Energy and the direction of chemical process

 

Energy. Energy can be defined as the capacity to supply heat or do work. One type of work (w) is the process of causing matter to move against an opposing force. For example, we do work when we inflate a bicycle tire—we move matter (the air in the pump) against the opposing force of the air already in the tire

Like matter, energy comes in different types. One scheme classifies energy into two types: potential energy, the energy an object has because of its relative position, composition, or condition, and kinetic energy, the energy that an object possesses because of its motion. Water at the top of a waterfall or dam has potential energy because of its position; when it flows downward through generators, it has kinetic energy that can be used to do work and produce electricity in a hydroelectric plant (figure 6.1). A battery has potential energy because the chemicals within it can produce electricity that can do work.

 

6.1.png 

Figure 6.1 (a) Water that is higher in elevation, for example, at the top of Victoria Falls, has a higher potential energy than water at a lower elevation. As the water falls, some of its potential energy is converted into kinetic energy. (b) If the water flows through generators at the bottom of a dam, such as the Hoover Dam shown here, its kinetic energy is converted into electrical energy.

 

Energy can be converted from one form into another, but all of the energy present before a change occurs always exists in some form after the change is completed. This observation is expressed in the law of conservation of energy: during a chemical or physical change, energy can be neither created nor destroyed, although it can be changed in form. (This is also one version of the first law of thermodynamics, as you will learn later.)

When one substance is converted into another, there is always an associated conversion of one form of energy into another. Heat is usually released or absorbed, but sometimes the conversion involves light, electrical energy, or some other form of energy. For example, chemical energy (a type of potential energy) is stored in the molecules that compose gasoline. When gasoline is combusted within the cylinders of a car’s engine, the rapidly expanding gaseous products of this chemical reaction generate mechanical energy (a type of kinetic energy) when they move the cylinders’ pistons.

According to the law of conservation of matter (seen in an earlier chapter), there is no detectable change in the total amount of matter during a chemical change. When chemical reactions occur, the energy changes are relatively modest and the mass changes are too small to measure, so the laws of conservation of matter and energy hold well. However, in nuclear reactions, the energy changes are much larger (by factors of a million or so), the mass changes are measurable, and matter-energy conversions are significant. This will be examined in more detail in a later chapter on nuclear chemistry. To encompass both chemical and nuclear changes, we combine these laws into one statement: The total quantity of matter and energy in the universe is fixed [2, p. 241].

 

Thermal Energy, Temperature and Heat. Thermal energy is kinetic energy associated with the random motion of atoms and molecules. Temperature is a quantitative measure of “hot” or “cold.” When the atoms and molecules in an object are moving or vibrating quickly, they have a higher average kinetic energy (KE), and we say that the object is “hot.” When the atoms and molecules are moving slowly, they have lower KE, and we say that the object is “cold” (figure 6.2). Assuming that no chemical reaction or phase change (such as melting or vaporizing) occurs, increasing the amount of thermal energy in a sample of matter will cause its temperature to increase. And, assuming that no chemical reaction or phase change (such as condensation or freezing) occurs, decreasing the amount of thermal energy in a sample of matter will cause its temperature to decrease.

 

6.2.png

Figure 6.2 (a) The molecules in a sample of hot water move more rapidly than (b) those in a sample of cold water.

 

Most substances expand as their temperature increases and contract as their temperature decreases. This property can be used to measure temperature changes, as shown in Figure 6.3. The operation of many thermometers depends on the expansion and contraction of substances in response to temperature changes.

 

6.3.png

Figure 6.3 (a) In an alcohol or mercury thermometer, the liquid (dyed red for visibility) expands when heated and contracts when cooled, much more so than the glass tube that contains the liquid. (b) In a bimetallic thermometer, two different metals (such as brass and steel) form a two-layered strip. When heated or cooled, one of the metals (brass) expands or contracts more than the other metal (steel), causing the strip to coil or uncoil. Both types of thermometers have a calibrated scale that indicates the temperature.

 

Heat (q) is the transfer of thermal energy between two bodies at different temperatures. Heat flow (a redundant term, but one commonly used) increases the thermal energy of one body and decreases the thermal energy of the other. Suppose we initially have a high temperature (and high thermal energy) substance (H) and a low temperature (and low thermal energy) substance (L). The atoms and molecules in H have a higher average KE than those in L. If we place substance H in contact with substance L, the thermal energy will flow spontaneously from substance H to substance L. The temperature of substance H will decrease, as will the average KE of its molecules; the temperature of substance L will increase, along with the average KE of its molecules. Heat flow will continue until the two substances are at the same temperature (figure 6.4).

 

6.4.png

Figure 6.4 (a) Substances H and L are initially at different temperatures, and their atoms have different average kinetic energies. (b) When they are put into contact with each other, collisions between the molecules result in the transfer of kinetic (thermal) energy from the hotter to the cooler matter. (c) The two objects reach “thermal equilibrium” when both substances are at the same temperature, and their molecules have the same average kinetic energy.

 

Matter undergoing chemical reactions and physical changes can release or absorb heat. A change that releases heat is called an exothermic process. For example, the combustion reaction that occurs when using an oxyacetylene torch is an exothermic process—this process also releases energy in the form of light as evidenced by the torch’s flame. A reaction or change that absorbs heat is an endothermic process. A cold pack used to treat muscle strains provides an example of an endothermic process. When the substances in the cold pack (water and a salt like ammonium nitrate) are brought together, the resulting process absorbs heat, leading to the sensation of cold.

 

Exothermic reactions:

·         Heat is transferred from the system to the surroundings.

·         Δ𝐻 is negative.

·         The enthalpy of the products is lower than the enthalpy of the reactants.

·         The products are energetically more stable than the reactants.

·         This can be represented in an energy diagram (figure 6.5):

 

 6.5.png

Figure 6.5 Exothermic

 

Endothermic reactions:

·         Heat is absorbed from the surroundings.

·         Δ𝐻 is positive.

·         The enthalpy of the products is greater than the enthalpy of the reactants.

·         The products are energetically less stable that the reactants.

·         This can also be represented in a similar energy diagram (figure 6.6): [21, p. 1]

 

6.6.png 

Figure 6.6 Endothermic

 

Historically, energy was measured in units of calories (cal). A calorie is the amount of energy required to raise one gram of water by 1 degree C (1 kelvin). However, this quantity depends on the atmospheric pressure and the starting temperature of the water. The ease of measurement of energy changes in calories has meant that the calorie is still frequently used. The Calorie (with a capital C), or large calorie, commonly used in quantifying food energy content, is a kilocalorie. The SI unit of heat, work, and energy is the joule. A joule (J) is defined as the amount of energy used when a force of 1 newton moves an object 1 meter. It is named in honor of the English physicist James Prescott Joule. One joule is equivalent to 1 kg m2/s2, which is also called 1 newton–meter. A kilojoule (kJ) is 1000 joules. To standardize its definition, 1 calorie has been set to equal 4.184 joules.

We now introduce two concepts useful in describing heat flow and temperature change. The heat capacity (C) of a body of matter is the quantity of heat (q) it absorbs or releases when it experiences a temperature change (ΔT) of 1 degree Celsius (or equivalently, 1 kelvin) (1):

 

C = q/∆T (1)

 

Heat capacity is determined by both the type and amount of substance that absorbs or releases heat. It is therefore an extensive property—its value is proportional to the amount of the substance.

For example, consider the heat capacities of two cast iron frying pans. The heat capacity of the large pan is five times greater than that of the small pan because, although both are made of the same material, the mass of the large pan is five times greater than the mass of the small pan. More mass means more atoms are present in the larger pan, so it takes more energy to make all of those atoms vibrate faster. The heat capacity of the small cast iron frying pan is found by observing that it takes 18,150 J of energy to raise the temperature of the pan by 50.0 °C:

 

Csmall pan = 18,140 J/ 50,0 oC = 363 J/oC

 

The larger cast iron frying pan, while made of the same substance, requires 90,700 J of energy to raise its temperature by 50.0 °C. The larger pan has a (proportionally) larger heat capacity because the larger amount of material requires a (proportionally) larger amount of energy to yield the same temperature change:

 

Csmall pan = 90,700 J/ 50,0 oC = 1814 J/oC

 

   The specific heat capacity (c) of a substance, commonly called its “specific heat,” is the quantity of heat required to raise the temperature of 1 gram of a substance by 1 degree Celsius (or 1 kelvin) (2):

 

C = q/m∆T (2)

 

Specific heat capacity depends only on the kind of substance absorbing or releasing heat. It is an intensive property—the type, but not the amount, of the substance is all that matters. For example, the small cast iron frying pan has a mass of 808 g. The specific heat of iron (the material used to make the pan) is therefore:

 

Ciron = 18,140 J/ (808g)(50,0 oC) = 0,449 J/oC

 

The large frying pan has a mass of 4040 g. Using the data for this pan, we can also calculate the specific heat of iron:

 

Ciron = 90,700 J/ (4040g)(50,0 oC) = 0,449 J/oC

 

Although the large pan is more massive than the small pan, since both are made of the same material, they both yield the same value for specific heat (for the material of construction, iron). Note that specific heat is measured in units of energy per temperature per mass and is an intensive property, being derived from a ratio of two extensive properties (heat and mass). The molar heat capacity, also an intensive property, is the heat capacity per mole of a particular substance and has units of J/mol °C.

 

Liquid water has a relatively high specific heat (about 4.2 J/g °C); most metals have much lower specific heats (usually less than 1 J/g °C). The specific heat of a substance varies somewhat with temperature. However, this variation is usually small enough that we will treat specific heat as constant over the range of temperatures that will be considered in this chapter. Specific heats of some common substances are listed in Table 6.1.

 

Table 6.1

Specific Heats of Common Substances at 25 °C and 1 bar

 

Substance

 

Symbol (state)

Specific Heat (J/g °C)

helium

He(g)

5.193

 

water

H2O(l)

4.184

 

ethanol

C2H5OH(l)

2.376

 

ice

H2O(s)

2.093 (at −10 °C)

 

water vapor

H2O(g)

 

1.864

nitrogen

N2(g)

1.040

 

air

 

1.007

 

oxygen

O2(g)

 

0.918

aluminum

Al(s)

 

0.897

carbon dioxide

CO2(g)

 

0.853

Argon

Ar(g)

 

0.522

iron

Fe(s)

 

0.449

copper

Cu(s)

0.385

 

lead

Pb(s)

 

0.130

Gold

Au(s)

0.129

 

silicon

Si(s)

0.712

 

 

If we know the mass of a substance and its specific heat, we can determine the amount of heat, q, entering or leaving the substance by measuring the temperature change before and after the heat is gained or lost (3):

 

q = (specific hea ) × (mass of substance) × (temperature change)

q = c × m × ΔT = c × m × (TfinaTinitial) (3)

 

In this equation, c is the specific heat of the substance, m is its mass, and ΔT (which is read “delta T”) is the temperature change, TfinalTinitial. If a substance gains thermal energy, its temperature increases, its final temperature is higher than its initial temperature, TfinalTinitial has a positive value, and the value of q is positive. If a substance loses thermal energy, its temperature decreases, the final temperature is lower than the initial temperature, TfinalTinitial has a negative value, and the value of q is negative [2, p. 242].

 

Enthalpy. Thermochemistry is a branch of chemical thermodynamics, the science that deals with the relationships between heat, work, and other forms of energy in the context of chemical and physical processes. As we concentrate on thermochemistry in this chapter, we need to consider some widely used concepts of thermodynamics.

Substances act as reservoirs of energy, meaning that energy can be added to them or removed from them. Energy is stored in a substance when the kinetic energy of its atoms or molecules is raised. The greater kinetic energy may be in the form of increased translations (travel or straight-line motions), vibrations, or rotations of the atoms or molecules. When thermal energy is lost, the intensities of these motions decrease and the kinetic energy falls. The total of all possible kinds of energy present in a substance is called the internal energy (U), sometimes symbolized as E.

As a system undergoes a change, its internal energy can change, and energy can be transferred from the system to the surroundings, or from the surroundings to the system. Energy is transferred into a system when it absorbs heat (q) from the surroundings or when the surroundings do work (w) on the system. For example, energy is transferred into room-temperature metal wire if it is immersed in hot water (the wire absorbs heat from the water), or if you rapidly bend the wire back and forth (the wire becomes warmer because of the work done on it). Both processes increase the internal energy of the wire, which is reflected in an increase in the wire’s temperature. Conversely, energy is transferred out of a system when heat is lost from the system, or when the system does work on the surroundings.

The relationship between internal energy, heat, and work can be represented by the equation (4):

 

ΔU = q + w (4)

 

as shown in figure 6.7. This is one version of the first law of thermodynamics, and it shows that the internal energy of a system changes through heat flow into or out of the system (positive q is heat flow in; negative q is heat flow out) or work done on or by the system. The work, w, is positive if it is done on the system and negative if it is done by the system.

 

6.7.png 

Figure 6.7 The internal energy, U, of a system can be changed by heat flow and work. If heat flows into the system, qin, or work is done on the system, won, its internal energy increases, ΔU < 0. If heat flows out of the system, qout, or work is done by the system, wby, its internal energy decreases, ΔU > 0.

 

A type of work called expansion work (or pressure-volume work) occurs when a system pushes back the surroundings against a restraining pressure, or when the surroundings compress the system. An example of this occurs during the operation of an internal combustion engine. The reaction of gasoline and oxygen is exothermic. Some of this energy is given off as heat, and some does work pushing the piston in the cylinder. The substances involved in the reaction are the system, and the engine and the rest of the universe are the surroundings. The system loses energy by both heating and doing work on the surroundings, and its internal energy decreases. (The engine is able to keep the car moving because this process is repeated many times per second while the engine is running.) We will consider how to determine the amount of work involved in a chemical or physical change in the chapter on thermodynamics.

Chemists ordinarily use a property known as enthalpy (H) to describe the thermodynamics of chemical and physical processes. Enthalpy is defined as the sum of a system’s internal energy (U) and the mathematical product of its pressure (P) and volume (V):

 

H = U + PV

 

Since it is derived from three state functions (U, P, and V), enthalpy is also a state function. Enthalpy values for specific substances cannot be measured directly; only enthalpy changes for chemical or physical processes can be determined. For processes that take place at constant pressure (a common condition for many chemical and physical changes), the enthalpy change (ΔH) is (5):

 

ΔH = ΔU + PΔV (5)

 

The mathematical product PΔV represents work (w), namely, expansion or pressure-volume work as noted. By their definitions, the arithmetic signs of ΔV and w will always be opposite:

 

PΔV = −w

 

Substituting this equation and the definition of internal energy into the enthalpy-change equation yields:

 

ΔH = ΔU + PΔV = qp + w w = qp

 

where qp is the heat of reaction under conditions of constant pressure. And so, if a chemical or physical process is carried out at constant pressure with the only work done caused by expansion or contraction, then the heat flow (qp) and enthalpy change (ΔH) for the process are equal. The heat given off when you operate a Bunsen burner is equal to the enthalpy change of the methane combustion reaction that takes place, since it occurs at the essentially constant pressure of the atmosphere. On the other hand, the heat produced by a reaction measured in a bomb calorimeter is not equal to ΔH because the closed, constant-volume metal container prevents expansion work from occurring. Chemists usually perform experiments under normal atmospheric conditions, at constant external pressure with q = ΔH, which makes enthalpy the most convenient choice for determining heat.

The following conventions apply when we use ΔH:

1.      Chemists use a thermochemical equation to represent the changes in both matter and energy. In a thermochemical equation, the enthalpy change of a reaction is shown as a ΔH value following the equation for the reaction. This ΔH value indicates the amount of heat associated with the reaction involving the number of moles of reactants and products as shown in the chemical equation. For example, consider this equation:

 

H2(g) + 1/2O2(g) H2O(l) ΔH = −286 kJ

 

This equation indicates that when 1 mole of hydrogen gas and 12 mole of oxygen gas at some temperature and pressure change to 1 mole of liquid water at the same temperature and pressure, 286 kJ of heat are released to the surroundings. If the coefficients of the chemical equation are multiplied by some factor, the enthalpy change must be multiplied by that same factor (ΔH is an extensive property):

 

(two-fold increase in amounts)

2H2(g) + O2(g) 2H2O(l) ΔH = 2 × (−286 kJ) = −572 kJ

(two-fold decrease in amounts)

1/2H2(g) + 1/4O2(g) 12H2O(l) ΔH = 12; × (−286 kJ) = −143 kJ

 

2.      The enthalpy change of a reaction depends on the physical state of the reactants and products of the reaction (whether we have gases, liquids, solids, or aqueous solutions), so these must be shown. For example, when 1mole of hydrogen gas and 12 mole of oxygen gas change to 1 mole of liquid water at the same temperature and pressure, 286 kJ of heat are released. If gaseous water forms, only 242 kJ of heat are released.

 

H2(g) + 1/2O2(g) H2O(g) ΔH = −242 kJ

 

3.      A negative value of an enthalpy change, ΔH, indicates an exothermic reaction; a positive value of ΔH indicates an endothermic reaction. If the direction of a chemical equation is reversed, the arithmetic sign of its ΔH is changed (a process that is endothermic in one direction is exothermic in the opposite direction).

 

Enthalpy changes are typically tabulated for reactions in which both the reactants and products are at the same conditions. A standard state is a commonly accepted set of conditions used as a reference point for the determination of properties under other different conditions. For chemists, the IUPAC standard state refers to materials under a pressure of 1 bar and solutions at 1 M, and does not specify a temperature. Many thermochemical tables list values with a standard state of 1 atm. Because the ΔH of a reaction changes very little with such small changes in pressure (1 bar = 0.987 atm), ΔH values (except for the most precisely measured values) are essentially the same under both sets of standard conditions. We will include a superscripted “o” in the enthalpy change symbol to designate standard state. Since the usual (but not technically standard) temperature is 298.15 K, we will use a subscripted “298” to designate this temperature. Thus, the symbol (ΔH°298) is used to indicate an enthalpy change for a process occurring under these conditions. (The symbol ΔH is used to indicate an enthalpy change for a reaction occurring under nonstandard conditions) [21, p. 263].

 

Enthalpy of Combustion. Standard enthalpy of combustion (ΔHC°) is the enthalpy change when 1 mole of a substance burns (combines vigorously with oxygen) under standard state conditions; it is sometimes called “heat of combustion.” For example, the enthalpy of combustion of ethanol, −1366.8 kJ/mol, is the amount of heat produced when one mole of ethanol undergoes complete combustion at 25°C and 1 atmosphere pressure, yielding products also at 25°C and 1 atm.

 

C2H5OH(l) + 3O2(g)2CO2 + 3H2O(l) ΔH°298 = −1366.8 kJ

 

Enthalpies of combustion for many substances have been measured; a few of these are listed in table 6.2. Many readily available substances with large enthalpies of combustion are used as fuels, including hydrogen, carbon (as coal or charcoal), and hydrocarbons (compounds containing only hydrogen and carbon), such as methane, propane, and the major components of gasoline [2, p. 267].

 

Table 6.2

Standard Molar Enthalpies of Combustion

Substance

Combustion Reaction

Enthalpy of Combustion, ΔHc°

( kJ/mol at 25 °C)

Carbon

 

C(s) + O2(g) CO2(g)

−393.5

Hydrogen

 

H2(g) + 1/2O2(g) H2O(l)

−285.8

Magnesium

 

Mg(s) + 1/2O2(g) MgO(s)

−601.6

Sulfur

 

S(s) + O2(g) SO2(g)

−296.8

Carbon monoxide

 

CO(g) + 1/2O2(g) CO2(g)

−283.0

Methane

 

CH4(g) + 2O2(g) CO2(g) + 2H2O(l)

−890.8

Acetylene

 

C2H2(g) + 5/2O2(g) 2CO2(g) + H2O(l)

−1301.1

Ethanol

 

C2H5OH(l) + 2O2(g) CO2(g) + 3H2O(l)

−1366.8

Methanol

 

CH3OH(l) + 3/2O2(g) CO2(g) + 2H2O(l)

−726.1

Isooctane

 

C8H18(l) + 25/2O2(g) 8CO2(g) + 9H2O(l)

−5461

 

Standard Enthalpy of Formation. A standard enthalpy of formation is an enthalpy change for a reaction in which exactly 1 mole of a pure substance is formed from free elements in their most stable states under standard state conditions. These values are especially useful for computing or predicting enthalpy changes for chemical reactions that are impractical or dangerous to carry out, or for processes for which it is difficult to make measurements. If we have values for the appropriate standard enthalpies of formation, we can determine the enthalpy change for any reaction, which we will practice in the next section on Hess’s law.

The standard enthalpy of formation of CO2 (g) is −393.5 kJ/mol. This is the enthalpy change for the exothermic reaction:

 

C(s) + O2 (g) CO2 (g) ΔHf ° = ΔH°298 = −393.5 kJ

 

starting with the reactants at a pressure of 1 atm and 25 °C (with the carbon present as graphite, the most stable form of carbon under these conditions) and ending with one mole of CO2, also at 1 atm and 25 °C. For nitrogen dioxide, NO2 (g), ΔHf ° is 33.2 kJ/mol. This is the enthalpy change for the reaction:

 

1/2N2 (g) + O2 (g) NO2 (g) ΔHf ° = ΔH°298 = +33.2 kJ

 

A reaction equation with ½ mole of N2 and 1 mole of O2 is correct in this case because the standard enthalpy of formation always refers to 1 mole of product, NO2(g).

   You will find a table of standard enthalpies of formation of many common substances in Appendix G. These values indicate that formation reactions range from highly exothermic (such as −2984 kJ/mol for the formation of P4O10) to strongly endothermic (such as +226.7 kJ/mol for the formation of acetylene, C2H2). By definition, the standard enthalpy of formation of an element in its most stable form is equal to zero under standard conditions, which is 1 atm for gases and 1 M for solutions [2, p. 271].

 

Hess’s Law. There are two ways to determine the amount of heat involved in a chemical change: measure it experimentally, or calculate it from other experimentally determined enthalpy changes. Some reactions are difficult, if not impossible, to investigate and make accurate measurements for experimentally. And even when a reaction is not hard to perform or measure, it is convenient to be able to determine the heat involved in a reaction without having to perform an experiment.

 

This type of calculation usually involves the use of Hess’s law, which states: If a process can be written as the sum of several stepwise processes, the enthalpy change of the total process equals the sum of the enthalpy changes of the various steps.

 

Hess’s law is valid because enthalpy is a state function: Enthalpy changes depend only on where a chemical process starts and ends, but not on the path it takes from start to finish. For example, we can think of the reaction of carbon with oxygen to form carbon dioxide as occurring either directly or by a two-step process. The direct process is written:

 

C(s) + O2(g)CO2(g) ΔH°298 = −394 kJ

 

In the two-step process, first carbon monoxide is formed:

 

C(s) + 1/2O2(g)CO(g) ΔH°298 = −111 kJ

 

Then, carbon monoxide reacts further to form carbon dioxide:

 

CO(g) + 1/2O2(g)CO(g) ΔH°298 = −283 kJ

 

The equation describing the overall reaction is the sum of these two chemical changes:

 

Step 1: C(s) + 1/2O2(g) CO(g)

Step 2: CO(g) + 1/2O2(g) CO2(g)

Sum: C(s) + 1/2O2(g) + CO(g) + 1/2O2(g) CO(g) + CO2(g)

 

Because the CO produced in Step 1 is consumed in Step 2, the net change is:

 

C(s) + O2(g) CO2(g)

 

According to Hess’s law, the enthalpy change of the reaction will equal the sum of the enthalpy changes of the steps. We can apply the data from the experimental enthalpies of combustion in table 6.2 to find the enthalpy change of the entire reaction from its two steps:

 

C(s) + 1/2O2(g)CO(g) ΔH°298 = −111 kJ

CO(g) + 1/2O2(g) CO2(g) ΔH°298 = −283 kJ

 

C(s) + O2(g) CO2(g) ΔH°298 = −394 kJ

 

   The result is shown in figure 6.8. We see that ΔH of the overall reaction is the same whether it occurs in one step or two. This finding (overall ΔH for the reaction = sum of ΔH values for reaction “steps” in the overall reaction) is true in general for chemical and physical processes.

 

6.8.png

Figure 6.8 The formation of CO2 (g) from its elements can be thought of as occurring in two steps, which sum to the overall reaction, as described by Hess’s law. The horizontal blue lines represent enthalpies. For an exothermic process, the products are at lower enthalpy than are the reactants.

 

Before we further practice using Hess’s law, let us recall two important features of ΔH.

1.      ΔH is directly proportional to the quantities of reactants or products. For example, the enthalpy change for the reaction forming 1 mole of NO2(g) is +33.2 kJ:

 

1/2N2 (g) + O2 (g) NO2 (g) ΔH = +33.2 kJ

 

When 2 moles of NO2 (twice as much) are formed, the ΔH will be twice as large:

 

N2(g) + 2O2(g) 2NO2(g) ΔH = +66.4 kJ

 

In general, if we multiply or divide an equation by a number, then the enthalpy change should also be multiplied or divided by the same number.

 

2.      ΔH for a reaction in one direction is equal in magnitude and opposite in sign to ΔH for the reaction in the reverse direction. For example, given that:

 

H2(g) + Cl2(g) 2HCl(g) ΔH = −184.6 kJ

 

Then, for the “reverse” reaction, the enthalpy change is also “reversed”:

 

2HCl(g) H2(g) + Cl2(g) ΔH = +184.6 kJ

 

We also can use Hess’s law to determine the enthalpy change of any reaction if the corresponding enthalpies of formation of the reactants and products are available. The stepwise reactions we consider are: (i) decompositions of the reactants into their component elements (for which the enthalpy changes are proportional to the negative of the enthalpies of formation of the reactants), followed by (ii) re-combinations of the elements to give the products (with the enthalpy changes proportional to the enthalpies of formation of the products). The standard enthalpy change of the overall reaction is therefore equal to: (ii) the sum of the standard enthalpies of formation of all the products plus (i) the sum of the negatives of the standard enthalpies of formation of the reactants. This is usually rearranged slightly to be written as follows, with Σ representing “the sum of” and n standing for the stoichiometric coefficients:

 

ΔHreaction ° =Σn × ΔHf ° (products) − Σ n × ΔHf ° (reactants) [2, p. 273]

 

Gibbs Energy. This function has been introduced by Gibbs. The symbol of Gibbs energy is G. Its significance and use are related to an isothermal and isobaric process. More precisely, the Gibbs energy is related to a process in which the temperature and the pressure of the studied system remain equal to the temperature and pressure of its surroundings (at least at the beginning and at the end of the process provided, in this case, that during it, the surroundings remain at the constant temperature Text and pressure pext):

 

p = pext and T = Text

 

The Gibbs energy is defined by the expression (6)

 

G = U + pV - TS (6)

 

where U, p, V, T, and S are, respectively, the internal energy, pressure, volume, temperature, and entropy of the system. Its unity is the Joule J. Owing to its definition, the Gibbs energy is a state function.

The interest of the introduction of this function is the following one: it turns out that the Gibbs function may constitute a criterion of equilibrium and also of evolution specially convenient for any process at constant pressure and temperature.

 

Let us, for example, study the process with the aid of which we want to recover useful work (every work other than that stemming from the change in the volume of the system), starting from the system. (A good example is that of an electrochemical cell producing electrical energy which is connected to an electrical motor. The cell has the property to transform the chemical energy—coming from the two electrochemical reactions which simultaneously take place at each of both electrodes—to electrical work.) It is demonstrated in an absolute general manner that the work given to the surroundings is always weaker than the change in internal energy of the system. In other words, the Gibbs energy of the system cannot do anything else than to decrease when it supplies work to the surroundings, in any case when the process is spontaneous. Hence, we can deduce that

 

ΔG ≤ 0

 

or in differential writing

 

dG ≤ 0

 

At equilibrium

 

dG = 0

 

For a system at equilibrium at given pressure and temperature, the Gibbs energy is at its minimum value.

Hence, with the introduction of the function of Gibbs energy, the criterion of spontaneous evolution of a system, that is to say that of the change in the total entropy (that of the system plus that of its surroundings—both forming an isolated system), is transformed into another one which is the criterion of the decrease of the Gibbs energy of the studied system alone. The latter criterion is evidently less heavy than the former and is easier to handle because it does not require the knowledge of the thermodynamic parameters defining the state of the surroundings. However, the criterion of the Gibbs energy is by far less general than that of entropy because, for its handling, it implies that the process evolves at constant temperature and pressure.

The Gibbs energy function can be defined in another way. Of course, all its definitions are equivalent. For example, taking into account the fact that the enthalpy of a system is defined by the expression

 

H = U + pV

 

the Gibbs energy can also be written (7)

 

G = H – TS (7)

 

or

dG = dH - TdS – SdT [22, p.3]