Sunteți pe pagina 1din 44

Thermochemistry

and
Chemical Thermodynamics
General Chemistry II
Jose Emmanuel P. Chico
Contents

Thermochemistry
Energy Changes in Chemical Reactions
 Energy
 Types of Energy
 Exothermic Reactions
 Endothermic Reactions
 Systems and Surroundings
First Law of Thermodynamics
 First Law of Thermodynamics
 Heat
 Work
Enthalpy of Chemical Reactions
 Enthalpy
 Enthalpy change in Reactions
 Thermochemical Equations
Contents

Thermochemistry
Calorimetry
 Heat Capacity
 Molar and Specific Heat Capacities
 Heat, Enthalpy, Temperature
 Specific heat and latent heat
 The bomb calorimeter
 Constant-pressure calorimetry
Standard enthalpy of formation
 Standard enthalpy of formation
 Hess’s Law
 Heat of Solution
 Heat of dilution
Contents

Chemical Thermodynamics
Spontaneous and Nonspontaneous Processes
 Spontaneous and Nonspontaneous Processes
 Dispersal of Matter and Energy
 The Reverse Process
Entropy
 Entropy and efficiency limits
 Entropy and heat death
 Entropy and the arrow of time
 Entropy at exact differential
The Second Law of Thermodynamics
Gibbs free energy and Chemical
Equilibrium
Energy Changes in Chemical Reactions
Energy
Energy, in physics, the capacity for doing work. It may exist in potential, kinetic, thermal,
electrical, chemical, nuclear, or other various forms. There are, moreover, heat and work—i.e.,
energy in the process of transfer from one body to another. After it has been transferred, energy is
always designated according to its nature. Energy is the ability to do work or transfer heat. Energy
is used to do every task, from cooking a meal to playing volleyball to launching spacecraft into
space.

Figure 1 Energy

Types of energy
There are two main types of energy- potential energy and kinetic energy. All forms
of energy are associated with motion. For example, any given body has kinetic energy if it
is in motion. A tensioned device such as a bow or spring, though at rest, has the potential
for creating motion; it contains potential energy because of its configuration. Similarly,
nuclear energy is potential energy because it results from the configuration of subatomic
particles in the nucleus of an atom.

Figure 2 Thermal Energy Figure 3 Chemical Energy Figure 4 Electrical Energy


Figure 6 Radiant Energy
Figure 7 Gravitational Energy Figure 5 Mechanical Energy

Table 1 Types of Energy

Potential energy Kinetic energy


Potential energy is stored energy and the Kinetic energy is the motion of waves, electrons,
energy of position. atoms, molecules, substances, and objects.
Chemical energy is energy stored in the bonds Radiant energy is electromagnetic energy that
of atoms and molecules. Batteries, biomass, travels in transverse waves. Radiant energy includes
petroleum, natural gas, and coal are examples visible light, x-rays, gamma rays, and radio waves.
of chemical energy. Chemical energy is Light is one type of radiant energy. Sunshine is
converted to thermal energy when people burn radiant energy, which provides the fuel and warmth
wood in a fireplace or burn gasoline in a car's that make life on earth possible.
engine.
Thermal energy, or heat, is the energy that comes
Mechanical energy is energy stored in objects from the movement of atoms and molecules in a
substance. Heat increases with increases in the
by tension. Compressed springs and stretched
speed that these particles move. Geothermal energy
rubber bands are examples of stored
is the thermal energy in the earth.
mechanical energy.
Motion energy is energy stored in the movement of
Nuclear energy is energy stored in the nucleus objects. The faster they move, the more energy is
of an atom—the energy that holds the nucleus stored. It takes energy to get an object moving, and
together. Large amounts of energy can be energy is released when an object slows down. Wind
released when the nuclei are combined or split is an example of motion energy. A dramatic example
apart. of motion energy is a car crash—a car comes to a
total stop and releases all of its motion energy at
Gravitational energy is energy stored in an once in an uncontrolled instant.
object's height. The higher and heavier the
object, the more gravitational energy is stored. Sound is the movement of energy through
When a person rides a bicycle down a steep hill substances in longitudinal (compression/rarefaction)
and picks up speed, the gravitational energy is waves. Sound is produced when a force causes an
converting to motion energy. Hydropower is object or substance to vibrate. The energy is
another example of gravitational energy, where transferred through the substance in a wave.
gravity forces water down through a Typically, the energy in sound is smaller than in other
hydroelectric turbine to produce electricity. forms of energy.

Electrical energy is delivered by tiny charged


particles called electrons, typically moving through a
wire. Lightning is an example of electrical energy in
nature.
Energy Changes in Chemical
Reactions
Almost all chemical reactions absorb or produce (release)
energy, generally in the form of heat. It is important to understand the
distinction between thermal energy and heat. Heat is the transfer of
thermal energy between two bodies that are at different
temperatures. Thus, we often speak of the ―heat flow‖ from a hot
object to a cold one. Although the term ―heat‖ by itself implies the
transfer of energy, we customarily talk of ―heat absorbed‖ or ―heat
released‖ when describing the energy changes that occur during a
process. Thermochemistry is the study of heat change in chemical
reactions. Figure 9 Global Heat flow

Due to the absorption of energy when chemical bonds are


broken, and the release of energy when chemical bonds are formed,
chemical reactions almost always involve a change in energy
between products and reactants. By the Law of Conservation of
Energy, however, we know that the total energy of a system must
remain unchanged, and that oftentimes a chemical reaction will
absorb or release energy in the form of heat, light, or both. The
energy change in a chemical reaction is due to the difference in the
amounts of stored chemical energy between the products and the
reactants. This stored chemical energy, or heat content, of the
system is known as its enthalpy.
Figure 8 Infrared Image showing heat

Exothermic Reactions
Exothermic reactions release heat and light into their surroundings. For example, combustion
reactions are usually exothermic. In exothermic reactions, the products have less enthalpy than the
reactants, and as a result, an exothermic reaction is said to have a negative enthalpy of reaction. This
means that the energy required to break the bonds in the reactants is less than the energy released
when new bonds form in the products. Excess energy from the reaction is released as heat and light..
Examples of Exothermic Reactions
•any combustion reaction
•a neutralization reaction
•rusting of iron (rust steel wool with vinegar)
•the thermite reaction
•dissolving laundry detergent in water
•adding water to anhydrous copper(II) sulfate
•freezing water into ice cubes
Figure 12 Freezing water into cubes
•snow forming inside clouds
•respiration
•rain forming from water vapor in clouds
•formation of ion pairs
•making a gas molecule from atoms
•dehydrating sugar with sulfuric acid
•a burning candle
•nuclear fission

Figure 13 Nuclear fission

Figure 11 Thermite reaction Figure 10 Burning candle


Endothermic Reactions
Endothermic reactions, on the other hand, absorb heat and/or light from their surroundings. For
example, decomposition reactions are usually endothermic. In endothermic reactions, the products
have more enthalpy than the reactants. Thus, an endothermic reaction is said to have a positive
enthalpy of reaction. This means that the energy required to break the bonds in the reactants is more
than the energy released when new bonds form in the products; in other words, the reaction requires
energy to proceed.

Examples of endothermic reactions:

•Photosynthesis: As a tree grows, it absorbs energy from the environment to break apart CO2
and H2O.

•Evaporation: Sweating cools a person down as water draws heat to change into gas form.

•Cooking an egg: Energy is absorbed from the pan to cook the egg.

Figure 15 Endothermic reaction Figure 14 Photosynthesis

Identifying Endothermic and Exothermic Reactions


There are two methods for distinguishing between exothermic and endothermic reactions.

1. Monitor temperature change

When energy is released in an exothermic reaction, the temperature of the reaction mixture
increases. When energy is absorbed in an endothermic reaction, the temperature decreases. You can
monitor changes in temperature by placing a thermometer in the reaction mixture
2. Calculate the enthalpy of reaction (ΔH)

To classify the net energy output or input of chemical reactions, you can calculate something
called the enthalpy change (ΔH) or heat of reaction, which compares the energy of the reactants with
the energy of the products. Enthalpy is a measure of internal energy. So, when you calculate the
difference between the enthalpy of the products and the enthalpy of the reactants, you find the
enthalpy change (ΔH), which can be represented mathematically as: ΔH = energy used in reactant
bond breaking + energy released in product bond making . If ΔH is negative (−) then the chemical
reaction is exothermic, because more energy is released when the products are formed than energy is
used to break up the reactants. If ΔH is positive (+) then the chemical reaction is endothermic, because
less energy is released when the products are formed than the energy is used to break up the
reactants.

Figure 16 Exothermic Reaction

Systems and Surroundings


The universe is made up of systems and surroundings. When energy is transferred, it is
necessary to determine what is losing the energy and what is gaining it. The system is the part of the
universe, the chosen portion of space that is under consideration. The system usually includes the
reactants and products of a chemical reaction or a single material that undergoes physical change like
melting or freezing. The surroundings are everything else in the universe around the system. The
surroundings usually include the container that holds the reactants and products, and the rest of the
universe
Figure 18 System and Surroundings

The system can be classified as open, closed, or isolated. Open systems allows flow and
interaction of matter and energy across its boundaries. A closed system does not allow matter to pass
through the boundaries, but energy can pass and interact through them. An isolated system has no
interchange of matter and energy within its surroundings.

Figure 17 Types of System


First Law of Thermodynamics
Introduction to Thermodynamics
Thermochemistry is part of a broader subject called
thermodynamics, which is the scientific study of the
interconversion of heat and other kinds of energy. The laws of
thermodynamics provide useful guidelines for understanding the
energetics and directions of processes.

Figure 20 Thermodynamics

In thermodynamics, we study changes in the state of a system, which is defined by the values of
all relevant macroscopic properties, for example, composition, energy, temperature, pressure, and
volume. Energy, pressure, volume, and temperature are said to be state functions— properties that
are determined by the state of the system, regardless of how that condition was achieved. In other
words, when the state of a system changes, the magnitude of change in any state function depends
only on the initial and final states of the system and not on how the change is accomplished.

Figure 19 State functions


First Law of Thermodynamics

Figure 21 First Law of Thermodynamics

The First Law of Thermodynamics states that energy can be converted from one form to another
with the interaction of heat, work and internal energy, but it cannot be created nor destroyed, under any
circumstances. Mathematically, this is represented as

ΔU=q+w

• ΔU is the total change in internal energy of a system,

• q is the heat exchanged between a system and its surroundings, and

• w is the work done by or on the system.

Work is also equal to the negative external pressure on the system multiplied by the change in
volume:

w=−pΔV

where P is the external pressure on the system, and ΔV is the change in volume. This is
specifically called "pressure-volume" work.

Equation says that the change in the internal energy, ∆E, of a system is the sum of the heat
exchange q between the system and the surroundings and the work done w on (or by) the system. The
sign conventions for q and w are as follows: q is positive for an endothermic process and negative for
an exothermic process and w is positive for work done on the system by the surroundings and negative
for work done by the system on the surroundings.

The transfers and transformations of energy take place around us all the time. For instance, light
bulbs transform electrical energy into light energy, and gas stoves transform chemical energy from
natural gas into heat energy. Plants perform one of the most biologically useful transformations of
energy on Earth: they convert the energy of sunlight into the chemical energy stored within organic
molecules.
Figure 23 Energy Tranformations

The exact energy of any system cannot be determined but the change in internal energy when a
system undergoes transformation from state 1 to state 2 can be determined. The internal energy of a
system refers to the sum of all possible forms of energy in the system.

There are two possible types of transformation that a system may undergo:

a. A physical change

b. A chemical change

The change in energy, ∆E (read as delta E), is defined as the change in the total energy of the
system as is transforms from one state to another. The equation is written as:

∆E = Efinal – Einitial

For a chemical reaction, the initial energy is the energy of the reactants and the final energy is
that of the products. If ∆E is positive, it means that the system has absorbed energy. If ∆E is negative,
it means that the system has released energy.

Figure 22 Physical and Chemical change


Heat and Work
Heat
When energy is exchanged between thermodynamic systems by thermal interaction, the
transfer of energy is called heat. The units of heat are therefore the units of energy, or joules (J). Heat
is transferred by conduction, convection, and/or radiation.

Heat is transfer by conduction occurs when an object with high thermal energy comes into
contact with an object with low thermal energy. Heat transfer by convection occurs through a medium.
For example, when heat transfers from the hot water at the bottom of the pot to the cooler water at the
top of the pot. Lastly, heat can also be transferred by radiation; a hot object can convey heat to
anything in its surroundings via electromagnetic radiation.

When a high temperature body is brought into contact with a low temperature body, the
temperatures equilibrate: there is heat flow from higher to lower temperature, like water flowing
downhill, until the temperatures of the bodies are equivalent. The high temperature body loses thermal
energy, and the low temperature body acquires this same amount of thermal energy. The system is
then said to be at thermal equilibrium.

Figure 24 Heat transfer

An illustration of thermal equilibrium: The can of cola and ice cube start at different
temperatures. When they come into contact, heat is transferred from the cola can to the ice cube until
both bodies reach thermal equilibrium.
Work
Work is the transfer of energy by any process other than
heat. Like heat, the unit measurement for work is joules (J).
There are many forms of work, including but not limited to
mechanical, electrical, and gravitational work. For our purposes,
we are concerned with P-V work, which is the work done in an
enclosed chemical system. In this type of system, work is
defined as the change in the volume (V) in liters within the
system multiplied by a pressure (P). Assuming the system is at
constant pressure, this equates to the following:

W=PΔV

Most often, we are interested in the work done by


expanding gases. Assuming the gases are ideal, we can apply
the ideal gas law to the above equation to get the following:
Figure 25 Work
W=PΔV=nRΔT

Relationship Between Heat and Work


Heat and work are related. Work can be completely converted into heat, but the reverse is not
true: heat energy cannot be wholly transformed into work energy. Scientists and engineers have been
able to exploit the principles of thermochemistry to develop technologies ranging from hot/cold packs to
gasoline powered combustion engines.

For a closed system, the change in internal energy (∆U) is related to heat (Q) and work (W) as
follows:

ΔU=Q+W

This means that the total energy within a system is affected by the sum of two possible energy
transfers: heat and work.
Table 2 sign Conventions

Sign Conventions for work and heat


Work done by the system on the _
surroundings
Work done on the system by the surrounding +
Heat absorbed by the system from the +
surrounding
Heat absorbed by the surroundings from the -
system

EXAMPLE

A certain gas expands in volume from 2.0 L to 6.0 L at constant temperature. Calculate the
work done by the gas if it expands (a) against a vacuum and (b) against a constant pressure of 1.2
atm. Strategy A simple sketch of the situation is helpful here:

The work done in gas expansion is equal to the product of the external, opposing pressure and
the change in volume. What is the conversion factor between L ? atm and J? Solution (a) Because
the external pressure is zero, no work is done in the expansion.

w =P∆V

=(0)(6.0 2 2.0) L

=0

(b) The external, opposing pressure is 1.2 atm, so

w =P∆V

=(1.2 atm)(6.0 2 2.0) L

=4.8 L-atm

To convert the answer to joules, we write

w =4.8 L-atm x101.3 J /1 L-atm

=4.9 x 102 J
Enthalpy of chemical reactions
Enthalpy
Enthalpy (signified as H) is a measure of the total energy of a system and often expresses and
simplifies energy transfer between systems. Since the total enthalpy of a system cannot be measured
directly, we most often refer to the change in enthalpy for a particular chemical reaction. At constant
pressure, the change in enthalpy is equal to the heat given off, or the heat absorbed, in a given
chemical reaction:

ΔHqrxn = ∑Hproducts - ∑H¬reactants

Due to this relation, the change in enthalpy, ΔH, is often referred to as the ―heat of reaction.‖

Enthalpy Change in Reactions


Since most of the chemical reactions in laboratory are nothing but the constant-pressure
processes, we can write the change in enthalpy (also known as enthalpy of reaction) for a reaction.
Consider the following general type of reaction

reactant → products

The ∆H is defined as the difference between the enthalpies of products and the reactants. Thus

∆H = Hproducts - Hreactants

The enthalpy of reaction can be positive or negative or zero depending upon whether the heat is
gained or lost or no heat is lost or gained:

∆H > 0, if Hproducts > Hreactants endothermic reaction (∆H is positive (+))

∆H < 0, if Hproducts < Hreactants, exothermic reaction (∆H is negative (-))

∆H = 0, if Hproducts = Hreactants, no heat is lost or gained (∆H is zero)


Figure 26 Enthalpy change
Standard Enthalpy of Reaction

The enthalpy of reaction is defined as the internal energy of the reaction system, plus the
product of pressure and volume. It is given by:

H=U+PV

By adding the PV term, it becomes possible to measure a change in energy within a chemical
system, even when that system does work on its surroundings. Most often, we are interested in the
change in enthalpy of a given reaction, which can be expressed as follows:

ΔH=ΔU+PΔV

When you run a chemical reaction in a laboratory, the reaction occurs at constant pressure,
because the atmospheric pressure around us is relatively constant. We will examine the change in
enthalpy for a reaction at constant pressure, in order to see why enthalpy is such a useful concept for
chemists.

Enthalpy of Reaction at Constant Volume

Let‘s examine the internal energy change, ΔU, at constant volume. At constant volume, ΔV=0,
the equation for the change in internal energy reduces to the following:

ΔU=QVΔ

The subscript V is added to Q to indicate that this is the heat transfer associated with a chemical
process at constant volume. This internal energy is often very difficult to calculate in real life settings,
though, because chemists tend to run their reactions in open flasks and beakers that allow gases to
escape to the atmosphere. Therefore, volume is not held constant and calculating ΔU becomes
problematic. To correct for this, we introduce the concept of enthalpy, which is much more commonly
used by chemists.
Enthalpy of Reaction at Constant Pressure

Let‘s look once again at the change in enthalpy for a given chemical process. It is given as
follows:

ΔH=ΔU+PΔV

However, we also know that:

ΔU=Q−W=Q−PΔV

Substituting to combine these two equations, we have:

ΔH=Q−PΔV+PΔV=QP

Thus, at constant pressure, the change in enthalpy is simply equal to the heat
released/absorbed by the reaction. Due to this relation, the change in enthalpy is often referred to
simply as the ―heat of reaction.‖

Thermochemical Equations
When we write chemical equations to represent chemical reactions, we simply write the
balanced chemical equations. However, within the realm of the thermodynamics, we must write the
chemical equations with change in heat (enthalpy change). Such equations are known as
thermochemical equations. Consider the process of melting the ice, for which we write the following
chemical equation:

H 2O (s) → H2O (l)

This equation simply says that one mole of ice melts and one mole of water forms. But in
reality, the melting process involves absorption of certain amount of heat by the ice. Now we have to
incorporate that heat as a part of the chemical reaction that is indicated by the enthalpy change. When
you do that it becomes a thermochemical equation.

H2O (s) → H2O (l) ∆H = 6.01 kJ/mol

First of all, the ∆H is positive (∆H >0) because it is an endothermic reaction (heat is absorbed by
the ice from the surrounding), and secondly, ∆H is written per mole of reaction. In this example, 1 mole
of ice is converted to 1mole of water.

As another example, consider the oxidation of graphite to produce carbon dioxide gas:

C (graphite) + O2 (g) → CO2 (g) ∆H = - 393.5 kJ/mol

Here the ∆H is negative (∆H <0) because it is an exothermic reaction. It means that when 1
mole of graphite is burned, 393.5 kJ heat is liberated to the surrounding.

The above written equations for the melting of ice and oxidation of graphite are examples of
thermochemical equations that show mass relationship as well as enthalpy change.
There are certain properties of thermochemical equations that you should know:

1. You must always specify the physical state of all reactants and products because it helps to
determine the actual enthalpy change. As you know the chemical formula for water is H2O whether it is
in solid state (ice) or liquid state (water) or gaseous state (vapor). However, the amount of energy
involved going from one state to another state is not the same. The above mentioned 6.01 kJ refers to
the enthalpy change going from the solid state to the liquid sate only.

2. If we multiply both sides of the thermochemical equation by a certain number, then the ∆H
must also be multiplied by the same number. Thus if we multiply the melting of ice equation by 2, then
we also have to multiply the ∆H by 2.

2H2O (s) → 2H2O (l) ∆H = 2 x 6.01 kJ/mol = 12.02 kJ/mol

3. When we reverse the thermochemical equation, we make the reactants to become products
and products to become reactants. In that case, the sign on ∆H must be reversed without changing the
magnitude of ∆H. For examples, freezing of water to ice is represented by

H2O (l) → H2O (s) ∆H = - 6.01 kJ/mol

This now becomes an exothermic reaction. This makes sense because when the ice is formed
the heat is released (lost) to the surrounding by the water. If the water does not lose the heat, it will not
become an ice.

Given the thermochemical equation

2SO 2(g) 1 O2(g) ¡ 2SO3(g) ¢H 52 198.2 kJ/mol

Calculate the heat evolved when 87.9 g of SO 2 (molar mass 5 64.07 g/mol) is converted to SO
3.

-198.2 kJ /2 mol SO2

How many moles of SO 2 are in 87.9 g of SO 2 ? What is the conversion factor between grams
and moles? Solution We need to fi rst calculate the number of moles of SO 2 in 87.9 g of the
compound and then fi nd the number of kilojoules produced from the exothermic reaction. The
sequence of conversions is as follows:

grams of SO2 → moles of SO2 → kilojoules of heat generated

Therefore, the enthalpy change for this reaction is given by

∆H = 87.9 g SO2 x1 mol SO2 /64.07 g SO2 x -198.2 kJ /2 mol SO2

=2136 kJ

and the heat released to the surroundings is 136 kJ.


Calorimetry
Heat Capacity
Heat capacity is an intrinsic physical property of a
substance that measures the amount of heat required to
change that substance‘s temperature by a given amount. In the
International System of Units (SI), heat capacity is expressed
in units of joules per kelvin (J⋅K−1)(J⋅K−1). Heat capacity is an
extensive property, meaning that it is dependent upon the
size/mass of the sample. For instance, a sample containing
twice the amount of substance as another sample would
require twice the amount of heat energy (Q) to achieve the
same change in temperature (ΔT) as that required to change
the temperature of the first sample.
Figure 27 Boiling water

Molar and Specific Heat Capacities


There are two derived quantities that specify heat capacity as an intensive property (i.e.,
independent of the size of a sample) of a substance. They are:

• the molar heat capacity, which is the heat capacity per mole of a pure substance. Molar
heat capacity is often designated CP, to denote heat capacity under constant pressure conditions, as
well as CV, to denote heat capacity under constant volume conditions. Units of molar heat capacity are
JK⋅ mol.

• the specific heat capacity, often simply called specific heat, which is the heat capacity per
unit mass of a pure substance. This is designated cP and cV and its units are given in Jg⋅K.

Heat, Enthalpy, and Temperature


Given the molar heat capacity or the specific heat for a pure substance, it is possible to calculate
the amount of heat required to raise/lower that substance‘s temperature by a given amount. The
following two formulas apply:

q=mcpΔT

q=nCPΔT

In these equations, m is the substance‘s mass in grams (used when calculating with specific
heat), and n is the number of moles of substance (used when calculating with molar heat capacity).
Specific Heat and Latent Heat
Specific heat capacity is the measure of the heat
energy required to raise the temperature of a given
quantity of a substance by one kelvin. Latent heat of
melting describes the amount of heat required to melt a
solid. When a solid is undergoing melting, the
temperature basically remains constant until the entire
solid is molten.

Figure 28 Latent heat

Example

A 466-g sample of water is heated from 8.50°C to 74.60°C. Calculate the amount of heat
absorbed (in kilojoules) by the water. Strategy We know the quantity of water and the specifi c
heat of water. With this information and the temperature rise, we can calculate the amount of
heat absorbed ( q ). Solution Using Equation (6.12), we write

q = ms∆t

=(466 g)(4.184 J/g?°C)(74.60°C 2 8.50°C)

=1.29 x 105 J x 1 kJ /1000 J

=129 kJ

Check The units g and °C cancel, and we are left with the desired unit kJ. Because heat
is absorbed by the water from the surroundings, it has a positive sign. Practice Exercise An
iron bar of mass 869 g cools from 94°C to 5°C. Calculate the heat released (in kilojoules) by the
metal.
The Bomb Calorimeter
Bomb calorimetry is used to measure the heat that a reaction absorbs or releases, and is
practically used to measure the calorie content of food. A bomb calorimeter is a type of constant-
volume calorimeter used to measure a particular reaction‘s heat of combustion. For instance, if
we were interested in determining the heat content of a sushi roll, for example, we would be
looking to find out the number of calories it contains. In order to do this, we would place the sushi
roll in a container referred to as the ―bomb‖, seal it, and then immerse it in the water inside the
calorimeter. Then, we would evacuate all the air out of the bomb before pumping in pure oxygen
gas (O2). After the oxygen is added, a fuse would ignite the sample causing it to combust,
thereby yielding carbon dioxide, gaseous water, and heat. As such, bomb calorimeters are built
to withstand the large pressures produced from the gaseous products in these combustion
reactions.

Figure 29 Bomb calorimeter

Heat of combustion is usually measured by placing a known mass of a compound in a


steel container called a constant-volume bomb calorimeter, which is fi lled with oxygen at about
30 atm of pressure. The closed bomb is immersed in a known amount of water, as shown in
Figure 6.8. The sample is ignited electrically, and the heat produced by the combustion reaction
can be calculated accurately by recording the rise in temperature of the water. The heat given off
by the sample is absorbed by the water and the bomb. The special design of the calorimeter
enables us to assume that no heat (or mass) is lost to the surroundings during the time it takes to
make measurements. Therefore, we can call the bomb and the water in which it is submerged an
isolated system.
A quantity of 1.435 g of naphthalene (C 10 H 8 ), a pungent-smelling substance used in
moth repellents, was burned in a constant-volume bomb calorimeter. Consequently, the
temperature of the water rose from 20.28°C to 25.95°C. If the heat capacity of the bomb plus
water was 10.17 kJ/°C, calculate the heat of combustion of naphthalene on a molar basis; that is,
fi nd the molar heat of combustion. Strategy Knowing the heat capacity and the temperature
rise, how do we calculate the heat absorbed by the calorimeter? What is the heat generated by
the combustion of 1.435 g of naphthalene? What is the conversion factor between grams and
moles of naphthalene? Solution The heat absorbed by the bomb and water is equal to the
product of the heat capacity and the temperature change. From Equation (6.16), assuming no
heat is lost to the surroundings, we write

qcal = Ccal∆t

= (10.17 kJ/°C)(25.95°C 2 20.28°C)

= 57.66 kJ

Because q sys = q cal + q rxn = 0, q cal = - q rxn . The heat change of the
reaction is 257.66 kJ. This is the heat released by the combustion of 1.435 g of C 10 H 8 ;
therefore, we can write the conversion factor as

-57.66 kJ /1.435 g C10H8

The molar mass of naphthalene is 128.2 g, so the heat of combustion of 1 mole of


naphthalene is

molar heat of combustion =-57.66 kJ/ 1.435 g C10H8 x 128.2 g C10H8 /1 mol C10H8

=-5.151 3 103 kJ/mol

Constant-Pressure Calorimetry
A constant-pressure calorimeter measures the
change in enthalpy of a reaction occurring in a liquid
solution. In that case, the gaseous pressure above the
solution remains constant, and we say that the reaction is
occurring under conditions of constant pressure. The heat
transferred to/from the solution in order for the reaction to
occur is equal to the change in enthalpy (ΔH=qP), and a
constant-pressure calorimeter thus measures this heat of
reaction. In contrast, a bomb calorimeter ‗s volume is
constant, so there is no pressure-volume work and the heat
measured relates to the change in internal energy (ΔU=qV).

Figure 30 Coffee-cup calorimeter


A simple example of a constant-
pressure calorimeter is a coffee-cup
calorimeter, which is constructed from two
nested Styrofoam cups and a lid with two
holes, which allows for the insertion of a
thermometer and a stirring rod. The inner cup
holds a known amount of a liquid, usually
water,that absorbs the heat from the reaction.
The outer cup is assumed to be perfectly
adiabatic, meaning that it does not absorb any
heat whatsoever. As such, the outer cup is
assumed to be a perfect insulator.
Figure 31 Coffe-cup calorimeter

Table 3Heats of some typical reactions measured at constant pressure

Type of Reaction ∆H(kJ/mol)


Heat of neutralization -56.2
Heat of ionization 56.2
Heat of fusion 6.01
Heat of evaporation 44.0
Heat of reaction -180.2

Standard Enthalpy of Formation


The standard enthalpy of formation or standard heat of formation, of a compound is the change
in enthalpy that accompanies the formation of one mole of the compound from its elements in their
standard states. For example, the standard enthalpy of formation for carbon dioxide would be the
change in enthalpy for the following reaction:

C(s)(graphite)+O2(g)→CO2(g)ΔH⊖f=−394 kJ/mol
Table 4 Heat of formation values

Hess’s Law
Hess‘s Law sums the changes in enthalpy for a series of intermediate reaction steps to find
the overall change in enthalpy for a reaction.

Hess‘s law is a relationship in physical chemistry named after Germain Hess, a Swiss-born
Russian chemist and physician. This law states that if a reaction takes place in several steps, then
the standard reaction enthalpy for the overall reaction is equal to the sum of the standard
enthalpies of the intermediate reaction steps, assuming each step takes place at the same
temperature.

Hess‘s law derives directly from the law of conservation of energy, as well as its expression
in the first law of thermodynamics. Since enthalpy is a state function, the change in enthalpy
between products and reactants in a chemical system is independent of the pathway taken from
the initial to the final state of the system. Hess‘s law can be used to determine the overall energy
required for a chemical reaction, especially when the reaction can be divided into several
intermediate steps that are individually easier to characterize. Negative enthalpy change for a
reaction indicates exothermic process, while positive enthalpy change corresponds to endothermic
process.
Figure 32 Hess's law

Calculating Standard Enthalpies of Reaction Using Hess‘s Law

C(s){graphite}→C(s){diamond}ΔHrxn=?

Turning graphite into diamond requires extremely high temperatures and pressures, and therefore
is impractical in a laboratory setting. The change in enthalpy for this reaction cannot be determined
experimentally. However, because we know the standard enthalpy change for the oxidation for these two
substances, it is possible to calculate the enthalpy change for this reaction using Hess‘s law. Our
intermediate steps are as follows:

C(s){graphite}+O2(g)→CO2(g)ΔH∘=−393.41 kJ/mol

C(s){diamond}+O2(g)→CO2(g)ΔH∘=−395.40 kJ/mol

In order to get these intermediate reactions to add to our net overall reaction, we need to reverse
the second step. Keep in mind that when reversing reactions using Hess‘s law, the sign of ΔH will
change. Sometimes, you will need to multiply a given reaction intermediate through by an integer. In
such cases, you need always multiply your ΔH value by that same integer. Restating the first equation
and flipping the second equation, we have:

C(s){graphite}+O2(g)→CO2(g)ΔH∘=−393.41 kJ/mol

CO2(g)→C(s){diamond}+O2(g)ΔH∘=+395.40 kJ/mol

Adding these equations together, carbon dioxides and oxygens cancel, leaving us only with our
net equation. By Hess‘s law, we can sum the ΔH values for these intermediate reactions to get our final
value, ΔH∘rxnΔHrxn∘.

C(s){graphite}→C(s){diamond}ΔH∘rxn=1.89 kJ/mol
Figure 34 Hess's law

Heat of Solution
Heat of solution refers to the change in enthalpy when a solute is
dissolved into a solvent.

The heat of solution, also referred to the enthalpy of solution or enthalpy


of dissolution, is the enthalpy change associated with the dissolution of a solute
in a solvent at constant pressure, resulting in infinite dilution. The heat of
solution, like all enthalpy changes, is expressed in kJ/mol for a reaction taking
place at standard conditions (298.15 K and 1 bar).

Figure 33 Solution process for NaCl


Heat of Dilution
When a previously prepared solution is diluted, that is, when more solvent is added to
lower the overall concentration of the solute, additional heat is usually given off or absorbed.
The heat of dilution is the heat change associated with the dilution process. If a certain
solution process is endothermic and the solution is subsequently diluted, more heat will be
absorbed by the same solution from the surroundings. The converse holds true for an
exothermic solution process—more heat will be liberated if additional solvent is added to dilute
the solution. Therefore, always be cautious when working on a dilution procedure in the
laboratory. Because of its highly exothermic heat of dilution, concentrated sulfuric acid (H 2 SO
4 ) poses a particularly hazardous problem if its concentration must be reduced by mixing it with
additional water. Concentrated H 2 SO 4 is composed of 98 percent acid and 2 percent water
by mass. Diluting it with water releases considerable amount of heat to the surroundings. This
process is so exothermic that you must never attempt to dilute the concentrated acid by
adding water to it. The heat generated could cause the acid solution to boil and splatter. The
recommended procedure is to add the concentrated acid slowly to the water (while constantly
stirring).
Chemical Thermodynamics
Spontaneous and Nonspontaneous Processes
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.

Figure 35 Examples of Spontaneous and Non spontaneous process


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.

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.

No intervention is required because these processes are thermodynamically favorable. In other


words, the initial energy is higher than the final energy.

Note how quickly a process occurs has no bearing on whether or not it is spontaneous. It may
take a long time for rust to become obvious, yet when iron is exposed to air the process will occur. A
radioactive isotope may decay instantly or after thousands or millions or even billions of years.

Figure 37 Converting graphite into diamond Figure 36 Radioactive alpha decay


Nonspontaneous Processes
An endergonic reaction (also called a nonspontaneous reaction or an unfavourable 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..

Dispersal of Matter and Energy


As we extend our discussion of thermodynamic concepts toward the objective of predicting
spontaneity, consider now an isolated system consisting of two flasks connected with a closed valve.
Initially there is an ideal gas on the left and a vacuum .When the valve is opened, the gas
spontaneously expands to fill both flasks. Recalling the definition of pressure-volume work from the
chapter on thermochemistry, note that no work has been done because the pressure in a vacuum is
zero.

w=−PΔV=0(P=0inavaccum)(19.1.2)(19.1.2)w=−PΔV=0(P=0inavaccum)

Note as well that since the system is isolated, no heat has been exchanged with the
surroundings (q = 0). The first law of thermodynamics confirms that there has been no change in the
system‘s internal energy as a result of this process.

ΔU=q+w=0+0=0(19.1.3)(19.1.3)ΔU=q+w=0+0=0

The spontaneity of this process is therefore not a consequence of any change in energy that
accompanies the process. Instead, the driving force appears to be related to the greater, moreuniform
dispersal of matter that results when the gas is allowed to expand. Initially, the system was comprised
of one flask containing matter and another flask containing nothing. After the spontaneous process took
place, the matter was distributed both more widely (occupying twice its original volume) and more
uniformly (present in equal amounts in each flask).

Now consider two objects at different temperatures: object X at temperature TX and object Y at
temperature TY, with TX > TY (Figure 19.1.419.1.4). When these objects come into contact, heat
spontaneously flows from the hotter object (X) to the colder one (Y). This corresponds to a loss of
thermal energy by X and a gain of thermal energy by Y.

qX<0andqY=−qX>0(19.1.4)(19.1.4)qX<0andqY=−qX>0

From the perspective of this two-object system, there was no net gain or loss of thermal energy,
rather the available thermal energy was redistributed among the two objects. This spontaneous process
resulted in a more uniform dispersal of energy.
The Reverse Process
A process that is spontaneous in one direction is not spontaneous in the reverse direction. Here
are a few examples:

• A sugar cube will spontaneously dissolve in hot coffee, but dissolved sugar will not
spontaneously come out of the coffee and form a sugar cube.

• Popcorn kernels will spontaneously pop when heated, but the popped popcorn will not
spontaneously go back to being a kernel.

• If you add a drop of blue food coloring to a glass of water, the dye will spontaneously mix
with the water to form blue water. In contrast, a cup of blue water will not spontaneously separate into
clear water and a drop of blue food coloring.

It's also important to remember that the direction of the spontaneous process can depend on
temperature: below freezing water spontaneously freezes, whereas above freezing ice spontaneously
melts. In both of these cases, the reverse process is not spontaneous.

Just because a process is spontaneous, doesn't mean it has to happen quickly. Spontaneous
processes sometimes happen very slowly. For instance, have you ever had to polish your
grandmother's tarnished silver? I have, and I am very thankful that even though the tarnishing of silver
is a spontaneous process, it's also a very slow process. I therefore only had to help my grandmother
polish silver about once a year. Some spontaneous processes happen very quickly, like the example of
the glass of water falling to the ground; others, like the rusting of iron, can happen slowly over the
course of years.

Entropy
Entropy and efficiency limits
The concept of entropy was first introduced in 1850
by Clausius as a precise mathematical way of testing
whether the second law of thermodynamics is violated by a
particular process. The test begins with the definition that if
an amount of heat Q flows into a heat reservoir at constant
temperature T, then its entropy S increases by ΔS = Q/T.
(This equation in effect provides a thermodynamic definition
of temperature that can be shown to be identical to the
conventional thermometric one.) Assume now that there are
two heat reservoirs R1 and R2 at temperatures T1 and T2.

Figure 38 Rudolf Clausius


Figure 39 Entropy

The condition ΔS ≥ 0 determines the maximum possible efficiency of heat engines. Suppose
that some system capable of doing work in a cyclic fashion (a heat engine) absorbs heat Q1 from R1
and exhausts heat Q2 to R2 for each complete cycle. Because the system returns to its original state
at the end of a cycle, its energy does not change. Then, by conservation of energy, the work done
per cycle is W = Q1 − Q2, and the net entropy change for the two reservoirs is (4)To make W as
large as possible, Q2 should be kept as small as possible relative to Q1. However, Q2 cannot be
zero, because this would make ΔS negative and so violate the second law of thermodynamics. The
smallest possible value of Q2corresponds to the condition ΔS = 0, yielding (5)This is the fundamental
equation limiting the efficiency of all heat engines whose function is to convert heat into work (such
as electric power generators). The actual efficiency is defined to be the fraction of Q1 that is
converted to work (W/Q1), which is equivalent to equation (2).

The maximum efficiency for a given T1 and T2 is thus (6)A process for which ΔS = 0 is said to
be reversible because an infinitesimal change would be sufficient to make the heat engine run
backward as a refrigerator.

As an example, the properties of materials limit the practical upper temperature for thermal
power plants to T1 ≅ 1,200 K. Taking T2 to be the temperature of the environment (300 K), the
maximum efficiency is 1 − 300/1,200 = 0.75. Thus, at least 25 percent of the heat energy produced
must be exhausted into the environment as waste heat to avoid violating the second law of
thermodynamics. Because of various imperfections, such as friction and imperfect thermal insulation,
the actual efficiency of power plants seldom exceeds about 60 percent. However, because of the
second law of thermodynamics, no amount of ingenuity or improvements in design can increase the
efficiency beyond about 75 percent.
Entropy and heat death
The example of a heat engine illustrates one of the many ways in which the second law of
thermodynamics can be applied. One way to generalize the example is to consider the heat engine
and its heat reservoir as parts of an isolated (or closed) system—i.e., one that does not exchange
heat or work with its surroundings. For example, the heat engine and reservoir could be encased in a
rigid container with insulating walls. In this case the second law of thermodynamics (in the simplified
form presented here) says that no matter what process takes place inside the container, its entropy
must increase or remain the same in the limit of a reversible process. Similarly, if the universe is an
isolated system, then its entropy too must increase with time. Indeed, the implication is that the
universe must ultimately suffer a ―heat death‖ as its entropy progressively increases toward a
maximum value and all parts come into thermal equilibrium at a uniform temperature. After that point,
no further changes involving the conversion of heat into useful work would be possible. In general,
the equilibrium state for an isolated system is precisely that state of maximum entropy. (This is
equivalent to an alternate definition for the term entropy as a measure of the disorder of a system,
such that a completely random dispersion of elements corresponds to maximum entropy, or minimum
information.

Figure 40 Heat death

Entropy and the arrow of time


The inevitable increase of entropy with time for isolated systems plays a fundamental role in
determining the direction of the ―arrow of time.‖ Everyday life presents no difficulty in distinguishing
the forward flow of time from its reverse. For example, if a film showed a glass of warm water
spontaneously changing into hot water with ice floating on top, it would immediately be apparent that
the film was running backward because the process of heat flowing from warm water to hot water
would violate the second law of thermodynamics. However, this obvious asymmetry between the
forward and reverse directions for the flow of time does not persist at the level of fundamental
interactions. An observer watching a film showing two water molecules colliding would not be able to
tell whether the film was running forward or backward.
Figure 41 Relation of entropy with the arrow of time

So what exactly is the connection between entropy and the second law? Recall that heat at
the molecular level is the random kinetic energy of motion of molecules, and collisions between
molecules provide the microscopic mechanism for transporting heat energy from one place to
another. Because individual collisions are unchanged by reversing the direction of time, heat can flow
just as well in one direction as the other. Thus, from the point of view of fundamental interactions,
there is nothing to prevent a chance event in which a number of slow-moving (cold) molecules
happen to collect together in one place and form ice, while the surrounding water becomes hotter.
Such chance events could be expected to occur from time to time in a vessel containing only a few
water molecules. However, the same chance events are never observed in a full glass of water, not
because they are impossible but because they are exceedingly improbable. This is because even a
small glass of water contains an enormous number of interacting molecules (about 1024), making it
highly unlikely that, in the course of their random thermal motion, a significant fraction of cold
molecules will collect together in one place. Although such a spontaneous violation of the second law
of thermodynamics is not impossible, an extremely patient physicist would have to wait many times
the age of the universe to see it happen.

The foregoing demonstrates an important point: the second law of thermodynamics is


statistical in nature. It has no meaning at the level of individual molecules, whereas the law becomes
essentially exact for the description of large numbers of interacting molecules. In contrast, the first
law of thermodynamics, which expresses conservation of energy, remains exactly true even at the
molecular level.
Figure 42 Arrow of time

The example of ice melting in a glass of hot water also demonstrates the other sense of the
term entropy, as an increase in randomness and a parallel loss of information. Initially, the total
thermal energy is partitioned in such a way that all of the slow-moving (cold) molecules are located in
the ice and all of the fast-moving (hot) molecules are located in the water (or water vapour). After the
ice has melted and the system has come to thermal equilibrium, the thermal energy is uniformly
distributed throughout the system. The statistical approach provides a great deal of valuable insight
into the meaning of the second law of thermodynamics, but, from the point of view of applications, the
microscopic structure of matter becomes irrelevant. The great beauty and strength of classical
thermodynamics are that its predictions are completely independent of the microscopic structure of
matter.The foregoing demonstrates an important point: the second law of thermodynamics is
statistical in nature. It has no meaning at the level of individual molecules, whereas the law becomes
essentially exact for the description of large numbers of interacting molecules. In contrast, the first
law of thermodynamics, which expresses conservation of energy, remains exactly true even at the
molecular level.

Entropy as an exact differential


Because the quantity dS = d′Qmax/T is an exact differential, many other important
relationships connecting the thermodynamic properties of substances can be derived. For example,
with the substitutions d′Q = T dS and d′W = P dV, the differential form (dU = d′Q − d′W) of the first law
of thermodynamics becomes (for a single pure substance)dU = T dS − P dV.
The advantage gained by the above formula is that dU is now expressed entirely in terms of
state functions in place of the path-dependent quantities d′Q and d′W. This change has the very
important mathematical implication that the appropriate independent variables are S and V in place
of T and V, respectively, for internal energy.

This replacement of T by S as the most appropriate independent variable for the internal
energy of substances is the single most valuable insight provided by the combined first and second
laws of thermodynamics. With U regarded as a function U(S, V), its differential dU is

A comparison with the preceding equation shows immediately that the

partial derivatives are Furthermore, the cross partial derivatives,

must be equal because the order of differentiation in


calculating the second derivatives of U does not matter. Equating the right-hand sides of the above

pair of equations then yields

This is one of four Maxwell relations (the others will follow shortly). They are all extremely
useful in that the quantity on the right-hand side is virtually impossible to measure directly, while the
quantity on the left-hand side is easily measured in the laboratory. For the present case one simply
measures the adiabatic variation of temperature with volume in an insulated cylinder so that there is
no heat flow (constant S).

The other three Maxwell relations follow by similarly considering the differential expressions for
the thermodynamic potentials F(T, V), H(S, P), and G(T, P), with independent variables as indicated.

The results are

As an example of the use of these equations, equation (35) for CP − CV contains the partial

derivative which vanishes for an ideal gas and is difficult to evaluate directly from experimental
data for real substances. The general properties of partial derivatives can first be used to write it in

the form

Combining this with equation for the partial derivatives together with the first of the Maxwell

equations from equation then yields the desired result

The quantity comes directly from differentiating the equation of state. For an ideal gas

(and so is zero as expected. The departure of from zero reveals directly


the effects of internal forces between the molecules of the substance and the work that must be done
against them as the substance expands at constant temperature.
The Second Law of Thermodynamics
The first law of thermodynamics asserts that energy must be conserved in any process
involving the exchange of heat and work between a system and its surroundings. A machine that
violated the first law would be called a perpetual motion machine of the first kind because it would
manufacture its own energy out of nothing and thereby run forever. Such a machine would be
impossible even in theory. However, this impossibility would not prevent the construction of a
machine that could extract essentially limitless amounts of heat from its surroundings (earth, air, and
sea) and convert it entirely into work. Although such a hypothetical machine would not violate
conservation of energy, the total failure of inventors to build such a machine, known as a perpetual
motion machine of the second kind, led to the discovery of the second law of thermodynamics. The
second law of thermodynamics can be precisely stated in the following two forms, as originally
formulated in the 19th century by the Scottish physicist William Thomson (Lord Kelvin) and the
German physicist Rudolf Clausius, respectively:

A cyclic transformation whose only final result is to transform heat extracted from a source
which is at the same temperature throughout into work is impossible.

A cyclic transformation whose only final result is to transfer heat from a body at a given
temperature to a body at a higher temperature is impossible.

The two statements are in fact equivalent


because, if the first were possible, then the work
obtained could be used, for example, to generate
electricity that could then be discharged through an
electric heater installed in a body at a higher
temperature. The net effect would be a flow of heat from
a lower temperature to a higher temperature, thereby
violating the second (Clausius) form of the second law.
Conversely, if the second form were possible, then the
heat transferred to the higher temperature could be
used to run a heat engine that would convert part of the
heat into work. The final result would be a conversion of
heat into work at constant temperature—a violation of
Figure 43 Perpetual motion machine
the first (Kelvin) form of the second law.
Central to the following discussion of entropy is the
concept of a heat reservoir capable of providing essentially
limitless amounts of heat at a fixed temperature. This is of
course an idealization, but the temperature of a large body of
water such as the Atlantic Ocean does not materially change
if a small amount of heat is withdrawn to run a heat engine.
The essential point is that the heat reservoir is assumed to
have a well-defined temperature that does not change as a
Figure 45 Perpetual motion machine
result of the process being considered.

Gibbs free Energy and Chemical Equilibrium


All batteries depend on some chemical reaction of the form reactants → products for the
generation of electricity or on the reverse reaction as the battery is recharged. The change in free
energy (−ΔG) for a reaction could be determined by measuring directly the amount of electrical
work that the battery could do and then using the equation Wmax = −ΔG. However, the power of
thermodynamics is that −ΔG can be calculated without having to build every possible battery and
measure its performance. If the Gibbs free energies of the individual substances making up a
battery are known, then the total free energies of the reactants can be subtracted from the total
free energies of the products in order to find the change in Gibbs free energy for the reaction,ΔG
= Gproducts − Greactants. (16)Once the free energies are known for a wide variety of
substances, the best candidates for actual batteries can be quickly discerned. In fact, a good part
of the practice of thermodynamics is concerned with determining the free energies and other
thermodynamic properties of individual substances in order that ΔG for reactions can be
calculated under different conditions of temperatureand pressure.

Figure 44 Willard Gibbs


A criterion for equilibrium is that the total free energy (Gibbs free energy, Gr)of the
reaction is at a minimum:

•If we add more reactant or more product, the reaction will proceed spontaneously(without
external help) as long as the value for G decreases

•Thus, a reaction in the direction of decreasing Gris spontaneous. A reaction in the


direction of increasing Gris not spontaneous, and will not occur in a closed system.

1.If ∆Gr< 0, (i.e., ∆Gris negative and thus G decreases as the reaction proceeds), then the
reaction proceeds spontaneously as written

2.If ∆Gr> 0, (i.e., ∆Gris positive and thus G increases as the reaction proceeds), then the
reaction proceeds spontaneously in the opposite direction as written

3.If ∆Gr= 0, (i.e., ∆Gris at a minimum), then the reaction is at equilibrium and will not
proceeds spontaneously in either direction
Bibliography
Hamby, Marcy. "Understanding the language: Problem solving and the first law of
thermodynamics." J. Chem. Educ. 1990: 67, 923.

Chang, Raymond. Chemistry: Ninth Edition

Petrucci, Harwood, Herring, Madura.General Chemistry: Ninth Edition

Lucas, J. (2015, May 19). What Is the First Law of Thermodynamics? Retrieved January 30, 2018, from
https://www.livescience.com/50881-first-law-thermodynamics.html

(n.d.). Retrieved January 30, 2018, from https://www.cliffsnotes.com/study-


guides/chemistry/chemistry/thermodynamics/introduction-to-thermodynamics

L. (2017, August 28). 5.4: Enthalpy of Reaction. Retrieved January 30, 2018, from
https://chem.libretexts.org/Textbook_Maps/General_Chemistry_Textbook_Maps/Map%3A_Chemistry%
3A_The_Central_Science_(Brown_et_al.)/05._Thermochemistry/5.4%3A_Enthalpy_of_Reaction

Drake, G. W. (2017, October 20). Thermodynamics. Retrieved January 30, 2018, from
https://www.britannica.com/science/thermodynamics/Isothermal-and-adiabatic-processes

Thermodynamics:Thermochemical Equations. (n.d.). Retrieved January 30, 2018, from


https://www.chem.wisc.edu/deptfiles/genchem/netorial/modules/thermodynamics/chemical/chemical3.ht
m

Harwood , William S., F. Geoffrey Herring, Ralph H. Petrucci, and Jeffry D. Madura. General
Chemistry: Principles and Modern Applications. 9th ed.. Upper Saddle River, NJ: Pearson
Education, inc., 2007.

Marzzacco, Charles J. "The Enthalpy of Decomposition of Hydrogen Peroxide: A General


Chemistry Calorimetry Experiment." J. Chem. Educ. 1999: 76, 1517.

Skoog, Douglas A. Principles of Instrumental Analysis. Brooks/Cole, 1985. Print.

Chang, Raymond. Physical Chemistry for the Biosciences. New York: University Science, 2005.
Print.
Petrucci, et al. General Chemistry: Principles & Modern Applications: AIE (Hardcover). Upper
Saddle River: Pearson/Prentice Hall, 2007.

Kotz, John C., Treichel, Paul M., and Townsend, John. Chemistry & Chemical Reactivity. 7th
Ed. Belmont: Thomson Higher Education, 2006. Print.
L. (2018, January 23). Calorimetry. Retrieved January 30, 2018, from
https://chem.libretexts.org/Core/Physical_and_Theoretical_Chemistry/Thermodynamics/Calorimetry

Calorimetry. (n.d.). Retrieved January 30, 2018, from


http://www.science.uwaterloo.ca/~cchieh/cact/c120/calorimetry.html

Calorimeters and Calorimetry. (n.d.). Retrieved January 30, 2018, from


http://www.physicsclassroom.com/class/thermalP/Lesson-2/Calorimeters-and-Calorimetry

Helmenstine, P. A. (n.d.). What Is a Spontaneous Process? Retrieved January 30, 2018, from
https://www.thoughtco.com/definition-of-spontaneous-process-604657

Drake, G. W. (2016, November 11). Entropy. Retrieved January 30, 2018, from
https://www.britannica.com/science/entropy-physics

S-ar putea să vă placă și