be formed.

Significance of the Equilibrium Constant

The equilibrium constant, which relates the concentrations of reactants and products,
provides information about the equilibrium mixture as well as the extent to which the reaction has
taken place. Equilibrium constant value can be very large or very small. Usually, it is possible to
determine the feasibility of a reaction just by knowing the magnitude of the equilibrium constant.
In other words, the K value shows how far a reaction proceeds toward completion.

A large K value (K>>1), indicates that the equilibrium position lies more to the right and
the concentrations of the products of the forward reaction are greater than those of the reactants.
On the other hand, if the K value is small (K<<1), the equilibrium position lies to the left, and the
concentrations of the reactants predominate over the products. Aside from indicating to what
extent the reaction has taken place, the equilibrium constant also allows us to calculate the
equilibrium concentrations of products and reactants.
Le Chatelier’s Principle

The study of chemical equilibrium is very important especially in industry. Manufacturers


generally choose the particular reaction and appropriate conditions that would favor the formation
of more products. Hence, aside from knowing the magnitude of K, it is also important to know the
various factors that may affect equilibrium as well as the amount of products formed. A chemical
system in a state of equilibrium remains in that state until it is disturbed by a change in condition.
How systems in equilibrium respond to disturbances or various changes in external conditions was
put forward by a French industrial chemist, Henri-Louis Le Chatelier.

Le Chatelier’s Principle states that of a stress or disturbance is applied to a system in


equilibrium, its tendency is to shift in a direction that best reduces the stress so that equilibrium
will be re-established. Disturbing an equilibrium due to changes in conditions may cause the
equilibrium to shift to either to the right or the left. Le Chatelier’s principle allows qualitative
predictions about the response of an equilibrium system to various stress such as a) concentration
changes, b)pressure and volume changes, c) temperature changes, and d) presence of catalyst.
CHAPTER 7: ACID-BASE EQUILIBRIA AND SALT EQUILIBRIA

Bronsted-Lowry Acids and Bases

Brønsted-Lowry theory of acid and bases took the Arrhenius definition one step further, as
a substance no longer needed to be composed of hydrogen (H+) or hydroxide (OH-) ions in order
to be classified as an acid or base. For example, consider the following chemical equation:
𝑯𝑪𝒍 (𝒂𝒒) + 𝑵𝑯𝟑 (𝒂𝒒) → 𝑵𝑯𝟒+ (𝒂𝒒) + 𝑪𝒍− (𝒂𝒒)
Here, hydrochloric acid (HCl) "donates" a proton (H+) to ammonia (NH3) which "accepts"
it, forming a positively charged ammonium ion (NH4+) and a negatively charged chloride ion (Cl-
). Therefore, HCl is a Brønsted-Lowry acid (donates a proton) while the ammonia is a Brønsted-
Lowry base (accepts a proton). Also, Cl- is called the conjugate base of the acid HCl and NH4+is
called the conjugate acid of the base NH3.
• A Brønsted-Lowry acid is a proton (hydrogen ion) donor.
• A Brønsted-Lowry base is a proton (hydrogen ion) acceptor.

For a reaction to be in equilibrium a transfer of electrons needs to occur. The acid will give
an electron away and the base will receive the electron. Acids and Bases that work together in this
fashion are called a conjugate pair made up of conjugate acids and conjugate bases.
𝐻𝐴 + 𝑍 ⇌ 𝐴− + 𝐻𝑍 +
A stands for an Acidic compound and Z stands for a Basic compound
• A Donates H to form HZ+.
• Z Accepts H from A which forms HZ+
• A- becomes conjugate base of HA and in the reverse reaction it accepts a H from HZ to
recreate HA in order to remain in equilibrium
• HZ+ becomes a conjugate acid of Z and in the reverse reaction it donates a H to
A- recreating Z in order to remain in equilibrium

Acid Base Properties of Water


Water (H2O) is an interesting compound in many respects. Here, we will consider its ability
to behave as an acid or a base.
In some circumstances, a water molecule will accept a proton and thus act as a Brønsted-Lowry
base. We saw an example in the dissolving of HCl in H2O:
𝐻𝐶𝑙 + 𝐻2 𝑂 (𝑙) → 𝐻3 𝑂+ (𝑎𝑞) + 𝐶𝑙 − (𝑎𝑞)
In other circumstances, a water molecule can donate a proton and thus act as a Brønsted-
Lowry acid. For example, in the presence of the amide ion, a water molecule donates a proton,
making ammonia as a product:
𝐻2 𝑂 (𝑙) + 𝑁𝐻2 (𝑎𝑞) → 𝑂𝐻− (𝑎𝑞) + 𝑁𝐻3 (𝑎𝑞)
In this case, NH2− is a Brønsted-Lowry base (the proton acceptor).
So, depending on the circumstances, H2O can act as either a Brønsted-Lowry acid or a
Brønsted-Lowry base. Water is not the only substance that can react as an acid in some cases or a
base in others, but it is certainly the most common example—and the most important one. A
substance that can either donate or accept a proton, depending on the circumstances, is called an
amphiprotic compound.

Auto-Ionization of Water
Because of its amphoteric nature (i.e., acts as both an acid or a base), water does not always
remain as H2OH2O molecules. In fact, two water molecules react to form hydronium and
hydroxide ions:
2𝐻2 𝑂 (𝑙) ⇌ 𝐻3 𝑂+ (𝑎𝑞) + 𝑂𝐻− (𝑎𝑞)

This is also called the self-ionization of water. The concentration of 𝐻3 𝑂+ and 𝑂𝐻 − are
equal in pure water because of the 1:1 stoichiometric ratio of the equation above. The molarity of
𝐻3 𝑂+ and 𝑂𝐻 − in water are also both 1.0 𝑥 10−7 𝑀 at 25° C. Therefore, a constant of water (Kw)
is created to show the equilibrium condition for the self-ionization of water. The product of the
molarity of hydronium and hydroxide ion is always 1.0 𝑥 10−7 (at room temperature).
𝐾𝑤 = [𝐻3 𝑂+ ][𝑂𝐻− ] = 1.0 𝑥 10−14
It also applies to all aqueous solutions. However, Kw does change at different
temperatures, which affects the pH range discussed below.

H+H+ and H3O+H3O+ is often used interchangeably to represent the hydrated proton, commonly
call the hydronium ion.

The equation for water equilibrium is:


𝐻2 𝑂 ⇌ 𝐻 + + 𝑂𝐻−
• If an acid (𝐻 ) is added to the water, the equilibrium shifts to the left and the 𝑂𝐻 − ion
+

concentration decreases
• If base ( 𝑂𝐻 − ) is added to water, the equilibrium shifts to left and the 𝐻 + concentration
decreases.

The pH Scale
Because the constant of water, Kw is always 1.0 X 10-14 (at 25° C), the pKw is 14, the
constant of water determines the range of the pH scale. To understand what the pK w is, it is
important to understand first what the "p" means in pOH and pH. The Danish biochemist Søren
Sørenson proposed the term pH to refer to the "potential of hydrogen ion." He defined the "p" as
the negative of the logarithm, -log, of [𝐻 + ]. Therefore, the pH is the negative logarithm of the
molarity of H. The pOH is the negative logarithm of the molarity of OH - and the pKw is the
negative logarithm of the constant of water. These definitions give the following equations:
𝑝𝐻 = −𝑙𝑜𝑔 [𝐻+ ]
𝑝𝑂𝐻 = −𝑙𝑜𝑔 [𝑂𝐻− ]
𝑝𝐾𝑤 = −𝑙𝑜𝑔 [𝐾𝑤 ]

EXAMPLE:
a. Find the pH of a solution of 0.002 M of HCl.
b. Find the pH of a solution of 0.00005 M NaOH.

SOLUTION
a. The equation for pH is -log [H+]
[𝐻+ ] = 2.0 𝑥 10−3 𝑀
𝑝𝐻 = − 𝑙𝑜𝑔 [2.0 𝑥 10−3 ] = 2.70

b. The equation for pOH is -log [OH-]


[𝑂𝐻 − ] = 5.0 𝑥 10−5 𝑀
𝑝𝑂𝐻 = − 𝑙𝑜𝑔 [5.0 𝑥 10−5 ] = 4.30
𝑝𝐾𝑤 = 𝑝𝐻 + 𝑝𝑂𝐻
and
𝑝𝐻 = 𝑝𝐾𝑤 − 𝑝𝑂𝐻
then
𝑝𝐻 = 14 − 4.30 = 9.70

Strength of Acids and Bases

We can rank the strengths of acids by the extent to which they ionize in aqueous solution.
The reaction of an acid with water is given by the general expression:

𝐻𝐴 (𝑎𝑞) + 𝐻2 𝑂 (𝑙) ⇌ 𝐻3 𝑂+ (𝑎𝑞) + 𝐴− (𝑎𝑞)

Water is the base that reacts with the acid HA, A− is the conjugate base of the acid HA, and
the hydronium ion is the conjugate acid of water. A strong acid yields 100% (or very nearly so)
of and A− when the acid ionizes in water; Figure 1 lists several strong acids. A weak acid
gives small amounts of and A−.

Figure 1. Some of the common strong acids and bases are listed here.

The relative strengths of acids may be determined by measuring their equilibrium constants
in aqueous solutions. In solutions of the same concentration, stronger acids ionize to a greater
extent, and so yield higher concentrations of hydronium ions than do weaker acids. The
equilibrium constant for an acid is called the acid-ionization constant, Ka. For the reaction of an
acid HA:

𝐻𝐴 (𝑎𝑞) + 𝐻2 𝑂 (𝑙) ⇌ 𝐻3 𝑂+ (𝑎𝑞) + 𝐴− (𝑎𝑞)

we write the equation for the ionization constant as:

[𝐻3 𝑂+ ][𝐴− ]
𝐾𝑎 =
[𝐻𝐴]

where the concentrations are those at equilibrium. Although water is a reactant in the
reaction, it is the solvent as well, so we do not include [H 2O] in the equation. The larger the Ka of
an acid, the larger the concentration of and A− relative to the concentration of the non
ionized acid, HA. Thus, a stronger acid has a larger ionization constant than does a weaker acid.
The ionization constants increase as the strengths of the acids increase.

Dissociation Constant of Weak Bases

Consider a hypothetical base B. A weak base partially ionizes in solution and the equation
is shown below. Since the forward reaction is not completed, a double arrow is used for the
backward reaction.

𝐵 + 𝐻𝑂𝐻 ⇄ 𝐻𝐵 + 𝑂𝐻−

The equilibrium expression for this is the same as the equilibrium expression for the weak
acid and is shown below. Instead of Ka, the ionization constant becomes Kb, ionization constant
for weak bases.

[𝐻𝐵][𝑂𝐻− ]
𝐾𝑏 =
[𝐵]

For NH3, the dissociation constant is 1.8 x 10-5. This is a measure of the ability of NH3 to
accept proton. The equation below.

𝑁𝐻3 + 𝐻2 𝑂 ⇆ 𝑁𝐻4+ + 𝑂𝐻 −

Another weak base is CH3NH2 and the equation is shown below.

𝐶𝐻3 𝑁𝐻2 + 𝐻2 𝑂 ⇆ 𝐶𝐻3 𝑁𝐻3+ + 𝑂𝐻 −

The dissociation constant expression for the two weak bases are given below.

[𝑁𝐻4+ ][𝑂𝐻 − ] [𝐶𝐻3 𝑁𝐻3+ ][𝑂𝐻 − ]


𝐾𝑏 = 𝐾𝑏 =
[𝑁𝐻3 ] [𝐶𝐻3 𝑁𝐻2 ]

Common Ion Effect

The solubility products Ksp's are equilibrium constants in hetergeneous equilibria (i.e., between
two different phases). If several salts are present in a system, they all ionize in the solution. If the
salts contain a common cation or anion, these salts contribute to the concentration of the common
ion. Contributions from all salts must be included in the calculation of concentration of the
common ion. For example, a solution containing sodium chloride and potassium chloride will have
the following relationship:
[Na+] + [K+] = [Cl−]
Consideration of charge balance or mass balance or both leads to the same conclusion.
Common Ions
When NaCl and KCl are dissolved in the same solution, the Cl− ions are common to both salts. In
a system containing NaCl and KCl, the Cl− ions are common ions.
NaCl ⇌ Na+ + Cl−
KCl ⇌ K+ + Cl−

For example, when AgCl is dissolved into a solution already


containing NaCl (actually Na+ and Cl− ions), the Cl− ions come from the ionization of
both AgCl and NaCl. Thus, [Cl−] differs from [Ag+]. The following examples show how the
concentration of the common ion is calculated.

EXAMPLE:
What are [Na+], [Cl−], [Ca2+], and [H+] in a solution containing 0.10 M each of NaCl, CaCl2,
and HCl?

SOLUTION
Due to the conservation of ions, we have
[Na+] = [Ca2+] = [H+] = 0.10M
but
[Cl−]= 0.10 (due to NaCl)
+ 0.20 (due to CaCl2)
+ 0.10 (due to HCl)
= 0.40M

Buffers
A buffer is a solution that can resist pH change upon the addition of an acidic or basic
components. It is able to neutralize small amounts of added acid or base, thus maintaining the pH
of the solution relatively stable. This is important for processes and/or reactions which require
specific and stable pH ranges. Buffer solutions have a working pH range and capacity which dictate
how much acid/base can be neutralized before pH changes, and the amount by which it will change.
What is a buffer composed of?
To effectively maintain a pH range, a buffer must consist of a weak conjugate acid-base pair,
meaning either a. a weak acid and its conjugate base, or b. a weak base and its conjugate acid. The
use of one or the other will simply depend upon the desired pH when preparing the buffer. For
example, the following could function as buffers when together in solution:
• Acetic acid (weak organic acid w/ formula CH3COOH) and a salt containing its conjugate
base, the acetate anion (CH3COO-), such as sodium acetate (CH3COONa)
• Pyridine (weak base w/ formula C5H5N) and a salt containing its conjugate acid, the
pyridinium cation (C5H5NH+), such as Pyridinium Chloride.
• Ammonia (weak base w/ formula NH3) and a salt containing its conjugate acid, the
ammonium cation, such as Ammonium Hydroxide (NH4OH)

How does a buffer work?


A buffer is able to resist pH change because the two components (conjugate acid and
conjugate base) are both present in appreciable amounts at equilibrium and are able to neutralize
small amounts of other acids and bases (in the form of H3O+ and OH-) when the are added to the
solution. To clarify this effect, we can consider the simple example of a Hydrofluoric Acid (HF)
and Sodium Fluoride (NaF) buffer. Hydrofluoric acid is a weak acid due to the strong attraction
between the relatively small F- ion and solvated protons (H3O+), which does not allow it to
dissociate completely in water. Therefore, if we obtain HF in an aqueous solution, we establish the
following equilibrium with only slight dissociation (Ka(HF) = 6.6x10-4, strongly favors reactants):
HF(aq)+H2O(l)⇌F−(aq)+H3O+(aq)
We can then add and dissolve sodium fluoride into the solution and mix the two until we
reach the desired volume and pH at which we want to buffer. When Sodium Fluoride dissolves in
water, the reaction goes to completion, thus we obtain:
NaF(aq)+H2O(l)→Na+(aq)+F−(aq)
Since Na+ is the conjugate of a strong base, it will have no effect on the pH or reactivity
of the buffer. The addition of NaFNaF to the solution will, however, increase the concentration of
F- in the buffer solution, and, consequently, by Le Châtelier’s Principle, lead to slightly less
dissociation of the HF in the previous equilibrium, as well. The presence of significant amounts of
both the conjugate acid, HFHF, and the conjugate base, F-, allows the solution to function as a
buffer. This buffering action can be seen in the titration curve of a buffer solution.
As we can see, over the working range of the buffer. pH changes very little with the
addition of acid or base. Once the buffering capacity is exceeded the rate of pH change quickly
jumps. This occurs because the conjugate acid or base has been depleted through neutralization.
This principle implies that a larger amount of conjugate acid or base will have a greater buffering
capacity.
If acid were added:
F−(aq)+H3O+(aq)⇌HF(aq)+H2O(l)
In this reaction, the conjugate base, F-, will neutralize the added acid, H3O+, and this
reaction goes to completion, because the reaction of F- with H3O+ has an equilibrium constant
much greater than one. (In fact, the equilibrium constant the reaction as written is just the inverse
of the Ka for HF: 1/Ka(HF) = 1/(6.6x10-4) = 1.5x10+3.) So long as there is more F- than H3O+,
almost all of the H3O+ will be consumed and the equilibrium will shift to the right, slightly
increasing the concentration of HF and slightly decreasing the concentration of F-, but resulting in
hardly any change in the amount of H3O+ present once equilibrium is re-established.
If base were added:
HF(aq)+OH−(aq)⇌F−(aq)+H2O(l)
In this reaction, the conjugate acid, HF, will neutralize added amounts of base, OH-, and
the equilibrium will again shift to the right, slightly increasing the concentration of F- in the
solution and decreasing the amount of HF slightly. Again, since most of the OH- is neutralized,
little pH change will occur.
These two reactions can continue to alternate back and forth with little pH change.
Selecting proper components for desired pH
Buffers function best when the pKa of the conjugate weak acid used is close to the desired
working range of the buffer. This turns out to be the case when the concentrations of the conjugate
acid and conjugate base are approximately equal (within about a factor of 10). For example, we
know the Ka for hydroflouric acid is 6.6 x 10-4 so its pKa= -log(6.6 x 10-4) = 3.18. So, a
hydrofluoric acid buffer would work best in a buffer range of around pH = 3.18.
For the weak base ammonia (NH3), the value of Kb is 1.8x10-5, implying that the Ka for
the dissociation of its conjugate acid, NH4+, is Kw/Kb=10-14/1.8x10-5 = 5.6x10-10. Thus, the
pKa for NH4+ = 9.25, so buffers using NH4+/NH3 will work best around a pH of 9.25. (It's always
the pKaof the conjugate acid that determines the approximate pH for a buffer system, though this
is dependent on the pKb of the conjugate base, obviously.)
When the desired pH of a buffer solution is near the pKa of the conjugate acid being used
(i.e., when the amounts of conjugate acid and conjugate base in solution are within about a factor
of 10 of each other), the Henderson-Hasselbalch equation can be applied as a simple
approximation of the solution pH, as we will see in the next section.

S-ar putea să vă placă și