Download as pdf or txt
Download as pdf or txt
You are on page 1of 45

10 Thermodynamic

Chapter Ten:
Thermodynamics is a funny subject.
The first time you go through it, you
don’t understand it at all. The second
time you go through it, you think you
understand it, except for one or two
small points. The third time you go
through it, you know you don’t
10.1 Chemical Energy and Heat understand it, but by that time you
are so used to it, it doesn’t bother
10.1.1 Everything has internal energy you anymore.
Arnold Sommerfeld
You might not know it, but a glass of water has internal energy. How so? It isn’t moving. Or is it? (Renown quantum physicist and
Think about the water molecules themselves: they are owing around in the liquid (“translational mentor to 7 Nobel Prize winners)
kinetic energy”), they are vibrating and rotating (“vibrational and rotational kinetic energies”), the
molecules attract each other through electrostatic forces (“electrical potential energy”), they have
energy due to the hydrogen and oxygen atoms being bonded together (“bond energy” or “chemical
potential energy”), and so on. There is no way we could experimentally measure all these
contributions to the internal energy. That’s okay. We won’t need to. Just keep in mind that
everything – whether unstable like nitroglycerin or stable like a grain of sand – has internal energy.

10.1.2 Energetics of chemical reactions are due to changes in internal energy


You have noticed that chemical reactions are usually accompanied by changes in temperature. You
might also notice that some reactions are accompanied by changes in volume (for example, the
rapid expansion of gases produced by chemical explosives). Where does the heat come from to
raise the temperature? Where does the “power” come from to cause the gas to expand? It was
there all along as the internal energy in the reactants. However, this internal energy might not be Sure, water molecules are moving
readily useable in its current form. about, but we can’t really tap into this
random KE and use it. How would
We can tap into this energy via chemical reactions. A chemical reaction involves a rearrangement we? Tie a rope around a molecule so
it can pull a cart?
of atoms – some old bonds (connections between atoms) are broken and new bonds are formed.
Different bonds have different strengths. Some bonds take more energy to break and release more
energy when formed compared to other bonds. Because there are different bonds in the products
and reactants, the products will have a different amount of internal energy than did the reactants.
Thus, when a chemical reaction happens, the chemicals will either release or absorb an amount of
energy equal to the difference in their internal energies, which mostly comes from the difference in
bond energies. The energy released or absorbed by chemical reactions is usually mostly in the
form of heat (thermal energy), which we’ll de ne more exactly a little later

In chemistry, we are less interested in the actual internal energies a substance has itself and more
interested in the changes of energy that occur during various processes – whether the process is a
chemical reaction such as an explosion or a physical change such as an ice cube melting. The
ability to calculate the changes of energy that will occur can allow us to make predictions about the
processes, including whether or not they will even occur. This study of energy changes that
accompany chemical or physical processes is called thermodynamics.

“Hi-diddily-ho, neighborinos! Why are only changes important? Remember gravitational potential
energy from physics? The actual GPE an object has isn’t as important as the change in GPE it
experiences as it falls. A kid at the top of a swing has 600J of GPE. So what? Maybe at the bottom
of the swing, she has 150J of GPE. Again, so what? But the change in GPE, 450J, tells us how
much kinetic energy she can gain, and therefore her velocity. Therefore, change in GPE is a more
useful quantity for us to know. For the same reasons, we care only about the changes in internal
energy. For example, the difference in bond energies of the reactants and products tells us how
much energy would be released or absorbed in a chemical reaction. That is useful. Okily dokily?

10.1.3 Detecting energy changes gives us information about the chemical reaction
As just discussed, reactants turning into products involves a change of internal energy. Let’s say
the products have less internal energy than the reactants. Thus, the chemicals have lost energy.
This energy is released by the chemicals, usually mostly in the form of heat. As this heat is
transferred to the environment surrounding the chemicals, it can be felt (“ah...warm”). By measuring
this heat given off, we can calculate how much energy was released by the reaction. In doing so,
this informs us about the reaction. Knowing energy changes of chemical reactions is a goal in
thermodynamics, so this is something we’ll want to learn how to do

Chapter 10 Thermodynamics 357

fi
s

fl
.

10.1.4 Systems, surroundings, and conservation of energy


One goal of this chapter is to come up with methods to determine the changes in energy that occur
in a reaction. But before going into that, we need to introduce some terminology. We call the piece
of the universe in which we are interested – the chemicals reacting for example – the system, and
we call everything else – stuff on the “outside” – the surroundings. If aqueous solutions of NaCl
and AgNO3 react in a beaker, the system would be the chemicals in solution (Na+, Cl¯, Ag+, NO3¯,
product AgCl). The surroundings would be the encompassing water, the beaker, your hand holding
the beaker, the air, etc. For chemical reactions, we always consider the reactants and products the
system: because they rearrange and break/form bonds, they are directly and actively responsible
for the absorption or release of energy. The surroundings are the passive part: they simply gain the
energy lost by the system or have energy taken away from them by the system

An important assumption we make is that if the system loses any energy, the surroundings must
gain that same amount of energy (where else would it go?). For example, the beaker and your
hand may be getting warmer as the reaction progresses. That must mean the system is losing
heat, and the beaker and your hand are gaining that heat. Let’s say that the system (reacting
chemicals) lose 100J of energy in forming the product. Well, how much energy must be gained by
the surrounding water, beaker, your hand, the air, etc.? 100J!

Diagrammatically
 
System Energy Surroundings
Transferred
Energy Energy
Decreases Increases

Mathematically
The negative sign means that if there ∆Esys = – ∆Esur
is a decrease in the energy of the
system, there must be an increase in (The Greek uppercase letter delta, ∆, means “change in.”) This is the familiar idea of conservation
the energy of the surroundings. of energy: energy can neither be created nor destroyed. In fact, this is more than an assumption:
it’s the law. We call the idea of conservation of energy the rst law of thermodynamics. For a
long time, conservation of energy was anything but obvious. For example, a bouncing ball comes to
a stop; a re dies out; energy is used to melt an ice cube. What happened to all that energy? It took
new ideas and careful measurements to realize energy is just changing from one form to another,
more hidden form. But taking into account all forms of energy, the total energy remains constant

10.1.5 Temperature, thermal energy, heat, and speci c heat


Recall that every particle (atom, ion, molecule) has some kind of motion, whether it is moving
across space like a gas particle or vibrating in place like a solid particle. Thus, each particle has a
kinetic energy. For a substance made up of gajillions of particles, its particles will have different
KE’s: some will be moving slowly and have little KE, some will be moving quickly and have a lot of
Temperature itself is not an energy. It KE. We de ne temperature as a measure of the average kinetic energy of its particles: in other
is just a number that is related to the words, the typical KE of one of its particles. For example, a substance at 20ºC has particles that
typical KE of a particle.
have an average KE of about 6 x 10–21 J. That’s a teeny bit of KE, but that makes sense because
each particle has such a teeny mass.

Thermal energy is the total combined kinetic energies of all the particles that make up a
An analogy is like a classroom of ten substance. For example, I just said that at 20ºC, the typical KE of a single particle is 6x10–21J. So if
16-year-olds. The average age is 16, you had a mole (6.02x1023) of that substance, the thermal energy would be (6x10–21J)(6.02x1023) =
and the combined age is 160. The
former is like temperature, the latter
3612 J. Thus, thermal energy depends on the temperature and how many particles you have.
is like thermal energy. Otherwise, temperature and thermal energy are closely related. If you increase the temperature of
a given substance, you increase its thermal energy. If you remove thermal energy from that
substance, you decrease its temperature.

During chemical reactions, the amount of thermal energy that a substance possesses can change:
it can lose or gain thermal energy. The loss or gain of thermal energy is known as heat. Really,
there’s not much difference between the two. “Heat” is just the name we give to the thermal energy
transferred from one place to another. If a system with 4000 J of thermal energy suddenly ends up
with 3000 J of thermal energy, it must have lost 1000 J of thermal energy to the surroundings. We
say that “1000 J of heat were transferred from the system to surroundings.”

358 Chapter 10 Thermodynamics


fi
fi
,

fi

fi

We symbolize the gain or loss of heat by q. When a substance gains heat (thermal energy
increases), its temperature increases. Similarly, a loss of heat/thermal energy would lead to a
decrease in temperature. The temperature change will also depend on how much stuff you have
(its mass) and the particular nature of the substance. We can relate the heat transfer to the
resultant temperature change with this equation
q = sm∆
where ∆T is the change in temperature of a substance of mass m, and s is the speci c heat
capacity of the substance changing temperature. The speci c heat capacity is the amount of
energy required to increase the temperature of one gram of a substance by one kelvin. For
example, liquid water has a speci c heat capacity of 4.184 J/gK: it takes 4.184 joules to increase You can also write 4.184 J/gºC since
one gram of water by 1K. The higher the speci c heat of a material, the more resistant it is to a change of 1ºC is the same as a
changes in temperature. A low speci c heat means it is easy to change its temperature. If we can change of 1K.
measure the mass and change in temperature of a substance, we can use them and the speci c
heat to calculate the heat lost or gained.

Because we use the speci c heat of water so often in chemistry, it would be nice to have a more
convenient number than 4.184. Thus, scientists made up a new energy unit called the calorie (cal),
where 1 cal = 4.184 J. This means that the speci c heat of liquid water can be expressed more
conveniently as 1 cal/gK (or 1 cal/gºC)

“Disco Stu says: don’t get confused by temperature, thermal energy, heat, and speci c heat.
Temperature is a measure of the average KE of the particles of a substance. That is, how much KE
a typical little atom or molecule has. Thermal energy is the total KE of all the particles that make up
the substance. This means that if you have two objects at the same temperature, but Object A has
twice the number of particles as Object B, Object A will have twice as much thermal energy. Heat is
used almost interchangeably with thermal energy. But technically, heat is a measure of the ow of
thermal energy. In other words, heat quanti es the gain or loss of thermal energy. We’d say that an
object possesses thermal energy, but it loses or gains heat. A subtle difference, but a difference
nonetheless. Lastly, speci c heat is neither heat nor thermal energy. It is the amount of heat/energy
needed to raise 1g of a substance by 1ºC. Boogie on!

10.1.6 Calculations involving q=sm∆T


In §10.3, we’ll use this equation to calculate the energy changes of reactions. For now, let’s just get
some practice using it. Let’s try a couple of questions, one solving for q and another solving for ∆T

1. You have 50 g of water initially at 20ºC. How much heat must it absorb in order to reach 30ºC?
Since we are talking about water, s = 4.184 J/gK. The mass is 50 g, and the change in temperature,
∆T, is 10ºC. But note that a change of 10ºC is the same thing as a change of 10K. Now plug in
q = sm∆T = (4.184J/gK)(50g)(10K) = 2092
Thus, 50g of water must absorb 2092J of heat to change by 10ºC.

In this next problem, I’ll be using calories. Also pay attention to how I deal with volumes

2. If you remove 600 cal of heat from 20mL of water at 40ºC, what will be the new temperature
Notice how we are given q in calories, not in joules. We could convert the 600 cal to joules using
the conversion factor 1cal = 4.184J, and then use 4.184J/gK for the speci c heat of water. But why
waste our time? Let’s use 1cal/gK for the speci c heat of water. Also notice that we are given a
volume of water, but the equation requires a mass. Liquid water has a density of 1g/mL (1g/cm3).
That is, each mL of liquid water has a mass of 1g. Thus, if we have 20mL of liquid water, we have Don’t think that everything has a
density of 1 g/mL. Liquid water does,
20g of water. Let’s solve for ∆T but most everything else doesn’t.
∆T = q/sm = (600cal) / [(1cal/gºC)(20g)] = 30º
This isn’t the answer! ∆T is the change in temperature. We want the new temperature. The water
started at 40ºC and it changed by 30ºC. So is the nal temperature 10ºC or is it 70ºC? Notice the
problem said we removed 600 cal from the water. Removing heat will lower the temperature. Thus,
the nal temperature will be 10ºC. Removing heat cools things down, adding heat warms things up

Different substances have different speci c heat values. You can experimentally determine the
speci c heat of a substance by using a calorimeter. A simple calorimeter used in science classes
is made up of two nested styrofoam cups, a lid, and a thermometer that pokes into the cups
through a hole in the lid. The following describes how you can nd the speci c heat of a metal

Chapter 10 Thermodynamics 359


fi
fi
T

fi
fi
:

fi
fi
.

fi
J

fi
C

fi
fi
fi
fi

fi
fi

fi
fi
.

fi
fi
fl
.

fi
.

What we do is place a measured mass of water in the calorimeter and record its initial temperature.
Then we heat up a known mass of the metal to a known starting temperature. Next, we place the
hot piece of metal into the cooler water in the calorimeter. Because the metal is hotter, heat will be
Heat always moves from a hotter transferred from the metal to the water. This happens until the metal and water arrive at a common
object to a cooler object. nal temperature, which we measure with our thermometer. At this point, we know the mass of the
water, the speci c heat of the water (1 cal/gK), and the change in temperature of the water. Thus,
we can calculate how much heat the water gained using q = sm∆T. How does this q gained by the
water compare to the q lost by the hot metal? We assume they are the same: if the water gained a
certain amount of heat, we assume the metal lost that amount of heat. Therefore, we now know the
q of the metal. We also know its mass and its change in temperature. Thus, we can use q = sm∆T
to solve for the one missing variable, s. We’ll do this in the following example.

Calorimetry is the name we give to We need to make a couple of assumptions in calorimetry. First, we will assume that the calorimeter
the experimental process of using a is a perfect insulator. That is, no heat will get into or out of the cup. The accuracy of this assumption
calorimeter. depends on the quality of the calorimeter. Needless to say, the styrofoam coffee cups we use in
class introduce a somewhat signi cant systematic error. Additionally, we assume that all the heat
lost by the metal will be gained by the water only. This is not really true, of course. For example,
some heat will be absorbed by the styrofoam cup, the thermometer immersed in the water, and the
air above the water. However, those things are dif cult to measure (and are hopefully minor). The
only thing that is easy to measure is the heat gained by the water. We assume that this represents
all the heat lost by the metal. We make the assumption for simplicity’s sake, understanding that our
results will be inexact. Nevertheless, we’ll still get a good ballpark gure for s

Example 1:
You heat up a 10-gram piece of silver to 120ºC, then place it in a calorimeter that contains 15mL of
water initially at 16ºC. After a few moments, the silver and water arrive at a temperature of 20ºC.
Calculate the speci c heat of the silver, assuming a perfectly insulted calorimeter
Solution:
For the water, we have its mass (15g), ∆T (4ºC), and we always know its s (1cal/gºC). Calculate its q
qwater = swatermwater∆Twater = (1cal/gºC)(15g)(4ºC) = 60 ca
We now assume that if the water gained 60 cal, the silver must have lost 60 cal. Thus, qAg = 60 cal. We can
use this, its mass (10g), and its ∆T (100ºC) to solve for its speci c heat, s
sAg = qAg / mAg∆TAg = (60cal) / (10g)(100ºC) = 0.06 cal/gºC

Section Overview

The Forest:
Thermodynamics is the study of energy changes in chemical processes. Most of the energy lost or
gained is in the form of heat. This can raise or lower the temperature of the surroundings. By
measuring such changes in temperature, we can determine the energy changes of reactions

The Trees:
All substances possess internal energy. During a chemical process, the amount of internal energy
changes. Energy is exchanged between the system (stuff we’re interested in, like the chemicals)
and the surroundings; but the total energy remains constant (the 1st law of thermodynamics). Most
of this energy is gained or lost in the form of heat, which is the thermal energy transferred from one
substance to another. Thermal energy is the total kinetic energy of all the particles of a substance,
whereas temperature is the average, or typical, KE of a particle in that substance. The speci c heat
of a material is how much energy is required to change 1 gram of it by 1K (or 1ºC). The equation
that relates energy changes to temperature changes is q = sm∆T. If a substance gains heat, its
temperature will increase; if it loses heat, its temperature will decrease

Test Yourself:
1. You put some chemicals in water and they react. The chemicals release energy and the water gains this heat. Will
the water decrease in temperature or increase in temperature?

2. A beaker of water loses 1000 cal of heat and its temperature drops from 30ºC to 10ºC. Find the mass of water
that you must have had in the beaker. (Look above if you forgot the speci c heat of water.

360 Chapter 10 Thermodynamics


fi

fi

fi

fi
fi
l

fi
fi
fi
:

fi
.

10.2 Enthalpy
In the last section, I talked about how knowing energy changes of reactions is important in
chemistry. Not only is it useful to know how much energy that, say, 10 grams of propane releases,
but knowing if a reaction releases or absorbs energy (and how much) can lead us to determine if
that reaction is likely to occur or not. We have a speci c name for the energy given off or absorbed
by a chemical reaction or process. This section de nes and describes it. Then in the next four
sections, we’ll learn four different ways to calculate it.

10.2.1 Enthalpy change is the measure of heat transfer


We now introduce a quantity called enthalpy (H). Enthalpy is closely related to internal energy, but
we won’t sweat the details.* So let’s stop talking about internal energy and start using enthalpy
from now on. We aren’t too concerned with the actual value of enthalpy. We are interested only in
the change in enthalpy, or ∆H. We de ne the change in enthalpy of a reaction as
∆Hrxn = qrxn

The change in enthalpy of a reaction is the heat gained or lost by the system (reacting chemicals).
For this reason, it is sometimes known as the change in heat of a reaction. Also, we often leave
off the words “change in” and say enthalpy of reaction or heat of reaction, although you should
always understand that change in implied.

You might be puzzled why we had to introduce a new term and symbol. Why not just use q? What’s If you want to know more about
the difference? For one, technically, ∆Hrxn is equal to the heat gained or lost by the system when enthalpy, read the next Interchapter,
the reaction occurs at constant pressure. Fortunately, that’s almost always the case anyways. You which goes into more detail on what
enthalpy really is and the difference
can disregard any differences. For our purposes, just think of ∆Hrxn as the q of a reaction. The only among ∆H, q, and internal energy.
important difference will be discussed in §10.3.

[*Note: You’ll recall I said chemical energy can manifest itself not only as heat but also as changes in volume as
product gases expand (“pressure-volume work”). So if we are saying ∆Hrxn is measuring q only, are we cheating
by ignoring PV work? Not really. PV work is usually negligible compared to thermal energy changes.

10.2.2 De nition of exothermic and endothermic


Keep in mind that ∆H is basically measuring how much heat is lost or gained by a system. While it We often write ∆H without rxn, but
is true that ∆H (change in enthalpy) is the important value, reactant and product chemicals do have we assume there’s an rxn subscript
enthalpy. Think of it as their energy state. Look at the diagram. Note that the reactants have more unless otherwise indicated.
enthalpy and the products have less enthalpy. That is,
reactants the system (chemicals involved in the reaction) has lost
enthalpy. If the chemicals of a reaction (the system)
enthalpy (energy)

heat flows to undergo a loss of enthalpy, we call it an exothermic


surroundings
get warmer
enthalpy lost

surroundings reaction. (Think of EXit: energy is exiting the system.)


Since the nal enthalpy is less than what it started with,
the value for ∆H would be negative. Okay, if the ∆H = H nal – Hinitia
products = Hproducts – Hreactant
chemicals lose energy, where does it go? To the
surroundings. Therefore, an exothermic reaction < 0 for exothermic rxns
SYSTEM SURROUNDINGS involves energy transfer from the system to the
surroundings. What, then, will happen to the temperature of the surroundings, which are in contact
with the system? It will increase. The clue for exothermic reactions is that we’d feel the container
getting warmer, an indication that the reaction is giving off energy.

If the chemicals of the reaction undergo a gain of enthalpy, we call it an endothermic reaction.
Since the nal enthalpy is more than what it started with, the value for ∆H would be positive. But if ∆H = H nal – Hinitia
the chemicals have gained energy, from where did it get it? From the surroundings. Therefore, an = Hproducts – Hreactant
endothermic reaction involves energy transfer from the > 0 for endothermic rxns
surroundings to the system. (Think of ENter: energy is
products
entering the system.) What, then, will happen to the
enthalpy (energy)

temperature of the surroundings, which are in contact


heat flows
surroundings
enthalpy gained

with the system? It decreases as the system steals


get colder

to system
energy from them. The clue for endothermic reactions
is that we’d feel the container getting cooler, an
reactants indication that the reaction is absorbing energy from our
hands (or thermometer) and turning it into enthalpy in
SYSTEM SURROUNDINGS the products

Chapter 10 Thermodynamics 361

fi
fi
fi
.

fi

fi
fi

A common source of confusion for endothermic reactions is thinking that if the system absorbs
energy from the surroundings, the system should get warmer. However, the energy it absorbs does
not remain as thermal energy. Instead, the heat the system steals from the surroundings turns into
enthalpy, which is a type of potential energy and is not “warmth energy.” We are part of the
surroundings. So when the system absorbs energy, all we feel is a drop in temperature due to the
thermal energy being stolen from us. Similarly, don’t think that an exothermic reaction should feel
cold because the system is losing energy. The reactants are releasing enthalpy, which turns into
heat (“warmth energy”) that we feel with our hands. Here is a summary. Please commit to memory

Process ∆H Direction of energy flow T of surroundings Example


exothermic – from system to surroundings increases fuel burning
endothermic + from surroundings to system decreases ice melting

10.2.3 Enthalpy of reaction is heat transfer associated with a reaction


A reaction will always give off or absorb the same amount of heat for a particular amount of
reactant used or product formed. Whenever you burn one mole of carbon to produce one mole of
carbon dioxide gas, 393.5 kJ of energy are released. Again, this is known as the enthalpy of
reaction, ∆Hrxn (aka heat of reaction). The enthalpy of reaction is the amount of heat transferred
during a reaction when you use (and produce) the number of moles given by the coef cients of the
balanced equation.

We can write this two ways. The rst way is to treat this energy as a reactant or product. Because
energy is released (produced) when you burn carbon, we will put it on the product side
C(graphite) + O2(g) ➞ CO2(g) + 393.5 k

The other way is to write the enthalpy of reaction next to the balanced chemical equation. Keep in
mind that this is an exothermic reaction, so the sign of ∆H is negative (see §10.2.2):
The unit of ∆H should be energy per C(graphite) + O2(g) ➞ CO2(g) ∆Hrxn = –393.5 k
mol (e.g. kJ/mol). We’ll often simplify
it by writing only kJ, but it is assumed These are called thermochemical equations. They both say, “393.5 kJ are released when exactly
that this is the energy when you use 1 mole of graphite reacts with exactly 1 mole of oxygen to produce exactly 1 mole of CO2.”
and make the number of moles given
by the balanced equation.
The gure here shows this schematically. The products are lower in
energy than the reactants. In going from the reactants to the enthalpy (energy) reactants
products, 393.5 kJ of energy must be released. You can think of it
like a ball being dropped from a hill. The ball on top of the hill

393.5kJ
(where the reactant level is) has potential energy. When it falls to
the bottom of the hill (where the product level is), it loses 393.5 kJ
of potential energy, which turns into KE, heat, etc. The same
applies to chemical reactions. In going from the top of the enthalpy
products
“hill” to the bottom, enthalpy is lost, which turns into heat that is
transferred to the surroundings

As another example, consider this endothermic reaction between boron and hydrogen, written two
ways
Again, this should be kJ/mol, but that 10B(s) + 9H2(g) + 125.6 kJ ➞ 2B5H9(g)
could cause confusion: 125.6kJ per
mole of what? It’s actually “per mole 10B(s) + 9H2(g) ➞ 2B5H9(g) ∆Hrxn = +125.6 kJ
of reaction,” meaning per 10 mol B, 9
mol H2, and 2 mol B5H9. This is
implied so we’ll leave off the “/mol.” In the rst equation, I wrote the energy as one of the reactants,
because it is being absorbed by the system during the course of the
enthalpy (energy)

reaction, turning into enthalpy. Like a reactant, it is used to help form products
the products. Also notice that the value of ∆Hrxn in the second
equation is positive (as it should be for endothermic reactions). Both
125.6kJ

of these equations say, “When 10 moles of boron react with 9 moles


of hydrogen to produce 2 moles of borane (B5H9), 125.6 kJ of energy reactants
will be absorbed from the surroundings.” What will happen to the
temperature of the surroundings? You better say they decrease! This
reaction is shown in the diagram. Notice how the reactants are
always written on the left and the products on the right

362 Chapter 10 Thermodynamics


fi
:

fi

fi
J

fi
:

As a caveat, when writing equations and determining heats of reactions, pay careful attention to the
physical states of all the reactants and products. Consider the production of CO2 shown above. If I
made the carbon dioxide from the diamond allotrope of carbon instead of graphite (remember that
allotrope is a fancy chem word for a particular element form), the enthalpy of reaction would be
different (and the reaction would have been a lot more expensive!)

Example 2:
Consider this reaction A + B ➞ C + 100k
a. Is it endothermic or exothermic
b. Would we write ∆Hrxn = +100kJ or ∆Hrxn = –100kJ
c. Which way is heat owing: from the system to the surroundings or vice versa
d. Would the temperature of the container in which the reaction is occurring increase or decrease?
Solution:
a. Because the energy is written as a product, it is producing (i.e. releasing) heat. It is exothermic
b. Exothermic reactions have negative enthalpies of reaction, so ∆Hrxn = –100kJ
c. If it is releasing energy, it is traveling from the system to the surroundings
d. If the surroundings are gaining energy, we’ll measure a temperature increase

Example 3:
Write a thermochemical equation that represents the
reaction indicated by this diagram.
Solution: 2A + B
Enthalpy (Energy)
The reactants are always on the left and the products on the
right. Notice that the products are lower in enthalpy. That
means that the system loses energy, indicating an exothermic

250kJ
reaction. (And as a result, energy is transferred to the
surroundings, increasing the temperature.) We can write a
thermochemical equation in two ways for this reaction A2B
2A + B ➞ A2B + 250kJ o
2A + B ➞ A2B ∆Hrxn = –250k

10.2.4 When you change the chemical equation, you must change ∆H
An equation in the last subsection said that when one mole of graphite combusts with one mole of
oxygen gas to produce one mole of carbon dioxide gas, 393.5 kJ of heat are released. How much
energy would be released if two moles of graphite combust with two moles of oxygen gas to
produce two moles of carbon dioxide gas? Twice as much, of course. We would writ
2C(graphite) + 2O2(g) ➞ 2CO2(g) ∆Hrxn = –787 k
What if instead of combusting graphite, we do the opposite reaction: we decompose the carbon
dioxide gas into graphite and oxygen gas? What would be the enthalpy of reaction? Well, if 393.5
kJ of energy are released in forming CO2, 393.5 kJ of energy must be absorbed to break it apart.
We would writ
CO2(g) ➞ C(graphite) + O2(g) ∆Hrxn = +393.5 k
o
CO2(g) + 393.5 kJ ➞ C(graphite) + O2(g)
In summary
★ If you multiply the coef cients of an equation by a constant, multiply ∆H by that same constant.
★ If you reverse an equation, reverse the sign of ∆H.

Example 4:
Consider the following reaction CD + 300kJ ➞ C +
What would be the ∆Hrxn for this reaction 2C + 2D ➞ 2CD
Solution:
The rst equation is an endothermic reaction that has a ∆Hrxn = +300kJ. What is the relationship between
the rst and second equations? For one, the second equation is reversed. Therefore we’d need to reverse
the sign of ∆Hrxn. Additionally, the second equation has twice the coef cients, so we’d need to multiply the
∆Hrxn by 2. Taking both into account, the second equation would have ∆Hrxn = –600kJ. Let’s think about it
conceptually. The rst equation says that 300kJ of energy are absorbed when one mole of CD decomposes. If
instead we synthesize 1 mole of CD, it would release 300kJ. But we’re actually synthesizing two moles of CD,
so it would release a total of 600kJ.

Chapter 10 Thermodynamics 363


r

fi
fi
:

fi

:
fl
fi
r

?
J

fi
.

Section Overview

The Forest:
The enthalpy of reaction is a measure of how much heat is lost or gained. An exothermic process is
one in which heat is lost. An endothermic process is one in which heat is gained.

The Trees:
The change in enthalpy of a reaction (aka enthalpy of reaction or heat of reaction), ∆Hrxn, is how
much heat the reacting system gives off or absorbs. If the system decreases in enthalpy during a
reaction (“exothermic”), ∆H is negative and the system loses heat to the surroundings, resulting in
an increase in temperature for the surroundings. If there is an enthalpy gain (“endothermic”), ∆H is
positive and the system absorbs heat from the surroundings, resulting in a decrease in temperature
of the surroundings (we’d feel things get colder). The enthalpy (or heat) of reaction is the heat
gained/lost when using the number of moles speci ed by the balanced equation. If you change a
chemical equation, you need to correspondingly change the value of ∆Hrxn. If you multiply the
coef cients by a certain factor, you must multiply ∆H by that same factor. If you reverse the
balanced equation, you must reverse the sign of ∆H

Test Yourself:
1. Given the following reaction A + 2B + 200kJ ➞ AB2
a. Is it endothermic or exothermic
b. What is ∆Hrxn
c. Which way is heat traveling: from the system to surroundings, or vice versa
d. Would the temperature of the surroundings increase or decrease

2. Consider the following reaction C + D ➞ CD ∆Hrxn = –400k


What would be the ∆Hrxn of this reaction 3CD ➞ 3C + 3

10.3 Finding ∆H by Calorimetry


This will be the rst of four sections that describe how enthalpy change can be calculated. In
addition to the relevance and usefulness of ∆H mentioned in these sections, we’ll learn later in the
chapter that ∆H can help tell us if a reaction is likely to occur at all. Alrighty then...let’s get started

10.3.1 ∆Hrxn vs. q


Our rst method of determining the value of ∆Hrxn is to do it experimentally. But before we start
describing this process, let’s discuss the subtle difference between ∆Hrxn and qrxn. To demonstrate,
let’s say that you react 24 grams of graphite in the presence of hydrogen gas to produce methane
C(gr) + 2H2(g) ➞ CH4(g)
You perform the reaction and calculate that the surroundings gain 150 kJ of heat. Remember that
we use the symbol qrxn for the heat gained or lost by a reaction. Since the reaction lost heat to the
surroundings, this is an exothermic reaction. Thus, we’d write qrxn = –150 kJ. Exothermic reactions
have a negative ∆Hrxn and a negative qrxn. Endothermic reactions, in which the system gains heat
from the surroundings, have a positive ∆Hrxn and a positive qrxn. Just remember ∆Hrxn and qrxn have
the same sign

Okay, if qrxn = –150kJ, is that also the value of ∆Hrxn? No. “What?!” you shout. “You told us ∆H is the
heat lost/gained by the system. Why do you lie to us? We trust you with our young, impressionable
minds.” Hold on. Yes, what I said earlier is true; but as mentioned in §10.2.3, technically ∆Hrxn is the
qrxn you’d measure if you used the number of moles given by the balanced chemical equation.
Speci cally for this reaction, ∆Hrxn is the qrxn you would get if you used 1 mole of carbon (see the
equation above). How many moles of carbon did we use? Carbon has a molar mass of 12 g/mol,
and we used 24 grams, so we actually reacted 2 moles of carbon. Okay, if 150 kJ of heat were
released by using 2 moles of carbon, how much would have been released if we used 1 mole? Half
Don’t forget the negative sign: it’s an that: 75 kJ. Therefore, we can nally say that ∆Hrxn = –75 kJ.
exothermic reaction. Remember to
always think about what sign your
∆Hrxn should have before “boxing
your answer.”

364 Chapter 10 Thermodynamics


fi
fi
fi

fi

:
:
?

fi
?

fi
.

We de ne ∆H this way – on a sort of “per mole” basis related to the coef cients of the balanced
equation – because we want a standardized value for reactions. By saying that the chemical
reaction C(gr) + 2H2(g) ➞ CH4(g) has a ∆Hrxn = –75kJ, we know that reacting one mole of graphite
with hydrogen releases 75kJ, reacting ten moles of graphite releases 750kJ, etc. If we didn’t do
this, everyone would report different values of ∆Hrxn, depending on how much graphite they used.
In other words, q is the value you obtain experimentally using whatever amount of reactants you
want. ∆H is the value standardized by the coef cients of the balanced equation

Let’s try another one. Consider the production of glucose in photosynthesis


6CO2(g) + 6H2O(l) ➞ C6H12O6(s) + 6O2(g)
By reacting 88 grams of CO2, the system absorbs 934 kJ of energy. What is the ∆Hrxn for the
reaction written above? Well, we know that it is endothermic and qrxn = +934 kJ. But ∆Hrxn is the qrxn
you would get if you used the amount of material shown by the balanced equation. How many
moles of CO2 did we use? Because each mole of CO2 weighs 44 grams, we used two moles of
CO2. The ∆Hrxn would be the qrxn we would get by using six moles of CO2 (see balanced equation).
Therefore, it would absorb three times as much energy if we used six moles instead of two: ∆Hrxn =
+2802 kJ (the positive sign indicating endothermic). We could even use a proportion to nd this.
934kJ is to 2 moles as “x” is to 6 moles, where “x” would be the ∆Hrxn
934kJ x
=
2mol 6mol

Example 5:
Consider the following reaction: A + 3B ➞ AB3

By reacting 0.1 mole of A, 30kJ of heat are released.What are qrxn and ∆Hrxn? (include the signs)
Solution:
By de nition, the reaction is exothermic since heat is being released. Therefore, the sign of both qrxn and
∆Hrxn are negative. qrxn = –30 kJ. However, we got this 30kJ by using only one-tenth of a mole of A. The
balanced equation says we should be using one mole of A. I think you can understand that if we used one
mole of A, we’d get ten times as much energy released. Therefore, ∆Hrxn = –300 kJ.

Example 6:
Consider the following reaction 2Al2O3 ➞ 4Al + 3O2
You perform the reaction using 41 grams of aluminum oxide. In doing so, 674 kJ of energy are absorbed.
Find the ∆Hrxn for this reaction as written.
Solution:
First, we know that the reaction is endothermic (∆Hrxn is positive) and qrxn = +674kJ. Now let’s nd out how
many moles we reacted. The molar mass of Al2O3 is 102g/mol. Therefore, we used 41/102 = 0.402 mol of
Al2O3. The reaction absorbed 674kJ when using 0.402 mol; ∆Hrxn is how much energy would be absorbed if
we reacted two moles (see balanced equation). Using a proportion
674kJ x
=  
0.402mol 2mol
This says 3353 kJ would be absorbed by the system if we used 2 moles of Al2O3. Thus, ∆Hrxn = +3354 kJ


10.3.2 Using calorimetry to nd ∆H
One way to nd ∆H is by calculating q when performing a reaction in a calorimeter, a device I
described in §10.1.6. We would place a measured amount of water in the calorimeter and record its
initial temperature. Then we put in it the chemicals we want to react. If the chemicals undergo an
exothermic reaction, for example, they release heat. What is the rst thing that absorbs that heat
released by the chemicals? The solution (water + dissolved chemicals). This will cause the
solution’s temperature to rise. We would then record the nal temperature of the solution. From the
solution’s change in temperature, its mass, and its known speci c heat (4.184 J/gK), we can use We’ll assume the aqueous solution is
the equation q = sm∆T to calculate the amount of heat that was absorbed by the solution. We then mostly H2O, so we’ll assume it has a
specific heat about the same as that
use this q to nd ∆H.
of liquid water.

That sounds confusing, but it ain’t so bad. Here’s how you do it, Hoss…

Chapter 10 Thermodynamics 365


fi
fi
fi
fi

:
fi

fi
fi
:

fi
fi
.

fi

fi
fi
.

Let’s say you have a calorimeter containing 90 mL of water initially at 20ºC. Then we dissolve 4
grams of sodium hydroxide.
NaOH(s) ➞ Na+(aq) + OH–(aq)
As a result, the water’s temperature increases to 31ºC. We rst nd q, the heat change of the
solution (water and dissolved chemicals). The mass of the entire solution is 94 g: the combined
mass of the water (90mL of liquid water has a mass of 90g) and dissolved NaOH (4g). The ΔT of
Changes are always “final minus the solution is 11ºC. We’ll assume that the speci c heat of the solution – because it is dilute – is the
initial,” so a temperature increase is same as that of pure water: 4.184J/gK. The heat gained by the solution i
positive. However, let’s not worry
about the sign of ∆T. At the end,
q = sm∆
we’ll use logic to slap a + or – on q q = (4.184J/gºC)(94g)(11ºC) = 4326 J = 4.326 k
and ∆H. That’s the important part.
But wait! Let’s get the correct sign for q. Because the solution increased in temperature, it must
have been exothermic. So q is going to be negative, and the correct answer is q = –4.326 kJ.
We’re almost home! This was the heat lost by using only 4 grams of NaOH, which is 0.1 mol (the
molar mass of NaOH is 40g/mol). Since the balanced equation shows 1 mole of NaOH reacting,
−4.326 kJ ∆Hrxn
Technically, dissolving a solute into a the ∆Hrxn will be the q if we used one mole: = Therefore, ∆Hrxn = –43.26 kJ.
solvent is not a reaction but a 0.1 mol 1 mol
physical process. However, we’ll
ignore that and still call this ∆Hrxn. That’s the general strategy to nd ∆Hrxn experimentally. Use the mass of solution in a calorimeter
and its ∆T to nd the heat gained/lost (q). Then use the number of moles of reactant to nd ∆Hrxn

Example 7:
Calcium chloride dissolves in water as shown here: CaCl2 ➞ Ca2+ + 2Cl–
If 3 grams of calcium chloride dissolve in a calorimeter lled with 50 grams of water at 18ºC, the
solution’s temperature increases to 28.1ºC. What is the ∆Hrxn?
Solution:
Let’s nd how much heat the solution gained. Note that ∆T = 10.1ºC (went from 18ºC to 28.1ºC). We must
also use the total mass of the solution, which is 53 g (50g of water + 3g of CaCl2)
q = sm∆
q = (4.184J/gºC)(53g)(10.1ºC) = 2240 J = 2.24 k
Since temperature increased, it is an exothermic reaction. Therefore, qrxn really is –2.24 kJ. Lastly, we want to
There is a more technical way of know qrxn if we used 1 mole of CaCl2, not just 3 grams. The molar mass of is 111.1 g, so 3 g = 0.027 mol.
doing calorimetry problems in which Now set up a proportion
you keep track of the heat gained or −2.24 kJ ∆Hrxn
lost by the system and by the =
surroundings separately, but we’ll
0.027 mol 1 mol
simplify it. You get the right answer This results in ∆Hrxn = –83.0 kJ. Yep, the negative should be there since the temperature increase indicates
without the mathematical confusion. it is an exothermic reaction. Don’t forget to think about your +/– signs for q and ∆H

Here’s a more dif cult problem. Let’s say we mix one liter of a 0.5-M HCl solution and one liter of a
0.5-M NaOH solution in a calorimeter and the temperature increases by 2.3ºC. What is the ∆Hrxn for
the reaction that occurs
HCl(aq) + NaOH(aq) ➞ H2O(l) + NaCl(aq) ∆Hrxn =
Where’s the information about the mass of water/solution? Are we lost? No. Remember that an
aqueous solution is simply a lot of water with a little bit of something dissolved in it. So when we
said we are mixing aqueous solutions of HCl and NaOH, there is water galore! When the HCl and
NaOH react, they’ll transfer heat to the solution all around them. We still use q = sm∆T. Since these
are dilute aqueous solutions (i.e. mostly water), we’ll assume that the speci c heat and density of
the solution are the same as those for water (4.184 J/gK and 1 g/mL, respectively). Let’s do it

The mass we should use is the total mass of the aqueous solution (that’s what’s changing
temperature). If we mixed 1L of HCl and 1L of NaOH, we have 2L of solution. What is the mass of
2000mL of solution, assuming that it has a density of 1 g/mL? It will be 2000g. Therefore
q = (4.184J/gºC)(2000g)(2.3ºC) = 19246J = 19.2k
What sign should it have? Because the temperature increased (again!), it is exothermic and should
be negative. Therefore
qrxn = –19.2 k
However, this is the amount of heat transferred when we used the amount of reactants we did. The
heat of reaction we are supposed to nd is for the number of moles given by the coef cients in the
equation above: 1 mole of HCl and 1 mole of NaOH. How many moles did we actually use?

366 Chapter 10 Thermodynamics


fi
T

fi
J

fi
,

fi
fi
J

fi
J

fi
fi
fi
s

fi
.

fi
fi
,

Since M = mol/L, then #mol = MV Remember that equation? The


For both reactants, #mol = (0.5mol/L)(1L) = 0.5mo number of moles of solute dissolved
in a solution is its molarity times the
That is, by using 0.5 mole of each reactant, 19.2 kJ of heat were released. How much heat would volume in liters. I told you you’d be
using it a lot!
be released if we had used 1 mole of each (as the coef cients of the equation speci es)? Twice as
much, or 38.4 kJ. So we can now write the thermochemical equation with the correct ∆Hrxn
HCl(aq) + NaOH(aq) ➞ H2O(l) + NaCl(aq) ∆Hrxn = –38.4k
Note that the value is negative because we said the temperature increased by 2.3ºC, indicating that
it was an exothermic reaction. (Of course, qrxn and ∆Hrxn always have the same sign.

No, this isn’t an easy problem to do. And, yes, if you get such a problem correct, you are a total
chemistry stud or studette.

10.3.3 Using ∆Hrxn to calculate energy changes


We have spent time calculating ∆Hrxn from experimental data. But what good is the value to us? For
one thing, it would let us know how much heat will be given off (or absorbed) when a certain
reaction takes place. This would be good, for example, if we needed to produce a certain amount of
heat and wanted to know how much reactant we needed.

Example 8:
The combustion of graphite has a ∆Hrxn = –393.5kJ C(gr) + O2(g) ➞ CO2(g)
If 120 grams of graphite combust, how much energy would be absorbed or released?
Solution:
First off, since ∆Hrxn is negative, we’ll be nding how much energy is released. Now how much? The balanced
equation and associated ∆Hrxn say, “393.5kJ of energy are released when one mole of graphite combusts.”
However, we’re using 120 grams of graphite, which is 10 moles. Therefore, the reaction will release
3935kJ of energy.

Example 9:
Consider the decomposition of hydrazine: N2H4(g) ➞ N2(g) + 2H2(g) ∆Hrxn = –50.4 kJ
If we need to produce 1 megajoule (1,000,000 J) of heat, how many grams of hydrazine are needed?
Solution:
The ∆Hrxn tells us it gives off 50,400 J for every 1 mole of N2H4 that reacts. If we need 1,000,000 J, we
require 1,000,000/50,400 = 19.8 moles of N2H4. Its molar mass is 32 g/mol, so we need 635 g of N2H4

If it were occurring in a calorimeter, we would be able to calculate the change in temperature we’re
expecting. It’s like doing Example 7 in reverse. Take a looksie..

Example 10:
The dissolution of calcium chloride has a ∆Hrxn = –83.0kJ CaCl2 ➞ Ca2+ + 2Cl–
If 11.1 grams of calcium chloride dissolve in a calorimeter lled with 100 grams of water, what will be
the change in temperature of the solution?
Solution:
Let’s translate into words what the equation and ∆H are telling us: “83kJ of heat are released when one mole
of CaCl2 dissolves.” But we’re not using one mole, are we? The molar mass of CaCl2 is 111 grams. 11.1
grams represents 0.1 mole. This will release only one-tenth as much energy: qrxn = –8.3kJ. We can plug
everything into q = sm∆T and solve for ∆T. The total mass is 111.1g (100g H2O + 11.1g CaCl2). We’ll
assume the speci c heat of water.
q = sm∆
8,300 J = (4.184J/gºC)(111.1g)∆
∆T = 17.9ºC
Did you notice that I did not use –8300J? Why did I drop the negative sign? Earlier in a sidebar note, I said Read this paragraph.
we are simplifying these problems by not worrying about the sign of ∆T. We still want to ignore it. I just
want a value of ∆T without a sign, so I don’t care at this point about q being negative. Just put in a positive
value and get a positive value for ∆T. Then we can nish the problem logically without worrying about the
sign of ∆T. Here’s how you nish. Note that this was an exothermic reaction. So there will be an increase in
temperature of the solution. Thus, the answer is that the temperature increases by 17.9ºC

Chapter 10 Thermodynamics 367


T

fi

fi

fi
fi

fi
J

fi
.

fi
.

Section Overview

The Forest:
Calorimetry is an experimental method in which reactions occur in an insulated device. By
measuring changes in temperature inside the calorimeter, we can calculate qrxn and ∆Hrxn

The Trees:
We can nd ∆H by using a calorimeter and q = sm∆T. By measuring the mass of water and its
temperature change due to a reaction occurring in it, we can nd the heat gained or lost (q). (We’ll
assume no heat is lost to or gained from the surroundings.) From this, we can calculate ∆Hrxn,
remembering that ∆Hrxn is the q you would achieve by using the number of moles speci ed by the
balanced equation. We can also work backwards from a given value of ∆Hrxn to nd the energy
given off or absorbed by a reaction

Test Yourself:
1. Consider the following reaction A + B ➞ A
By reacting 4 moles of A, 100 kJ of heat are transferred from the surroundings to the system
a. Is the reaction endothermic or exothermic
b. If the reaction occurred in a calorimeter, would the temperature of the solution increase or decrease
c. What is the value of qrxn? (Remember the sign.
d. What is the value of ∆Hrxn? (Remember the sign.

2. You perform the following reaction E + 2F ➞ EF2 ∆Hrxn = +200k


Given the value of the heat of reaction, how much energy the system would give off or absorb (state which one) if
0.25 mole of E reacts

3. You perform the following reaction NH4Cl ➞ NH4+ + Cl–


There are 30 grams of water in a calorimeter initially at 25ºC. After dissolving 3.1 grams of NH4Cl, the nal
temperature of the water is 19ºC
a. What is the total mass of solution used?

b. Calculate q (Assume it is dilute enough that s = 4.184 J/gK). Remember to add the appropriate sign.

c. Assuming no heat has been lost to the surroundings, calculate ∆Hrxn. Remember to use the appropriate sign. (Hint:
you’ll need to nd the number of moles of NH4Cl. Its molar mass is 53.5 g/mol.

10.4 Finding ∆H by Hess’s Law


10.4.1 State functions depend only on the initial and nal states
Imagine that you are trying to drive up to the peak of a mountain. You have two paths. From the left
1000 m

side, you have a long path to travel, going up and down along the way. From the right side, you
have a direct and steep path up to the top.

What is the difference in height between the starting point and the peak? ∆h = 1000m. It doesn’t
depend on whether you take the long way up and down and up and down and up or you take the
short route. The difference in height is still going to be 1000m. It doesn’t matter how you got to the
top: your net change in height is the same. We would call change in height a state function; that
is, it is a property that depends only on the initial and nal states of a system and not how it got
there. How about the amount of fuel: is that a state function? Does the amount of fuel you use to
get to the top of the mountain depend only on the initial and nal amounts of fuel in your gas tank?
No. You might have stopped to ll up the tank on the way. In fact, you might even have more fuel in
your tank at the end. It matters what you did on your way to the top

368 Chapter 10 Thermodynamics


fi

fi

:
.

:
:
fi
.

fi
fi
B

fi
fi

fi

fi
.

fi
10.4.2 Enthalpy is a state function
Enthalpy is another property that depends only on its initial and nal states. This means that ∆H for
a reaction depends only on the difference between the enthalpy of the products and reactants. It
doesn’t matter how we get from the reactants to the products: ∆H will always be the same. We can
take a quick direct route, or we can take some winding, scenic route from the reactants to the
products, but ∆H will be the same. This freedom of how we get from A to B allows us to nd ∆Hrxn
values without having to perform experiments such as using a calorimeter.

10.4.3 Finding ∆Hrxn using Hess’s Law


In §10.3, we saw how we could nd a heat of reaction experimentally. Now we are going to nd it
by an indirect method. Let’s take the generic reactio
(1) A+B➞D ∆Hrxn =
We want to nd ∆Hrxn for this reaction as written. Let’s say we know the ∆Hrxn values for the
following two reactions (by looking them up in a reference book)
(2) A+B➞C ∆Hrxn = +200k
(3) C➞D ∆Hrxn = +350k

How does this help us? Let’s look at an energy level diagram.
Equation (2) says that C is 200kJ higher in energy than A and
D
B (a positive ∆H means it is endothermic, which means that the
products have more enthalpy). Equation (3) says that D is

350kJ
350kJ higher in energy than C. I have shown this schematically
in the gure
Enthalpy

550kJ
As you can see, the difference between A+B and D is 550kJ. C C
So we can writ
200kJ

A + B ➞ D ∆Hrxn = +550

Another way to think of it is that instead of making D directly A+B


from A and B, we took a different, longer, more scenic route.
We rst used A and B to make an intermediate product, C. Then we took this C and made D out of
it. The net result: we took one A and one B and made one D (any C we made disappears when it
makes D). Does the reaction really occur like this? Who cares? As long as we get from A and B to
D, the ∆Hrxn will be the same. In effect, we are saying, “It takes 200kJ to turn A and B into C, and
then it takes an additional 350kJ to turn C into D, for a total of 550kJ.”

This method is known as Hess’s Law, which states that the enthalpy change for a reaction is the
sum of the enthalpy changes for a series of reactions that add up to the overall reaction. Um, right.
That doesn’t make too much sense yet. Let’s use the last example to show how this works without
drawing energy level diagrams. Instead, we’ll just use balanced equations. To make it more
general, we’ll just use the symbols ∆Hx instead of actual values. We are trying to get the equatio

(1) A+B➞D ∆H1

from the following two equations


(2) A+B➞C ∆H2
(3) C➞D ∆H3

How can we manipulate Equations (2) and (3) so that they add up to Equation (1)? Just add the
two
A + B ➞ C ∆H2
C➞D ∆H
A + B ➞ D ∆H1 = ∆H2 + ∆H3

In the problem, note that the two C’s cancel on opposite sides of the equations to give us the
desired equation. It is important to remember that if you add two equations to get a third, the ∆H of
the third is equal to the sum of the ∆H’s of the rst two. This example was a simple version. Here’s
a more dif cult one

Chapter 10 Thermodynamics 369


:

fi
fi
fi
.

fi
e

fi

fi
n

fi

fi
fi
n

Example 11:
Find the enthalpy of reaction for the following reaction:
2 H2O2(l) ➞ 2 H2O(l) + O2(g) ∆Hrxn =
given the following information
(1) 2 H2(g) + O2(g) ➞ 2 H2O(l) ∆H1 = –572k
(2) H2(g) + O2(g) ➞ H2O2(l) ∆H2 = –188k
Solution:
Our goal is to add the last two equations in a way to get the equation we want. Along the way, we might
have to manipulate the equations (multiplying them, reversing them). Be careful to change the value of ∆H
appropriately.

Okay, the easiest way to start is to look at the equation we are trying to get for hints. Notice that we need 2
moles of H2O2 on the left side. We can get this by using Equation (2). First, we need to multiply the
We learned about these rules in coef cients by 2. That means we need to multiply the value of ∆H2 by two as well. Second, we need to write
§10.2.4. Equation (2) in reverse in order to get the 2H2O2 on the left. Remember that when we write an equation in
reverse, we have to change the sign of the ΔH for the equation. So overall, we need to multiply ΔH2 by –2:
2H2O2(l) ➞ 2H2(g) + 2O2(g) ∆H = (–2)∆H2 = +376k
Okay, our left side is good, but we want “2H2O(g) + O2(g)” on the right, not “2H2(g) + 2O2(g).” How can we
get some H2O on the right? We can use Equation (1) as written. In fact, it has just the right number of H2O’s
that we need. We don’t need to do anything to the equation, so we don’t need to do anything to its ∆H
2H2O2(l) ➞ 2H2(g) + 2O2(g) ∆H = (–2)∆H2 = +376k
2H2(g) + O2(g) ➞ 2H2O(l) ∆H1 = –572kJ
2H2O2(l) ➞ 2H2O(l) + O2(g) ∆Hrxn = +376kJ –572kJ = –196kJ
Note how conveniently the H2’s cancel on each side and how the O2 from the left cancels one of the O2’s
on the right, so that we get the equation we wanted. Since we added those two equations, we need to add
their ∆H values, as shown

Section Overview

The Forest:
Hess’s Law is a way to determine a ∆Hrxn without having to perform experiments. It involves adding
other equations, whose ∆H‘s are known, to make the equation of interest

The Trees:
Other people have already performed calorimetric experiments to determine heats of reaction for
many different reactions. These can be found in a reference book or textbook. If we want to know a
∆Hrxn for a different reaction, we can go about it mathematically instead of experimentally. Hess’s
Law states that if two equations add up to give you a third equation, the ∆H of that third equation is
the sum of the ∆H’s of the rst two. We just manipulate known equations to “build” the equation we
want. Just keep in mind that we need to make the corresponding changes to an equation’s ∆H, as
learned in Section 10.2.4

Test Yourself:
Determine the enthalpy of reaction, ∆Hrxn, for this reactio H(g) + Br(g) ➞ HBr(g
given the following information
H2(g) ➞ 2H(g) ∆H = 436.4 k
Br2(g) ➞ 2Br(g) ∆H = 192.5 k
H2(g) + Br2(g) ➞ 2HBr(g) ∆H = –72.4 k

370 Chapter 10 Thermodynamics


fi

fi
:

10.5 Finding ∆H by Standard Enthalpies of Formation


10.5.1 The standard enthalpy of formation de ned
In §10.2, we came up with the idea of enthalpy, de ned in way that makes the change in enthalpy
simply the heat transferred between the system and surroundings. The ∆Hrxn value can change if,
for example, you vary the temperature and pressure of the experiment, or if you use a different
allotrope as a reactant (e.g. diamond instead of graphite). It would be good to have some
standardized set of conditions: not only for comparison, but – as you will see – they will become
useful for determining values of ∆H for other reactions. We de ne the standard enthalpy of
formation, ∆Hfº, as the change in enthalpy when one mole of a compound is formed from its
elements in their most stable forms at the standard state of 1 atm of pressure and 25ºC. The Note that the standard temperature
subscript “f” denotes that it is the heat of formation, and the superscript “º” denotes that it is for thermodynamics is 25ºC whereas
the standard temperature for gas
occurring at the standard conditions. In effect, ∆Hfº tells you how much heat is required or released laws is 0ºC.
when you form a mole of a compound from its elements at this temperature and pressure.

10.5.2 How to write the correct chemical equation for ∆Hfo


We’ve already seen an example of a thermochemical equation for a standard enthalpy of formation
reaction, although we didn’t call it that at the time. We wrote an equation for the synthesis of
gaseous carbon dioxide from the elements that make up CO2: carbon and oxygen. Moreover, the
forms of carbon and oxygen we used in the equation are in their most stable states: how they
normally occur at 1 atm and 25ºC. For carbon, “normal” is a solid, and more speci cally the
graphite allotrope. For oxygen, “normal” is a diatomic gas

C(graphite) + O2(g) ➞ CO2(g) ∆Hfº = –393.5 kJ/mo

This is called a standard enthalpy of formation equation, or sometimes simply a formation


equation. It says that the 393.5 kJ of heat are released when you form one mole of CO2 gas from Recall negative ∆H = exothermic =
its elements in their most stable states at 1 atm and 25ºC. Note that the following reactions are not energy given off = temperature of
equations representing the standard enthalpy of formation of CO2 gas surroundings increases

C(graphite) + O2(l) ➞ CO2(g) ∆H ≠ –393.5 kJ/mo


C(diamond) + O2(g) ➞ CO2(g) ∆H ≠ –393.5 kJ/mol
C(graphite) + 2O(g) ➞ CO2(g) ∆H ≠ –393.5 kJ/mo
CO(g) + ½O2(g) ➞ CO2(g) ∆H ≠ –393.5 kJ/mo

The rst one is incorrect because you are making CO2 from oxygen in a state (liquid) it wouldn’t be
in at 1 atm and 25ºC. The second is wrong because although carbon is still a solid, it isn’t in its
most stable state. The graphite allotrope is more stable than the diamond allotrope. The third is
wrong because oxygen needs to be in the diatomic gaseous form at the standard conditions. The
fourth is wrong because it isn’t making the CO2 gas from elements (it’s using a molecule, CO). All
four of these would have a different ∆Hrxn than –393.5 kJ/mol

There is one more wrong way I should point out. When we learned to balance chemical equations,
we didn’t want to use fractional coef cients. For example, the production of water from hydrogen
and oxygen gas is given by the equatio

2H2(g) + O2(g) ➞ 2H2O(g)

However, our de nition of the standard enthalpy of formation speci es that it is the heat change
when one mole of product is formed. The correct equation to use for standard enthalpies of
formation would b
H2(g) + ½O2(g) ➞ H2O(g)

The ∆H measured in this reaction would be equal to ∆Hfº, whereas the previous reaction would give
a ∆H that is twice as big. Remember that fractional coef cients are okay when balancing
thermochemical equations because we’re dealing with moles, not individual molecules

10.5.3 The ∆Hfo of elements in their most stable states is zero


How do we “make” elements, like carbon? Is there a ∆Hfº for this element? How would we write the
thermochemical equation? Maybe heat was transferred in forming a mole of graphite from protons,
neutrons, and electrons. But who cares? Remember that we are interested in changes in energy for
a reaction. We want to know the difference in energy between the products and reactants. The

Chapter 10 Thermodynamics 371


fi
fi
e

fi
n

fi




fi

fi

fi
fi
:

fi

difference in enthalpies between CO2 and C+O2 gives us the ∆H for the reaction. It doesn’t matter if
it took one joule or a billion joules to make the carbon and oxygen, the enthalpy of CO2 is always
393.5 kJ smaller. So for convenience, we arbitrarily assign a value of ∆Hfº = 0 for the standard
enthalpy of formation of any element in its most stable state at 1 atm and 25ºC. This will become
very useful later on!

Are you not convinced about this whole ∆Hfº = 0 thing? Let’s say we want to know the difference in
altitude between the peak of a mountain and some base camp part way up. We could nd the exact
altitude of the base camp relative to sea level, then measure the exact altitude of the peak relative
to sea level, then subtract. But why? Why not de ne the base camp as our “zero” and measure the
difference in height from the base camp to the peak? It’s much more convenient. We are not saying
that the base camp has zero altitude – just that we are using it as a point of reference. In the same
way, we are not saying that an element has no enthalpy or that it took no energy to produce it – just
that we are using it as a point of reference…a zero of energy.

“Hi. It’s Simpsonized Doc. Be careful what you de ne as having zero ∆Hfº. It has to be the elements
in their “usual” states at “normal” conditions (1 atm and 25ºC). Almost all elements are in their most
stable states as monatomic atoms in the solid state. For example, Na(s), Cu(s), etc. There are a
handful of exceptions you can memorize
★ the noble gases are, of course, in the gaseous stat
★ the BrINClHOF elements are in the diatomic form
★ the BrINClHOFs are gases except bromine, which is a liquid, and iodine, which is a soli
★ mercury is a liqui
★ the most stable allotrope of carbon is graphit
Think you’ll forget? There is an easy way of nding out what the most stable state of an element is.
Look at your table that lists ∆Hfº values. If you nd a ∆Hfº = 0 in the table, it must be for an element
in its most stable form. C(diamond) has ∆Hfº = 1.9 kJ/mol listed in the table, whereas C(graphite) has
∆Hfº = 0 listed. Which is the stable version of carbon? Pretty easy! Bye for now.

Example 12:
Write an equation representing the standard enthalpy of formation of liquid N2O3
Solution:
This sounds hard at rst, but it’s actually kinda simple. We know that our product is N2O3. So we already
have half the equation
____ + ____ ➞ N2O3
Now we just need to put in the reactants. The reactants we use are the elements that make up N2O3:
nitrogen and oxygen. So, then, do we do this
2N + 3O ➞ N2O3
No! We must use the elements how they exist in their normal, everyday state. Both nitrogen and oxygen are
BrINClHOFs: they are diatomic gases. So we’d writ
N2(g) + 3/2O2(g) ➞ N2O3(l)
Notice how I included the physical states. Also notice how I left a fractional coef cient for O2. We normally
don’t do this for chemical equations; but we do it for standard enthalpy of formations

10.5.4 The easy way of nding the value of ∆Hrxn


Let’s say you were going to use Hess’s Law to solve for the ∆Hrxn of
2N2O3(g) + 3CF4(g) ➞ 4NF3(g) + 3CO2(g)
given the following information
N2(g) + 3/2O2(g) ➞ N2O3(g) ∆Hfº(N2O3
C(gr) + 2F2(g) ➞ CF4(g) ∆Hfº(CF4
½N2(g) + 3/2F2(g) ➞ NF3(g) ∆Hfº(NF3
C(gr) + O2(g) ➞ CO2(g) ∆Hfº(CO2
You would solve by adding the equations as follows (check to make sure you understand what I’m
doing). To get the 4 NF3’s in the products, I’d have to multiply the coef cients of the third equation
by four. That means I’d also need to multiply its ∆Hfº by four (see below). Then I’d need to multiply

372 Chapter 10 Thermodynamics


fi

fi
:

fi
e

fi
fi

fi
e

fi
)

fi
.

fi
d

the fourth equation (and its ∆Hfº) by three to get the 3 CO2’s. To get 2 N2O3’s on the reactant side, NOTE: Don’t worry about all this. It’s
I’d need to multiply the rst equation by two and reverse it. As a result, I’d need to multiply its ∆Hfº going to all come together to show
by two and reverse the sign. And nally to get the 3 CF4’s, I have to multiply the second equation by you something convenient and easy!
three and reverse it (and likewise its ∆Hfº)
2N2(g) + 6F2(g) ➞ 4NF3(g) +4∆Hfº(NF3
3C(gr) + 3O2(g) ➞ 3CO2(g) +3∆Hfº(CO2
2N2O3(g) ➞ 2N2(g) + 3O2(g) –2∆Hfº(N2O3
3CF4(g) ➞ 3C(gr) + 6F2(g) –3∆Hfº(CF4)
2N2O3(g) + 3CF4(g) ➞ 4NF3(g) + 3CO2(g)
∆Hrxn = +4∆Hfº(NF3) + 3∆Hfº(CO2) – 2∆Hfº(N2O3) – 3∆Hfº(CF4)
Look at that value for the heat of reaction that results on the bottom line. Now look at the balanced
equation right above it. Notice a pattern or relationship? It ends up that all I’m doing is adding the
∆Hfº of each product, each multiplied by its coef cient, and then subtracting the ∆Hfº of each
reactant, each multiplied by its coef cient. In other words, for a general reactio
aA + bB ➞ cC + d
(where the lowercase letters are the coef cients from the balanced equation), the value of ∆Hrxn i
∆Hrxn = c∆Hfº(C) + d∆Hfº(D) – a∆Hfº(A) – b∆Hfº(B
o
∆Hrxn = [c∆Hfº(C) + d∆Hfº(D)] – [a∆Hfº(A) + b∆Hfº(B)
All you need to do to nd the ∆Hrxn is to look in a table for the heats of formation of each reactant
and product, multiply each by its coef cient, and then subtract the sum of the heats of formation of
the reactants from the sum of the heats of formation of the products. We often simply say “products
minus reactants,” but don’t forget about multiplying by the coef cients

If you think about it, it makes conceptual sense. Remember that the heat of formation of a
compound is the energy released (or absorbed) when you synthesize it from the elements that
make it up. The heat of formation of CO2 is –393.5 kJ/mol. This says that 393.5 kJ are released
when you make one mole of CO2 from C and O2. Then how much energy would need to be
absorbed if you did the reverse: split CO2 into C and O2? +393.5 kJ. It’s just the reverse reaction of
the standard enthalpy of formation, so it will have the opposite sign (positive instead of negative).
That is, when we make a compound like CO2, the energy involved is ∆Hfº. But when we break the
compound apart, the energy involved is –∆Hfº.

Now look back up at the sample balanced equation. How can we get from the reactants (N2O3 and
CF4) to the products (NF3 and CO2)? Each side has the same elements: they’re just bonded
together differently. One way we can get from reactants to products is to decompose all the
reactant compounds into their elements (break N2O3 and CF4 into N2, O2, C(gr), and F2). Then we’d
take these four elements and combine them to form the product molecules (NF3 and CO2). The
energy to do the rst part (breaking apart the reactant molecules) involves each reactant’s –∆Hfº
value. The energy to do the second part (forming product molecules from the elements) involves
each product’s ∆Hfº value. However, we’d need to multiply each by their coef cient, because we’re
decomposing or synthesizing more than one mole. What we’re saying is this: “It takes –2∆Hfº(N2O3)
and –3∆Hfº(CF4) to break apart the reactants and 4∆Hfº(NF3) and 3∆Hfº(CO2) to form the products.
The heat of reaction will be the sum of these four energies.

It is highly unlikely that the reaction actually occurs this way (all the reactants breaking into
elements, then recombining to form products). But who cares? The fact that enthalpy is a state
function means that the energy we calculate from this method will be the same as what we’d
calculate if we used the “real” method. So it’s all good.

A conceptual understanding is cool and all, but it will just come down to you using the “products
minus reactants” rule. For example, for the reactio
2 NO2 ➞ N2O4
we know right away that ∆Hºrxn = ∆Hfº(N2O4) – 2∆Hfº(NO2) (products – reactants, each multiplied Since we are calculating the ∆Hrxn by
by its coef cient). Look up the ∆Hfº values in a table, plug ‘em in, and presto, you got your answer. using the ∆Hfº values from the table
for elements in their standard states
[But conceptually, what we’re doing is this: “We’re breaking apart the two moles of NO2 into its
at 298K, we should use the º symbol
elements, which involves –2∆Hfº(NO2), and then forming one mole of N2O4 from these elements, and write ∆Hºrxn. This is important if
which involves ∆Hfº(N2O4). So the heat of reaction is the sum of these two energies.” you major in chemistry, but not so
much for our intro level course.

Chapter 10 Thermodynamics 373


r

fi
fi
D

fi
fi
fi
fi
fi
fi
.

fi
]

fi
.

fi
n

Example 13:
Determine the ∆Hºrxn for the reaction 2C2H6(g) + 7O2(g) ➞ 4CO2(g) + 6H2O(l)
Solution:
Look up the values of ∆Hfº in your table. For CO2(g), H2O(l), and C2H6(g) (recall that ∆Hfº = 0 for O2(g)),
those values are, respectively, –393.5 kJ/mol, –285.5 kJ/mol, and –84.7 kJ/mol. Just do “products minus
reactants,” using the appropriate coef cients
∆Hºrxn = [4(–393.5 kJ) + 6(–285.8 kJ)] – [2(–84.7 kJ)] = –3119.4 kJ
Be careful with those negative signs

Section Overview

The Forest:
By de ning a standard state of conditions, results can be tabulated and published for the enthalpy
changes due to the formation of compounds from their elements in their most stable states. We can
then use these known ∆Hfº values quickly and easily to nd ∆Hrxn of other reactions

The Trees:
The standard enthalpy of formation is the enthalpy change in producing one mole of a compound
from its elements in their most stable states at the standard conditions of 1 atm and 25ºC. Books
list values for the ∆Hfº of many compounds. Because of the way that Hess’s Law works out, we can
nd the ∆Hºrxn for any reaction simply by subtracting the ∆Hfº of the reactants, each multiplied by its
coef cient in the balanced equation, from the ∆Hfº of the products, each multiplied by its coef cient.
In other words, “products minus reactants.

Test Yourself:
1. Write an equation representing the standard enthalpy of formation of liquid ethanol C2H5OH.

2. Use your table of ∆Hfº‘s to nd the ∆Hºrxn of the following reaction: 6CoO(s) + O2(g) ➞ 2Co3O4(s

10.6 Finding ∆H by Bond Energies


So far, we’ve talked only qualitatively about the strength of bonds: which kinds are stronger than
which. For example, in Chapter 1, we said that a C=C double bond is stronger (harder to break)
than a C–C single bond. However, there is a quantitative measure of how strong a bond between
two atoms is. The bond energy is de ned as the amount of energy required to break a bond of one
Reminder: A covalent bond is due to mole of molecules. For example, the bond energy (BE) of the H–H covalent bond is 432 kJ/mol,
the sharing of electrons between two meaning you have to put in 432,000 J of energy to break all the H–H bonds of a mole (2.0 grams)
nonmetal atoms. of H2 gas.
H––H(g) ➞ H(g) + H(g) ΔHrxn = BE = +432kJ/mol
What if you want to form one mole of H2 from individual hydrogen atoms, as shown below

H(g) + H(g) ➞ H––H(g) ΔHrxn = –BE = –432kJ/mol

How much energy would this involve? Notice that this is the reverse of the rst equation. Therefore,
the ΔH of this reaction will be –432kJ/mol, the opposite of the bond energy. This means that when
the two separate hydrogen atoms come together to form a covalent bond (making a molecule of
H2), 432kJ of energy are released. While bond breaking is endothermic, bond formation is
exothermic. Keep that in mind for this section: breaking bonds absorbs energy (endo) while forming
bonds releases energy (exo). Below are some sample bond energies. Notice that they are all
positive: they represent how much energy it takes to break that bond. This is endothermic, so it is
positive. (Notice some values are for double or triple bonds.

374 Chapter 10 Thermodynamics


fi
fi
fi

fi
!

fi
fi
:

fi

)

fi

fi
Bond Energy (kJ/mol) Bond Energy (kJ/mol)
H–H 432 N–N 160
H–F 565 N≡N 945
H–Cl 427 N–H 391
C–H 413 O–H 467
C–F 453 F–F 159
C–Cl 339 Cl–Cl 243
C–C 347 C=C 614
C–O 358 C≡C 839
C=O 745 O=O 498
There is some usefulness to knowing the bond energies. Recall that the heat (enthalpy) of reaction
is the amount of energy absorbed or released in a chemical reaction. For example, the ΔHrxn in
forming HCl is given b

H2(g) + Cl2(g) ➞ 2HCl(g) ΔHrxn = –179kJ

When 2 moles of HCl are formed from 1 mole of hydrogen and 1 mole of chlorine, 179 kJ of heat
are produced. If you didn’t know the value of ΔHrxn, you can use the bond energies to determine it.

Let’s use an analogy. Say you have 2 blue Legos stuck together and 2 red Legos stuck together.
You want to make two pairs of Legos, each one with a red one stuck to a blue one. What do you
do? You pull the blue-blue pair apart and the red-red pair apart to get four individual Legos. Then
you attach each blue one to a red one. Now you have your two red-blue Lego pairs.

Now let’s switch from Legos to atoms. This description of “Lego building” is similar to building HCl
from H2 and Cl2. To do the reaction, we need to break apart the H–H molecule and the Cl–Cl
molecule to form four individual atoms. Then we take these now-separated atoms and attach each
H to a Cl. BOOM! We got our pair of HCl molecules. If we want to make two moles of HCl (as the
balanced equation says), we need to break apart one mole of H2 and one mole of Cl2, then form
two moles of HCl. Let’s follow the energy requirements, remembering that breaking bonds is
endothermic and forming bonds is exothermic

1. How much energy does it take to break apart one mole of H2 molecules? According to our table,
we must put in +432kJ (positive because this requires energy)
2. Now we put in another +243kJ to break apart the mole of Cl2 molecules (see table)
3. Lastly, we need to make two moles of HCl, which means we are forming two moles of H–Cl
bonds. According to our table, 2(–427kJ) = –854kJ of energy are released when this happens
(negative because energy is produced)
4. Altogether, it took an input of +675kJ to break apart the reactants, but we got –854kJ released Warning! Don’t get this method
when the products formed. Therefore, there is a net –179kJ released. That’s the ΔHrxn we originally confused with the ∆Hfº method of
last section. We don’t do “products
wrote minus reactants” here.

That’s the idea. You use bond energies to determine how much energy you use to break apart all
the reactant molecules completely into atoms (endothermic). Then you use the bond energies to
nd out how much energy is released when you put them together to form the products
(exothermic). Visually, this is what we did

Break apart H2: +432k Form 2 HCl molecules


Break apart Cl2: +243k 2(–427kJ) = –854kJ
Total: +675kJ

Total energy: +675kJ – 854kJ = –179kJ

Is this what really happens in the reaction? Probably not. But remember that enthalpy is a state
function: it doesn’t matter how we get from A to B (H2/Cl2 to 2HCl), the change in enthalpy will still
be the same

Chapter 10 Thermodynamics 375


fi
.

I’ll do one more for you. Let’s nd the heat of reaction for 3 H2(g) + N2(g) ➞ 2 NH3(g)
Before we start, I should tell you that the two N atoms of N2 are connected with a triple bond (you
discovered this in Chapter 5) and that NH3 is made up of a nitrogen atom bonded to three hydrogen
atoms, so NH3 has three N–H single bonds (you probably knew that). You’ll need to know enough
about bonding (which of them are single, double, or triple) to solve these types of questions.

We need to break apart 3 moles of H–H bonds [3(432kJ)=1296kJ] and one mole of N≡N triple
bonds [945kJ...use the correct value for the N≡N triple bond]. Altogether, then, we need to put in
+2241kJ to rip everything apart. We need to form a total of six N–H bonds (notice we’re making two
moles of NH3). Our table tells us that this will release 6(–391kJ) = –2346kJ. Now add them:
+2241kJ – 2346kJ = –105kJ. That’s the ΔHrxn

Example 14:
Find that ∆Hrxn for the following C2F4 + 2 H2 ➞ C2H4 + 2 F2
The reaction is shown here schematically
F F H H

C C + 2 H H ➞ C C + 2 F F

F F H H
Solution:
Let’s rst break apart all the bonds of the reactants. I’ll use the table to nd out how much energy it takes. I
have one C=C double bond (614kJ), four C–F bonds (453kJ•4), and two H–H bonds (432kJ•2). Altogether, it
takes +3290kJ to do this. Now let’s see how much energy is released when forming the products. I form one
C=C double bond (–614kJ), four C–H bonds (–413kJ•4), and two F–F bonds (–159kJ•2). Altogether, –2584kJ
are released in making the products. Overall, then, the ∆Hrxn is (3290kJ –2584kJ) = +706kJ.
Did you notice something? I added 614kJ to break apart the C=C double bond in the reactants, but
subtracted that same 614kJ when I formed the C=C double bond in the products. Since that double bond
never broke, I really didn’t need to do anything to it. That makes these problems even easier. If some bond
isn’t broken in going from the reactants to the products, just ignore it

Section Overview

The Forest:
It takes energy to break covalent bonds. Energy is released when they are formed. By analyzing
what bonds are broken and formed in a chemical reaction, you can determine the heat of reaction

The Trees:
The bond energy is the energy required to break a covalent bond. It is expressed as a positive
value since it is an endothermic process. Conversely, forming bonds is an exothermic process:
energy is released when atoms come together to form a bond. To determine the enthalpy of
reaction for a given equation, you simply add up the bond energies of all the bonds in the reactant
molecules then subtract from this the bond energies of all the bonds in the product molecules.

Test Yourself:
We want to nd the ∆Hrxn for the combustion reaction CH4 + 2 O2 ➞ CO2 + 2 H2O
Shown below is the reaction, where single lines represent single bonds and double lines are double bonds. I’ll help
guide you for this one
H

H C H + 2 O O ➞ O C O + 2 H O H

H
a. How many C–H single bonds need to be broken
b. How much energy will that take? (This should be positive.
c. How many O=O double bonds need to be broken
d. How much energy will that take? (This should be positive.
e. How many C=O double bonds need to be formed
f. How much energy will that release? (This should be negative.
g. How many O–H bonds need to be formed
h. How much energy will that release? (This should be negative.
i. Now add it all up and that’s the ∆Hrxn

376 Chapter 10 Thermodynamics


fi
fi

fi
:
.

fi

10.7 Spontaneity and Entropy


10.7.1 Spontaneous processes happen in a particular direction
Consider the following everyday observations
When you release a ball from your hand, it falls to the groun
When you ignite a fuel, it burns until consume
A gas spreads across a roo
Wet clothes dry when hung from a clotheslin
When you put solid NH4Cl in water, it dissolve
Ice melts at room temperatur
Water freezes below 0º
A drop of food coloring spreads throughout a cup of wate
These things just happen. We don’t force them – they are quite happy to do these things on their
own. They are spontaneous. By spontaneous, I do not mean that they happen quickly. I simply Some spontaneous processes are
mean that they start to happen in this way, not in the opposite way. We don’t see a pool of water agonizingly slow, like iron rusting.
turning into ice at room temperature. The food coloring of a uniformly blue liquid doesn’t all gather
into one little concentrated drop. We don’t see a ball on the ground suddenly jump up on its own to
your hand. Those are all non-spontaneous processes. Spontaneity refers to the direction of
physical or chemical processes. But why do some things happen in one way and not another? Let’s
nd out…

10.7.2 Processes in which energy is lowered tend to be spontaneous


Some spontaneous processes are not dif cult to explain. You may have heard how nature likes to
run downhill: everything wants to lower its potential energy (PE)...everything wants to be in a
stable, low-energy state. Does an object “know” where it needs to go to lower its PE? No. But
everything has somewhere it “prefers” to be – whether it prefers to be there because of the effect of
gravity or electrical attraction or whatever. This preferred state of the object is de ned to be a state
of low potential energy. Saying that something is going to a state of low potential energy is
equivalent to saying that it is going to where it likes to go.

Take the rst example in the list above. When you hold a ball in your hand, it has GPE. Gravity
wants to pull it to the ground. In other words, the ground is a position of low potential energy
because that is the preferred state of the ball. Your hand is preventing it from falling to the ground.
When you remove your hand, the ball is now free to return to this state of low potential energy.
Conversely, a ball that is already resting on the ground (low GPE) is highly unlikely to suddenly
jump up on its own to your raised hand, increasing its potential energy. We say that the process of
a ball falling is always a spontaneous process. The reverse process – a ball at rest on the ground
jumping into the air – is a non-spontaneous process. In general, a process involving a decrease in
energy tends to be spontaneous, and the reverse process tends to be non-spontaneous, meaning
we see things happen the former way and not the latter

10.7.3 Exothermic processes tend to be spontaneous



Let’s now see how this relates to chemical reactions, such as the second one on the list above.
Once lit, a beaker of ethanol continues to burn on its own, producing CO2 and H2O. That is, the
burning of ethanol is spontaneous. We never see the opposite: a bunch of CO2 and H2O gases
coming together and reacting to form liquid ethanol that falls into the beaker. If you think about the
enthalpy change of the reaction, you’ll see why burning ethanol is spontaneous.

ethanol + O2 In the diagram, notice how enthalpy decreases in going from the
reactants to the products: the system is lowering its internal energy.
This exempli es the idea “nature tends to run downhill.” As stated
enthalpy

above, we would expect this to be a spontaneous reaction.

We can say that spontaneous processes tend to be the ones in


CO2 + H2O which the reacting system lowers its enthalpy. What is it called when
a reaction has a decrease in enthalpy, releasing that extra energy to
the surroundings? It is called an exothermic process. We’ll
generalize and say that exothermic processes tend to be spontaneous. This explains the second
item in the list above. Combustion – an exothermic reaction – is spontaneous.

Chapter 10 Thermodynamics 377


fi
.

fi
fi
C

fi

fi
10.7.4 Okay, but not ALL spontaneous processes are exothermic
Notice that the last paragraph says that exothermic processes tend to be spontaneous. Just
because something is exothermic doesn’t guarantee it will be spontaneous. The same can be said
of the converse: just because something is spontaneous doesn’t mean that it is exothermic

Let’s examine some exceptions. Take the third example in the list that opens this section. Consider
a ask of gas. You open the ask and after a while, the gas has spread throughout the room.
Everyone can smell it. This always happens: a sample of gas will always want to diffuse evenly
throughout a room. We would say that this is a spontaneous process. We never see the reverse
process: a spread-out gas all gathering into one little corner of the room. Obviously, nature has
de ned a preferred direction: gases spread out – not concentrate – on their own. However, there is
no change in enthalpy! A con ned gas turning into a spread-out gas has a ∆H = 0. What is making
one process spontaneous and the opposite process non-spontaneous?

Another example is even stranger. Hang up a wet shirt and it will dry (the fourth example from the
Remember that spontaneous doesn’t list above). This must mean that liquid water is spontaneously turning into water vapor, right? But
mean “immediately,” only that it this is an endothermic process! It requires the absorption of energy to vaporize water
happens in that direction.
H2O(l) ➞ H2O(g) ∆H = +40.8 kJ/mo

It’s true that nature likes to go downhill – a system wants to lower its energy. Spontaneous
processes tend to be exothermic, and non-spontaneous processes tend to be endothermic. Here is
direct proof that it is not always true, and why it is that we should include “tend to” in our statement

Let’s take another example. Have you seen chemical cold packs? They get cold because the
solution process is endothermic for some solutes, such as ammonium chloride: the system absorbs
heat from the water (the fth example on our list)

NH4Cl(s) ➞ NH4+(aq) + Cl–(aq) ∆H = +15.2 kJ/mo

But we know this is a spontaneous process (the solid dissolves readily). Okay, so there must be
something else at work here. Determining if a process is endothermic or exothermic (positive ∆H or
negative ∆H) is not enough to determine if it will be spontaneous. And knowing that something is
spontaneous doesn’t necessarily mean it was exothermic. There must be another determining
factor that we must take into account

10.7.5 Pause for re ection: Is there a deeper reason for spontaneity?


We started this section by asking why some things are spontaneous. Why do some things naturally
happen one way without external in uence (such as a ball falling towards the ground) whereas it
requires an input of energy to occur in the other direction (such as lifting the ball up to a ledge)?
Why don’t objects ever spontaneously jump up to higher places on their own?

We thought we found an answer. We said that nature likes to run downhill: things want to lower
their energy. For example, we “explained” why a ball falling towards the ground is spontaneous by
saying it is going to a state of lower potential energy. The problem is that we de ne low potential
energy as the place where the ball will spontaneously want to go. This is circuitous reasoning. In
our attempt to explain why some things are spontaneous, we are saying things go in the direction
of lower potential energy: in other words, they go to a place that they want to spontaneously go. We
are simply begging the question.

You might say, “Of course a ball only There must be another deeper factor that causes one direction to be spontaneous. This became
falls DOWN. Gravity forces it that clearer in the last subsection: gases always spread out, although that involves nothing “going
way.” Even so, we need to explain downhill.” And ice cubes melt at room temperature, despite requiring to “go uphill” (endothermic).
why other things – such as gases
spreading out or ice melting – are We will discuss this new factor momentarily. However, please do keep in mind that, even if it isn’t a
spontaneous when there are no completely de nitive answer, spontaneous directions more often than not correspond to processes
forces “forcing” them to go that way. in which energy decreases. That is, spontaneous reactions really do tend to be exothermic ones.
So if you see a ∆Hrxn < 0, you can probably make a fairly good prediction that it will be
spontaneous. However, to be certain that it is spontaneous, we need to know about the second
factor that determines whether it will be spontaneous or not: entropy

378 Chapter 10 Thermodynamics


fl
fi
fi
fl
fi
fi
fl

fl
.

fi
:

10.7.6 Entropy is the dispersal of energy


That other determining factor is entropy, which is a measure of the dispersal of energy at a certain
temperature. To illustrate what this means, consider these everyday spontaneous examples
๏ a bulb gives off ligh
๏ a hot pan cools of
๏ a ball falls from your hand to the groun
๏ a wave spreads out when you drop a rock into a lak
In each of these examples, energy is going from being localized and concentrated to being spread
out. When you turn on a bulb, light radiates away from the hot lament. The energy concentrated
as thermal energy in the tiny lament disperses throughout the room, becoming less concentrated.
Next, consider the hot pan. Initially, the thermal energy is localized in the pan. It cools off by
transferring some heat to the cooler surroundings. By doing so, the energy becomes less
concentrated. Instead of being all in one place (the pan), the energy becomes dispersed throughout
the room. This “dispersal of energy” is what we call entropy.

Next consider the ball. As the ball falls, it converts its GPE to KE. The ball collides with the air
molecules, speeding them up. When it hits the ground, it causes the molecules in the ground (and
in the ball itself) to speed up. The energy that was localized in the ball as GPE is now spread out to
the random kinetic energy of the particles of the ball, the ground, and the air (that is, it was
converted into thermal energy/heat). Again, energy has been dispersed. The fourth example is
similar. When you drop a rock into a lake, the GPE concentrated in the rock is dispersed as
motional energy in the water waves (and sound waves) that spread away from the point of impact.

10.7.7 Spontaneous processes tend to be those for which entropy increases


Don’t the examples from the subsection above seem like common sense from experience? We
don’t ever see light in a room all gather at a cool lament to warm it up. We don’t see cooler air
transfer heat to a warmer pan, making the air even cooler and the pan even hotter. We don’t see
heat from the ground and air suddenly concentrate as KE into a tennis ball on the ground, causing
it to jump up. Things that we see happening in everyday life in one preferred direction and not In fact, our psychological perception
another are because of entropy, the dispersal of energy. of the passage of time could be due
to entropy. If we see a movie of a
ball jumping up into someone’s
From this practical experience approach, we will conclude that spontaneous processes tend to be hand, we’d say that the movie is
those that involve a dispersal of energy. Recall that entropy is a measure of this dispersal of playing in reverse, wouldn’t we?
energy. When energy is dispersed, we say that entropy has increased. Therefore, we can write a
similar statement as the italicized one above: spontaneous processes tend to be those that
involve an increase in entropy. Now we have two things to think about when considering
spontaneity: enthalpy and entropy. Spontaneous reactions tend to have a decrease in enthalpy and
an increase in entropy

Let’s symbolize entropy with the letter S. The change of entropy of some process is symbolized by
∆S. Because “change” is always “ nal minus initial,” a positive ∆S signi es an increase in entropy.
An increase in entropy is “good” for spontaneity. That is, a positive ∆S favors spontaneity. By that, I
mean that, as far as entropy is concerned, it would want the reaction to occur in this direction You should memorize this right now!
(although we would need to also analyze the effect of enthalpy). A negative ∆S means that there ∆S>0: entropy increases, favoring
would be a decrease in entropy. This is “bad” for spontaneity. Thus, a negative ∆S favors non- spontaneity
spontaneity. This means that entropy would not want the reaction to occur in this direction ∆S<0: entropy decreases, favoring
non-spontaneity
As stated above, entropy (or more correctly, the change in entropy) is a measure of the dispersal of
energy at a given temperature. It really does measure an energy dispersed (divided by the Specifically, we mean the dispersal
temperature at which it is dispersed). Do not worry about this equation, but the change in entropy of of motional energy (KE) of the
a system is equal to the heat transferred (dispersed) to the system divided by the temperature in particles.
kelvin:
∆S = q/
Entropy’s units are joules per kelvin (J/K). If 600J of energy are dispersed into a substance at
200K, the entropy increase of the substance is 3 J/K. That’s how much energy (per kelvin) was
dispersed into the substance. Again, don’t worry about this. I just wanted to show you that entropy Technically, the energy dispersed to
changes can be measured. You just have to nd a way to measure how much heat was transferred the system must occur reversibly.
That is, it must occur slowly enough
and divide by the temperature at which it was transferred. that we can stop or reverse the heat
transfer at any moment.

Chapter 10 Thermodynamics 379


T

fi
fi
d

fi
e

fi

fi
fi

10.7.8 Entropy-increasing physical processes


There’s another way to describe In this section, we’ll learn to understand some processes that involve an increase in entropy: that
entropy using statistics. Sometimes it is, those that tend to be spontaneous. After the explanations, we’ll come up with a concise, clear
is easier to explain things using summary. Once that is in place, we’ll have an easy time applying entropy to actual chemical
statistical thermodynamics and
sometimes it is easier to use
reactions in order to help decide if they are spontaneous (but we’ll need to consider enthalpy, too)
classical thermodynamics (“dispersal
of energy”). To avoid confusion, I’ll
[PLEASE NOTE: The next few pages are going to be heavy in exposition. Hang in there. By the
stick with classical thermo here. In a end of §10.7.10, I’ll simplify all this discussion into a few facts you need to memorize. The actual
future Interchapter, I’ll present a questions we’ll ask you are very straightforward and don’t require you to explain those few
quick description of stat thermo. memorized facts. But I want to give you the theory behind entropy so you can appreciate it more.

There are two ways in which energy is thought to be dispersed. One is when an actual quantity of
motional energy is dispersed: that is, there is some heat transferred, and we could quantify the
entropy change using q/T. Since motional energy was dispersed into the system, its entropy has
increased. The second way in which energy is dispersed is when the energy of the system itself
spreads out in 3D space. The material possesses motional energy, right? If the material spreads
out spatially, its energy is spreading out spatially, too. That is a dispersal of the system’s energy
and thus the system’s entropy has increased.

Practically, then, the dispersal of energy manifests itself in two principal ways
1. When motional energy (heat) is transferred from a hotter object to a colder objec
2. When motional energy (heat) spreads out spatially instead of being con ned
I want to emphasize how logical this is. Take #1. Isn’t it common sense that heat ows from hotter
That said, why does heat always to colder objects? We don’t see a cold table transferring heat to a hot pan in order to make the pan
travel from the hotter to colder hotter and the table colder, do we? No. We only see hotter things transfer heat to colder things.
substance? Read the following
Interchapter and find out.
Now take #2 above. This should also be understood as something totally natural and of common
sense. Consider a gas trapped in a jar with a lid on it. The particles are ying around, slamming into
each other and the walls/lid, wanting to move away from each other and spread out. It is only the lid
that is preventing that. Given the opportunity (opening the jar), it is only natural that – because of
the energy of their motion – the particles will tend to move apart and increase in volume. So while
“entropy” and “dispersal of energy” sound arcane, it is actually easy for anyone to understand

Each of the following processes involves an increase in entropy for the system of interest, which
means that ∆S is positive for all these processes. This also means that entropy favors these
processes being spontaneous (although enthalpy will have a say in the matter!). Keep in mind the
two reasons listed above why entropy of the system would increase

The volume of a gas increases


When a gas has a small volume, its particles are con ned to a small space. All the kinetic energy of
those particles is localized. By expanding to a larger volume, the particles – and their energy –
becomes spread out. The energy of the system has become dispersed spatially. This indicates an
increase in entropy: ∆S > 0

It should suffice to say entropy has


increased because the system’s Volume increases
energy has spread out spatially. But Energy of gas dispersed
this actually does involve a transfer
of heat to the system. This will be Entropy increases
shown in the following Interchapter.

Mixing of two materials


Imagine two gases in a container with a barrier between them. If the barrier is removed, each gas
will naturally spread out to occupy the full space. That is, the two gases will become mixed with
each other. The same thing can be said when two liquids mix. Each has its energy dispersed, so
entropy has increased

Each gas’s energy is


dispersed

Entropy increases

380 Chapter 10 Thermodynamics


fi
.

fl
:

fl
fi
t

A substance is heated (i.e. its temperature increases)


If something is being heated, it must be in contact with something hotter. Of course, heat will be
transferred from the hotter surroundings to the cooler system (Reason #1 above). By increasing the
motional energy of the particles of the object, we’ve dispersed energy into it and we’ve increased
its entropy

A solid melts into a liquid


Think of a piece of ice in a warm room. Of course, heat will ow from the room to the ice, causing it
to melt. Since the warm room disperses energy into the ice cube, that means the ice cube’s entropy
has increased. This favors spontaneity. (We say “favoring” because we need to consider enthalpy
and temperature. Sometimes a solid melting isn’t spontaneous, such as at low temperatures, even
if the entropy would increase. We’ll come back to this later.

A liquid vaporizes into a gas


This has the same explanation that we used for a solid melting. If you place a drop of cold water on
a hot plate, thermal energy will be dispersed into the liquid, causing it to vaporize. Again, this
means the entropy of the H2O has increased. Additionally, by turning into a gas – which has a
larger volume – the energy of the particles spreads out spatially. So entropy increases for both
reasons

Summarizing brie y, whenever a system increases in volume, mixes with another substance,
increases in temperature, melts, or vaporizes, the entropy of the system increases, which means
that entropy favors these processes being spontaneous. Two sections from now, we’ll apply these
ideas to chemical reactions, which will be our primary focus. Before that, let’s take a closer look at
an important idea that will be very useful later

10.7.9 Gases are the “kings” of entropy


Those last three changes (heating a substance, melting it, vaporizing it) all involve heat energy
moving from the warmer surroundings to the cooler system: an increase in entropy as de ned by
Reason #1 above. But there is a difference. When you heat a substance, the heat that is
transferred to that substance goes towards increasing the kinetic energy of its particles. That is, the Recall that temperature is a measure
temperature of the substance increases. Melting and vaporizing are different. of the average kinetic energy.

Consider melting. Take 1 gram of ice at 0ºC. When it melts, what is the initial temperature of that
liquid water that forms? 0ºC. The ice cube had to absorb energy from the surroundings just to melt
without even changing temperature yet. Why? Recall the difference between a solid and liquid:
solid particles vibrate rigidly in place while liquid particles are free to wander around. We need to
add energy to a solid to loosen some of these attractive forces among the solid particles in order to
allow them to wander away from each other as a liquid. For water, it takes 80 calories to melt 1 g of
ice at 0ºC into 1 g of liquid water at 0ºC. (Compare that to the measly 1 cal needed to raise the
temperature of 1 g of liquid water by 1ºC.) That means when 80 cal of energy are dispersed into
the 0ºC ice to form 0ºC liquid water, it did not turn into motional energy (KE/temperature) in the
water. Then were did those 80 cal go? It went into a type of potential energy associated with phase
change (change of state), called latent heat (aka phase change energy). Putting all the explanation The energy used to melt a solid ends
aside, the point I want to make is this: just to create a liquid water out of solid water (melt it), it up as latent heat in the liquid. The
energy used to vaporize a liquid
requires 80 cal/g. Energy needs to be dispersed into a material to melt it. In this sense, the ends up as latent heat in the gas.
resulting liquid has more entropy than the solid because it has that extra energy dispersed into it,
even if they are at the same temperature.

The same thing happens when a liquid vaporizes into a gas. It takes 540 cal to turn 1 g of liquid
water at 100ºC into 1 g of gaseous water at 100ºC. So the gas has all that extra energy dispersed Again, that added energy didn’t end
into it compared to the liquid. That means the gas has more entropy than the liquid (which has up as KE but as latent heat.
more entropy than the solid). But also notice how much energy was required: 6.75 times more than
it took to melt the ice. Think again of our model of liquids and gases. Gas particles are completely
free. In order to form a gas, we have to completely break every attractive force among the liquid
particles. That takes a lot of energy! So all that energy that is dispersed into the liquid to create the
gas gives gases lots of entropy

In summary (and this is crucial): for any given chemical, the liquid state has a higher entropy
than the solid state, and the gaseous state has a much higher entropy than the liquid state.
Always think of this fact: gases are the kings of entropy

Chapter 10 Thermodynamics 381


.

fl

fl

fi
10.7.10 Entropy in chemical reactions
Instead of talking about hot pans and dropped balls, let’s think about entropy as it relates to
chemistry. We would like to be able to determine if a system experiences an increase or decrease
of entropy as a result of a chemical reaction. That will help us decide – when combined with
knowledge about whether the reaction is endothermic or exothermic – if it will be spontaneous

Earlier, we learned that whenever a system increases in volume, mixes with another substance,
increases in temperature, melts, or vaporizes, its entropy increases. Let’s take a look at the main
examples of entropy in chemical reactions. I’ll keep my exposition brief because I already explained
the ideas above. All of these involve an increase of entropy (∆S > 0), which means entropy favors
that they are spontaneous.

A reaction in which a solid or liquid turns into a gas


Consider this reaction, in which solid calcium carbonate decomposes into solid calcium oxide and
gaseous carbon dioxide
CaCO3(s) ➞ CaO(s) + CO2(g)
What about if a solid turns into a Although this is not melting (the solid CaCO3 didn’t turn into liquid CaCO3), the major thing here is
liquid? I’d assume that the entropy we went from having only a solid to having a gas. Gases are the kings of entropy, baby! Whenever
increases, but solids and liquids
you see only solids or liquids in the reactants, but there is a gas in the products, you can rest
aren’t that different in entropy. It
depends on the reaction. But when a assured that entropy increases.
solid or liquid turns into a gas, that
pretty much guarantees ∆S > 0. A reaction in which the number of gas particles increases
Let’s say you have a two-mole sample of carbon dioxide gas. It decomposes into two moles of
carbon monoxide and one mole of oxygen
2 CO2(g) ➞ 2 CO(g) + O2(g)
The easiest way to see why this involves an increase of entropy is by noticing that we went from
having two moles of gas to having three moles of gas. More gas particles? What do we know about
gases? Say it with me: gases are the kings of entropy, baby! More gas particles = more overall
entropy. Thus, ∆S > 0 and this favors spontaneity. (For reactions that involve only solids and liquids,
the difference in entropy is not so pronounced, so we can safely say that entropy increases only
when the number of gaseous molecules increases.
[If you want other ways to explain why this is an increase of entropy, here you go: First, think of the motional
energy being con ned to just 2 moles of gas particles. By transforming into 3 moles of gas particles, the
motional energy is more spread out/dispersed in a sense that more particles are sharing the thermal energy.
Second, if this occurs at constant pressure (as we usually assume), there will be an increase of volume. As you
know, that’s an increase of entropy. Finally, we went from having a pure substance (only CO2) to having a
mixture of two gases. That’s an increase of entropy, too.

A solid or liquid solute dissolves in a solvent


Now let’s take a solid (say NaCl) and dissolve it in a solvent (like H2O). The solid chunk is broken
up into Na+ and Cl– ions that are now free to “swim” all around the solution
NaCl(s) ➞ Na+(aq) + Cl–(aq)
The NaCl has now “spread out” over the whole solution. Hey! What was once two separate entities
(the liquid H2O and solid NaCl) has now become thoroughly mixed with each other. As we learned,
There are some special instances when things mix, the entropy increases. Again, ∆S > 0 and this favors spontaneity. We also already
when dissolving a solid in a solvent learned that dissolving a liquid in another liquid (“mixing two liquids”) means an increase of entropy.
could result in a decrease of entropy,
but we won’t bother ourselves with Note, however, that dissolving a gas in a liquid solvent always involves a decrease of entropy,
those exceptions. because you are turning the kings of entropy into little dissolved atoms/molecules. Not so kingly

Simple Summary: This is the important conclusion of all the preceding!


Notice how the rst two examples above are basically the same idea: entropy increases when there
are more gas particles in the products than in the reactants. If you memorize/understand that, along
with solids dissolving and pure substances melting or vaporizing, you are all set to analyze
chemical reactions for changes in entropy: our goal for this entire section. In summary

These all involve an increase of entropy (∆S > 0), and thus they favor spontaneity
a reaction in which you end up with more gas particles than you started wit
when a solid dissolves to form a solution (we won’t worry about liquids dissolving
when a pure substance melts or vaporize

382 Chapter 10 Thermodynamics


fi
fi
:

Example 15:
For each of the following, state whether the entropy is increasing or decreasing for the system described
a. A piece of lead melting
b. Two gases react to form a liquid: 2H2(g) + O2(g) ➞ 2H2O(l)
c. Two ions precipitate out of a solution: Ag+(aq) + Cl–(aq) ➞ AgCl(s)
d. Gaseous propane combusting: C3H8(g) + 5O2(g) ➞ 3CO2(g) + 4H2O(g)
Solution:
a. To melt the solid, energy must be dispersed into it. That is the de nition of an entropy increase. Entropy is
increasing
b. We start with the kings of entropy and end up with a liquid? Psssh… entropy is de nitely decreasing.
c. This is the opposite of dissolving, so entropy is decreasing. (You can also think of it this way. The ions
started out thoroughly mixed with the liquid. Their energy is spread out everywhere. When they precipitate
to form a solid, their energy becomes con ned to a smaller space (in the solid precipitate).
d. Notice how we started with 6 moles of gas particles (one C3H8 and ve O2). We end up with 7 moles of
gas particles. The energy gets more widely dispersed when you have more gas particles. All hail the king! The
entropy is increasing

10.7.11 Spontaneity depends on BOTH enthalpy and entropy


We’ve been talking about entropy for so long, we need to now bring enthalpy back into play. We
must use both factors to determine if a reaction will be spontaneous or not (i.e. if it will happen in
the direction that the balanced equation is written). Here’s a reminder of the two factors that must
be considered if you want to determine if a reaction will be spontaneous:
★ Spontaneity is favored when there is a decrease in enthalpy (∆H<0); i.e. it is exothermi “Spontaneity is favored” means the
reaction would like to occur in the
★ Spontaneity is favored when there is an increase in entropy (∆S>0); i.e. energy is disperse direction written by the balanced
equation.
Sometimes enthalpy and entropy work together. For example, if you know a reaction is exothermic
and involves an increase in entropy, it WILL be spontaneous. On the other hand, if a reaction is
endothermic and the entropy is decreasing, it WILL be non-spontaneous. In fact, that would mean
the reaction would prefer to go in the opposite direction

At other times, enthalpy and entropy are opposed. For example, a reaction might be endothermic
(bad for spontaneity), but involve an increase in entropy (good). So, will it be spontaneous? That’s
the topic of the next section. Before that, let’s go back to some of the examples that were listed
earlier in this section, keeping in mind the in uence of both ∆H and ∆S. Remember how we brought
up a couple of endothermic processes that were spontaneous, despite us saying that exothermic
reactions tend to be spontaneous? Let’s explain them now
Wet clothes dry when hung from a clothesline, which we know is spontaneous. Although
water vaporizing is endothermic (∆H>0), the liquid is turning into a gas, which involves a
signi cant increase in entropy (∆S>0). It is the increase in entropy that causes the water to
evaporate. In other words, entropy “outrules” enthalpy here
When you put solid NH4Cl in water, it dissolves. Although the dissolving process is
endothermic, the NH4Cl breaks up, allowing the ions to spread out and mix to form a
solution. Again, the effect of entropy dominates the effect of enthalpy, so it is spontaneous
Okay, there are two driving forces for spontaneous processes: enthalpy and entropy. Which one is
more in uential? How do we predict whether or not a process will happen? We’ll soon see

Example 16:
Each of the following gives you information about either enthalpy or entropy. Furthermore, it tells you if
that property is favoring spontaneity or non-spontaneity. State which property (enthalpy or entropy) it is
describing and whether it favors spontaneity or non-spontaneity
a. A2(s) + B2(s) ➞ A2B2(g)
b. A reaction in which ∆S <
c. A reaction in which the surroundings increase in temperature
d. A reaction in which ∆H = +200 kJ.
Solution:
a. Two solids are reacting to form a gas. That’s giving us information about entropy change. Entropy is
increasing. That means that entropy favors spontaneity.
b. A negative ∆S means that the entropy is decreasing. This means entropy favors non-spontaneity.
c. If the surroundings get warmer, it’s an exothermic reaction. Thus, enthalpy favors spontaneity.
d. A positive ∆H signi es an endothermic reaction. Thus, enthalpy favors non-spontaneity.

Chapter 10 Thermodynamics 383


fi
fl
.

fi
.

fi
fl
.

fi
fi
fi
)

Section Overview

The Forest:
Thermodynamics can tell you whether a particular reaction will happen or not. Take a reaction that
could go either way, such as A + B ➞ C + D or C + D ➞ A + B. If you throw a bunch of A, B, C,
and D in a vessel, which way will they react? What determines the direction of a reaction, what we
call the spontaneous direction, are enthalpy and entropy. Spontaneous processes tend to be
exothermic (releases energy) and involve an increase in entropy (dispersal of energy).

The Trees:
A spontaneous process is one that happens (under current conditions), of its own accord, in a
certain direction and not the other. It describes the directionality of a process, not how rapidly it will
occur. One driving force that in uences spontaneity is how the energy changes: exothermic
processes tend to be spontaneous. Another natural tendency that many spontaneous processes
have in common is that energy gets dispersed. That is, energy goes from being localized and
concentrated to being spread out. Entropy is a measure of this “spreading of energy.” Entropy-
increasing processes include melting, vaporizing, mixing, and increases in temperature or volume.
Besides melting and vaporizing, what we look for in chemical reactions to indicate an increase of
entropy are the following: solids (or liquids) dissolving in a solvent, and reactions in which the
number of gas particles increases. Enthalpy and entropy are the two driving forces of a reaction,
deciding whether it will be spontaneous or not. Processes that are exothermic where the system
increases in entropy are always spontaneous; those that are endothermic where the system
decreases in entropy are always non-spontaneous

Test Yourself:
1. Indicate whether each involves an increase or decrease in entropy for the system described
a. When you cool a supersaturated solution of sodium acetate, lots of solid crystals form
b. A human being spontaneously combusts, turning into vapor
c. A person slowly compresses a sample of gas in a syringe (no change in temperature occurs)
d. Water vapor from the air condenses into little droplets of liquid water (“fog”) on your window

2. For each statement, indicate which factor (enthalpy or entropy) affects spontaneity and whether it favors
spontaneity or non-spontaneity
a. “That reaction would result in 3 reactant gas molecules turning into 8 product gas molecules.
b. “∆Hrxn = –400 kJ
c. “There are 2 gas molecules in the reactants and 3 solid molecules in the products.
d. “As a result of the reaction, the energy of the system spreads out spatially.

3. Do you know for sure if any of the things in Question 2 will actually be spontaneous

10.8 Spontaneity and Temperature


10.8.1 Enthalpy vs. Entropy: Who wins?
We have seen that there are two driving forces that will cause a process to be spontaneous: a
decrease in enthalpy and an increase in entropy. If a process has both a decrease in enthalpy and
increase in entropy, it is certain to be spontaneous. If a process involves both an increase of
enthalpy and decrease in entropy, it is going to be non-spontaneous

De nitely Spontaneous De nitely Non-Spontaneou


∆H < 0 and ∆S > ∆H > 0 and ∆S <

For example, take the reaction N2H4(g) ➞ N2(g) + 2H2(g) ∆H = –50.4 k


It is an exothermic reaction (∆H is negative), and entropy is increasing (more gas particles in the
products than in the reactants). Thus, this reaction will occur spontaneously, fo' shizzy

Once again, we remind you that “spontaneous” here does not mean “instantaneous.” It just means
it will occur in the direction as written without any force or impetus from an outside source. Iron
rusting is a spontaneous reaction (you never nd some rusted iron “un-rusting”), but it occurs very
slowly. Determining the spontaneity of a reaction tells you only which way a reaction will proceed,
not how fast it occurs.

Now take the reaction 3O2(g) ➞ 2O3(g) ∆H = +142.2 k

384 Chapter 10 Thermodynamics


fi

0



fl
fi

fi

It is endothermic (∆H is positive), and entropy is decreasing (fewer gas particles in the products
than in the reactants). Thus, this reaction will not be spontaneous in the direction written. In fact,
the reverse reaction will be spontaneous (remember that when you reverse a reaction, the sign of “What do you mean, ‘the reverse
∆H changes) reaction’?” you ask. In Ch. 8, we’ll
2O3(g) ➞ 3O2(g) ∆H = –142.2 k learn about reversible reactions:
reactions that can go both left to right
If you threw a mole of ozone (O3) and diatomic oxygen (O2) in a container together, the ozone will and right to left.
want to turn into oxygen, not vice versa.

Reiterating, we can be safe to conclude the following:


★ If ∆H < 0 and ∆S > 0, the reaction will be spontaneous as writte You best memorize this!
★ If ∆H > 0 and ∆S < 0, the reverse reaction will be spontaneou
But what happens if ∆H<0 and ∆S<0, or if ∆H>0 and ∆S>0? The two driving forces are competing
against each other, but which wins: enthalpy or entropy?

There is a qualitative way to decide. If the reaction is barely exothermic or endothermic, yet there is
a large change in entropy, entropy will have a greater effect and determines spontaneity. We would Most reactions are affected more by
enthalpy than by entropy. That is,
say that “entropic” effects win out. If the reaction is hugely exothermic or endothermic, yet there usually enthalpic effects win out. But
isn’t much difference between the entropy of the products and reactants, the enthalpy will have a if there is a large change in entropy,
greater effect and determine spontaneity. We would say that “enthalpic” effects win out. entropy might win.

However, this is an unreliable method in most cases. Besides, we need to study one more factor
that we must take into account when deciding if a reaction will be spontaneous: temperature

Example 17:
For each of the following, state whether the forward reaction will be spontaneous or non-spontaneous
a. 6C(graphite) + 3H2(g) ➞ C6H6(l) ∆Hrxn = +49 k
b. C3H8(g) + 5O2(g) ➞ 3CO2(g) + 4H2O(g) ∆Hrxn = –2043 kJ
Solution:
a. Notice how we go from having three gas molecules (and some solid atoms) to having only a liquid
molecule. Entropy is decreasing (∆S<0). This does not favor spontaneity. Also, the system is gaining enthalpy
(∆H>0), which is also bad for spontaneity. Therefore, we can predict this reaction is non-spontaneous
b. The entropy of the system is increasing (∆S>0) because there are more gas particles in the products than
in the reactants. This is good for spontaneity. Combine that with the fact that it is exothermic (∆H<0, also
good!) tells us that this reaction will be spontaneous

10.8.2 Reactions that could go either way depend on temperature


Let’s take a look at ice melting H2O(s) ➞ H2O(l) ∆H = +6.02 kJ/mo
∆H is positive (energy must be absorbed to melt a solid), but ∆S is also positive (liquids have more
entropy than solids). In other words, for this reaction

✓ Entropy favors spontaneit


✓ Enthalpy favors non-spontaneit
If entropy had its way, ice would melt spontaneously. If enthalpy had its way, it wouldn’t. In fact,
enthalpy says that the reverse reaction (liquid water freezing) should occur spontaneously. What do
we observe that could give us a clue to which factor wins out? Do we always see ice melting or do
we always see liquid water freezing? It would be very perceptive if you said, “It depends on the
temperature.” We could see either process happen, depending on the temperature
❖ Above 0ºC, ice melts spontaneously (in other words, ∆S won out
❖ Below 0ºC, liquid water freezes spontaneously (in other words, ∆H won out
So it appears that higher temperatures seem to give entropy more “weight,” but lower temperatures
seem to help enthalpy win.

Let’s try water vapor condensing H2O(g) ➞ H2O(l) ∆H = –40.8 kJ/mo


∆H is negative (energy is released when a gas condenses), but ∆S is also negative (respect the
king!). In other words, for this reaction
✓ Enthalpy favors spontaneit
✓ Entropy favors non-spontaneit

Chapter 10 Thermodynamics 385


:

:
y

:

Enthalpy wants the reaction to occur as written; entropy doesn’t. What do we observe? Do we see
gaseous water condensing, or do we see liquid water vaporizing? Once again, either process could
happen, depending on the temperature
❖ Below 100ºC, gaseous water condenses spontaneously (∆H won out
❖ Above 100ºC, liquid water vaporizes spontaneously (∆S won out
Again, it appears that higher temperatures make entropy win. But lower temperatures make
enthalpy win

Looking at these last two examples, we can conclude not only that temperature affects spontaneity,
We’ll see why that’s the case later in but something more speci c. At high temperatures, entropy dominates. At low temperatures,
the chapter. enthalpy dominates

Here is a table of what we learned so far (color coded red for “bad” and green for “good,” meaning
bad or good for the forward reaction to be spontaneous)

∆S > ∆S <
(entropy increasing) (entropy decreasing)
Don’t bother trying to “memorize” this ∆H > Spontaneous only at Non-spontaneous at all
table. Rather understand it so it (endothermic) high temperatures temperatures
makes sense to you.
∆H < Spontaneous at all Spontaneous only at
(exothermic) temperatures low temperatures

How high or low of a temperature though? Can we be more speci c? Yes. We’ll learn how to
de nitively determine if a reaction will be spontaneous; but before that, let’s see how to use the
knowledge we’ve acquired

Example 18:
Below are thermodynamic data for two reactions. Decide if each is always spontaneous, never
spontaneous, spontaneous only at high temperature, or spontaneous only at low temperatures. (Try to do
this without simply looking at the table above.
Reaction 1: ∆H = +225 kJ ∆S = –150 J/
Reaction 2: ∆H = –450 kJ ∆S = –340 J/K
Solution:
Reaction 1 is never spontaneous at any temperatures because it is endothermic and entropy decreases
Reaction 2 has an enthalpy change that favors spontaneity, but an entropy change that favors non-
spontaneity. Since enthalpy dominates at low temperatures, it is spontaneous at low temperatures

10.8.3 Making reactions happen by manipulating temperature


We can use the information in the last section practically. As an example, let’s say we are trying to
get the following reaction to occur
2 H2O(l) ➞ 2 H2(g) + O2(g) ∆H = +571.6 k

Analyzing this reaction, ∆H is positive (an endothermic reaction), and ∆S is also positive (more
gaseous particles in the products). This means that entropy favors spontaneity, but enthalpy favors
non-spontaneity. If we want to get the reaction to occur in the direction written, we have to give
extra “weight” to entropy to help it win. How do we do that? By increasing the temperature. At high
temperatures, we can make entropy win out over enthalpy.

What if we want this reaction to occur as written

N2(g) + 3 H2(g) ➞ 2 NH3(g) ∆H = –92.2 k

It is exothermic (good for spontaneity), but entropy is decreasing (bad for spontaneity). This means
that we do not want entropy to have much of an effect. We do this by lowering the temperature.
This reaction will occur at low temperatures. The reverse reaction might occur at high
temperatures. (How low or how high? We’ll learn in §10.10.

So to sum up: high temperatures maximize the effect of entropy, and low temperatures minimize
the effect of entropy

386 Chapter 10 Thermodynamics


fi
0

fi
:

fi
.

Example 19:
Below is thermodynamic data for a reaction. If you want the reaction to occur, should you choose to
perform the reaction at low temperature or at high temperature
∆H = +139 kJ ∆S = +203 J/K
Solution:
The system is gaining energy (endothermic). That’s bad for spontaneity. But entropy is increasing (positive
∆S). That’s good for spontaneity. To make the reaction occur, we have to help entropy “win.” We do this by
performing the reaction at high temperatures

Section Overview

The Forest:
Exothermic reactions involving an increase in entropy are always spontaneous and endothermic
reactions involving a decrease in entropy are never spontaneous. When enthalpy and entropy are
opposed, the temperature of the reaction will affect the outcome

The Trees:
Any reaction that is exothermic and also has energy dispersed will be spontaneous, no matter the
temperature. Any reaction that is endothermic and also has energy becoming localized will be non-
spontaneous, no matter the temperature. But when the two driving forces, enthalpy and entropy,
disagree, we must take into account the temperature. It turns out that high temperatures maximize
the effect of entropy. That is, as the temperature increases, entropy has a bigger and bigger “say”
about whether the reaction will occur. To make a particular reaction occur, we choose an
appropriate temperature that will increase or decrease the effect of entropy, depending on if entropy
is “good” or “bad” for spontaneity

Test Yourself:
1. Will this reaction be spontaneous? X3Z2(s) + 2A(l) ➞ 2AZ(s) + 3X(g) ∆Hrxn = –300 k

2. For each of the following, decide if it will be always spontaneous, never spontaneous, spontaneous at high
temperatures, or spontaneous at low temperatures
a. 4 B(s) + 3 O2(g) ➞ 2 B2O3(s) ∆Hrxn = –300 k
b. 4 C3H5N3O9(l) ➞ 12 CO2(g) + 10 H2O(g) + 6 N2(g) + O2(g) ∆Hrxn = –5680 k
c. 2 NO2(g) ➞ 2 NO(g) + O2(g) ∆Hrxn = +114 kJ

10.9 Calculating Entropy Changes Numerically


Earlier this chapter, we de ned the standard enthalpy of formation, ∆Hfº, as the energy required to
form one mole of a compound from its elements in their standard states at 298 K and 1 atm. There
is a corresponding property for entropy. And just as we used the tabulated values of ∆Hfº to
calculate the ∆Hºrxn, we can use the tabulated values of these “standard entropies” to nd the
∆Sºrxn. This will give us a de nitive, numerical answer for the change in entropy of a reaction rather
than just qualitatively saying it changes based on the number of gas particles or something

10.9.1 Standard molar entropy


Imagine taking a piece of silver and removing heat from it. The particles vibrate with less and less
energy. In other words, entropy will decrease as we lower its temperature. At absolute zero (zero Recall that entropy increases as T
kelvin), the particles should stop vibrating, right? What do you think is the entropy at that increases, so entropy must decrease
temperature? Zero. In fact, this idea is important enough to be the third law of thermodynamics: as T decreases.
the entropy of a perfect crystal at zero kelvin is zero

Recall that the change in entropy of a system is the energy dispersed into it divided by the
temperature at which the dispersal occurs: ∆S = q/T. Now let’s de ne the standard molar entropy,
∆Sº. It is the amount of heat (divided by the temperature) that needs to be dispersed to one mole of
a perfect crystal at zero kelvin to get it to its usual state at 298K. In other words, you start with a Technically speaking, the entropy of
perfect crystal of the substance, say silver, at zero kelvin. To get the silver to 298K, you add a bit of a substance is the motional energy
added to it plus any potential energy
heat in tiny increments, dividing by the kelvin temperature each time (this is to keep the process
(latent heat) from phase changes.
reversible). Do this over and over until you reach 298K. To get ∆Sº, you simply add up all these tiny
q/T values. For silver (see your table), ∆Sº = 42.6 J/Kmol. (That doesn’t mean you added only
42.6J to get the silver to 298K. You probably added thousands of joules.

Chapter 10 Thermodynamics 387

fi
fi

fi

fi
That’s ne and dandy that you got all those details, but understand the rough idea of what ∆Sº
represents. The bigger the value of ∆Sº, the more energy that was required to get that substance to
298K. For example, the table says that the ∆Sº of liquid water is 69.9 J/Kmol while that of gaseous
water is 188.8J/Kmol. That only makes sense. Both require the same amount of energy to get them
from a perfect solid H2O crystal at 0K to liquid H2O at 298K, but the gaseous water takes that extra
I talked about this in §10.7.9: how amount of energy to vaporize it. That extra energy is re ected in the larger ∆Sº value. In that sense,
gases have higher entropy than a substance with a bigger ∆Sº is at a “higher entropy state.” A substance with a smaller ∆Sº is at a
liquids because of the extra energy “lower entropy state.” That’s why we often talk in terms of the “entropy of a substance,” as if it is
that needs to be dispersed into the
liquid to turn it into a gas. something a substances possesses, despite having de ned entropy as an energy dispersal. So
don’t get thrown if we say “gases have a larger entropy than liquids.” We just mean a large energy
dispersal was required to form it

Our table lists values for the standard molar entropy of many chemicals. The larger the number, the
higher the entropy of that chemical. For example, solid aluminum has a value of 28.3 J/Kmol while
solid AlCl3 has a value of 110.7 J/Kmol. Thus, AlCl3 has a higher entropy than Al. As we’ll see in the
next subsection, we can compare the total entropy of the products to the total entropy of the
reactants to see if the overall entropy is increasing or decreasing. From that, we can decide if
entropy truly is favoring spontaneity or not. Also, armed with an actual numerical value of ∆S, we
can compare it to the numerical value of ∆H to see which one dominates and will “win.” We’ll do
that in §10.10. It’s all coming together nicely now

Before we proceed, here are a few notes.


The standard molar entropy is the difference between the entropy of the substance at 0K
and 298K (∆Sº = S298K – S0K). But the entropy at 0K is de ned to be zero. Thus, the
change in entropy to get the substance to 298K (∆Sº) is the same as its nal entropy. For
that reason, ∆Sº is typically written simply as Sº (without the delta), like in the
thermodynamic table we gave you.
Take notice that – unlike the ∆Hfº values – the standard molar entropy of an element in its
standard state is NOT equal to zero. When doing calculations, don’t mistakenly assume
they are zero
Also be aware that the energy unit for Sº in the table is joules per kelvin (per mole),
whereas the energy unit for ∆Hfº in the table is kilojoules (per mole). We’ll have to be
convert if we are going to add or subtract enthalpies and entropies

10.9.2 Calculating ∆Srxn


We just de ned the standard molar entropy, Sº. Basically, it numerically describes the entropy of a
substance at 298K. The bigger the value, the greater its entropy. For example, if you look at your
I’ll drop the “per mol” in the unit to table, liquid water has Sº = 69.9 J/K and gaseous water has Sº = 188.8 J/K: gaseous water has a
save some space, but remember larger standard molar entropy. Now consider the vaporization of liquid water at 298K
that all the data in the thermo table is
per mole. H2O(l) (Sº = 69.9 J/K) ➞ H2O(g) (Sº = 188.8 J/K)
Notice how we are going from a “lower entropy state” to a “higher entropy state”: the entropy is
increasing. This is exactly what we concluded qualitatively in the earlier section on entropy.

We can calculate entropy changes in a reaction the same way we calculated ∆Hrxn. For a given
reactio
aA + bB ➞ cC + d
The change in entropy of the reaction, ∆Sºrxn, is calculated by taking the standard molar entropy of
the products (each multiplied by its coef cient) and subtracting the standard molar entropy of the
reactants (each multiplied by its coef cient)
∆Sºrxn = cSº(C) + dSº(D) – aSº(A) – bSº(B
∆Sºrxn = ∑ Sºproducts – ∑ Sºreactants
(Remember that the standard molar entropy of elements in their standard states is NOT zero. Don’t
forget to include them!) When ∆Sºrxn > 0, entropy is increasing and entropy favors spontaneity;
when ∆Sºrxn < 0, entropy is decreasing and entropy favors non-spontaneity. This is exactly what we
concluded earlier. For the vaporization of water above, we’d writ
∆Sºrxn = 188.8 J/K – 69.9 J/K = 118.9 J/
Because ∆Sº > 0, entropy is increasing and entropy favors spontaneity. No surprise there.

388 Chapter 10 Thermodynamics


n

fi
fi
.

fi
K

fi

fl
fi
fi
e

fi
:

Example 20:
Given the following reaction, use your thermodynamic table to determine the ∆Sºrxn. Then state whether
entropy favors spontaneity or non-spontaneity
2 NO2(g) ➞ 2 NO(g) + O2(g)
Solution:
Simply look up the values for So in your table, then do “products minus reactants” (each multiplied by their
coef cients). Keep in mind that, unlike ∆Hfº values, the Sº of element in its standard state is NOT zero.

2 NO2(g) ➞ 2 NO(g) + O2(g)


So (J/Kmol): 240. 210.8 205.

∆Sºrxn = ∑ Sºproducts – ∑ Sºreactants = 2(210.8 J/Kmol) + (205.1 J/Kmol) – 2(240.1 J/Kmol) = 146.5 J/

Because ∆Sº > 0, entropy is increasing and we say that entropy favors spontaneity. It makes perfect sense that
∆Sº > 0 because we went from two gas molecules to three gas molecules. Energy is being dispersed, so
entropy is increasing

Section Overview

The Forest:
We can use our thermodynamic table to calculate numerical values for the entropy change of a
reaction. We just do “products minus reactants.” If ∆Sºrxn > 0, entropy is increasing and entropy
favors spontaneity; when ∆Sºrxn < 0, entropy is decreasing and entropy favors non-spontaneity

The Trees:
The standard molar entropy of a substance is related to how much energy we need to transfer to a
perfect crystal at zero kelvin to get it up to 298K. In other words, the values in our thermodynamic
table tells us how much energy had to dispersed into something to get it to normal temperatures. In
that way, it sort of tells us how much entropy something “has.” Something with a larger standard
molar entropy is in a “higher entropy state.” By comparing the standard molar entropies of reactants
and products, we can calculate if entropy is increasing or decreasing. We simply add the standard
molar entropies of the products (each multiplied by its coef cient) and subtract the standard molar
entropies of the reactants (each multiplied by its coef cient). A positive ∆Sºrxn means entropy is
increasing. In that case, spontaneity is favored. (However, we’ll need to take into account enthalpy.)

Test Yourself:
1. Calculate the change in entropy for the reaction 2 H2O(l) + O2(g) ➞ 2 H2O2(l

2. Let’s say that the reaction above has a ∆Hºrxn < 0. Is the reaction always spontaneous, spontaneous only at high
temperatures, spontaneous only at low temperatures, or never spontaneous (that is, the reverse reaction would be
spontaneous)

10.10 Gibbs Free Energy


10.10.1 Gibbs Free Energy will determine if a process is spontaneous
We have seen that the spontaneity of a process depends on the change in enthalpy, the change in
entropy, and the temperature at which the process occurs. We combine all of these together into an
equation that takes into account the temperature in uence. We de ne the Gibbs Free Energy, G, In the following Interchapter, I’ll
as explain what Gibbs Free Energy
means qualitatively.
G=H–T

More usefully, we de ne the change in Gibbs Free Energy of a reaction, ∆Grxn, a


ΔGrxn = ΔHrxn – TΔSrxn
...or, more simply..
∆G = ∆H – T∆

Chapter 10 Thermodynamics 389


fi
S

fi

1
.

fl
fi
fi
1

fi
)

(Again, we sometimes drop the “rxn” subscripts for convenience. We’ll assume that they always
refer to the reaction/system, unless stated.) Please note that you must use kelvin for temperature!

This might be a little confusing since For any process to be spontaneous, ∆G < 0 (what we call exergonic). If ∆G > 0 (what we call
we haven’t yet studied reversible endergonic), the reaction is non-spontaneous; however, the reverse reaction will be spontaneous
reactions (Ch. 8). But for now (∆G’s behave just like ∆H’s: if you reverse the chemical equation, the sign of ∆G changes).
understand that most reactions can
go “both ways”: the reactants can
turn into the products and the What happens if ∆G = 0? Let’s think about this. Take the generic reactio
products can turn into the reactants.
Both the forward and reverse X➞
reactions are possible.
If ∆G were less than zero, the forward reaction would be spontaneous and would proceed left to
right as written: X would turn into Y. If ∆G were greater than zero, the reverse reaction is
spontaneous: Y would turn into X. Exactly at ∆G = 0, which would happen? Both! X is forming Y
and Y is forming X at the same rate. This is called equilibrium, a topic we’ll get to soon. Whenever
we have a system at equilibrium (that is, there is no net change in the amount of reactant or
product), we assume that ∆G = 0. Summarizing
If ∆G < 0 (exergonic), the forward reaction is spontaneou
Another thing for you to memorize/ If ∆G > 0 (endergonic), the forward reaction is non-spontaneous (the reverse reaction is spontaneous
understand. If ∆G = 0, the reaction is at equilibrium (the topic of Chapter 12

10.10.2 Does the ∆G equation make sense?


First, let’s examine a spontaneous reaction. ∆G better return a result of less than zero, right? What
conditions de nitely make for a spontaneous reaction? A negative ∆H and a positive ∆S. Now
remembering that we must measure T in kelvin (i.e. we will never plug in a negative temperature),
the equation will look something lik

∆G = ∆H – T∆S = [–] – [(+)(+)] = negativ

It works so far. What about a de nite non-spontaneous reaction? ∆G better return a positive
number. For such a reaction, ∆H > 0 and ∆S < 0.

∆G = ∆H – T∆S = [+] – [(+)(–)] = positiv

Another thing that we discussed earlier is that when ∆H and ∆S have the same sign (they are
“competing”), high temperatures give entropy more weight, while low temperatures minimize the
effect of entropy. Does our equation do this? Yes. Let’s say that both ∆H and ∆S are positive. This
means enthalpy favors non-spontaneity while entropy favors spontaneity. We said that if we want
entropy to “win,” we better make sure the temperature is high. Look at the equation. With a large T,
the second term (T∆S) becomes very large. If the magnitude of this negative term is larger than the
magnitude of the ∆H term, the reaction is spontaneous. However, if T is too low, the term (T∆S) will
be correspondingly small, and ∆H will win out. Seems to work, so let’s use it!

10.10.3 Using ∆G to determine de nitively if a reaction is spontaneous (FINALLY!)


To determine if a reaction is spontaneous, all we have to do is plug into the Gibbs Free Energy
equation the values of ∆Hºrxn, ∆Sºrxn, and T (in kelvin!). Sometimes, the values of ∆H and ∆S will be
given to you. Other times, you’ll have to use your table and do “products minus reactants” to solve
for them.

You might have noticed that the table lists values of ∆Gfº too. Can we just skip using ∆H’s and
∆S’s? In other words, can we nd ∆Gºrxn just by doing the following

∆Gºrxn = ∑ ∆Gfºproducts – ∑ ∆Gfºreactants

Yes, we can, but only when the temperature of the reaction is at 298K. The values for ∆Gfº are very
dependent on temperatures. Those in the table are only valid for 25ºC. If the temperature is
something other than 25ºC, you must calculate ∆H, ∆S, and plug them in with T into the Gibbs Free
Energy equation. Bummer! Take a gander at the following example to see how it is done

Example 21:
Consider the following reaction.We looked up for you the ∆Hfº, Sº, and ∆Gfº of each and put them below
a. Determine whether the following reaction will be spontaneous at 25ºC by calculating ∆Gº.
b.Then determine if it will be spontaneous at 1000ºC.

390 Chapter 10 Thermodynamics


Y

fi

fi
e

fi
fi
e

2C2H6(g) + 7O2(g) ➞ 4CO2(g) + 6H2O(l


∆Hfº(kJ/mol): –84.68 0 –393.5 –285.
Sº(J/Kmol) 229.6 205.138 213.74 69.9
∆Gfº(kJ/mol): –32.82 0 –394.359 –237.129
Solution
a. Since we are at 25ºC, we can simply work with the ∆Gfº’s
∆Gº = 4∆Gfº(CO2) + 6∆Gfº(H2O) – 2∆Gfº(C2H6) – 7∆Gfº(O2
∆Gº = 4(–394.359 kJ/mol) + 6(–237.129 kJ/mol) – 2(–32.82 kJ/mol) – 7(0
∆Gº = –2934.57 kJ
It is exergonic. The negative ∆Gº means it is spontaneous. (By the way, if you had used the Gibbs Free
Energy equation by solving for ∆Hº and ∆Sº rst, you would get the same answer.

b. Because we now want to calculate ∆Gº at a temperature other than 25ºC, we must use the Gibbs
equation. That means we need to nd ∆Hº and ∆Sº. Notice that the units for Sº in the table are in J/K: we’ll
need to convert to kJ/K. We’ll also need to convert 1000ºC to 1273K
∆Hº = 4∆Hfº(CO2) + 6∆Hfº(H2O) – 2∆Hfº(C2H6) – 7∆Hfº(O2
∆Hº = 4(–393.5 kJ/mol) + 6(–285.8 kJ/mol) – 2(–84.68 kJ/mol) – 7(0
∆Hº = –3119.44 k

∆Sº = 4Sº(CO2) + 6Sº(H2O) – 2Sº(C2H6) – 7Sº(O2


∆Sº = 4(213.74 J/Kmol) + 6(69.91 J/Kmol) – 2(229.6 J/Kmol) – 7(205.138 J/Kmol
∆Sº = –620.746 J/K = –0.620746 kJ/K
NOTE: We converted entropy from
∆Gº = ∆Hº – T∆Sº = (–3119.44 kJ) – (1273 K)(–0.620746 kJ/K) = –2329 kJ J/K to kJ/K. We could have instead
converted enthalpy from kJ to J.
Spontaneous. Compare this to the value in part (a). Clearly, the result is affected by the temperature.

“Har Har! You have to do homework, nerds! Did your answer for (b) make sense compared to (a)?
Notice that enthalpy favors spontaneity for this reaction, but entropy does not favor spontaneity
(there are fewer gas molecules in products, which is why we got a negative ∆Sº). By increasing the
temperature to 1000ºC, we are increasing the effect of entropy. That is, we are making entropy
have a stronger say in the matter than it did at the lower temperature. That’s why the value for ∆Gº
isn’t as negative as at 25ºC. However, it is still negative and spontaneous. With a little thought, you
can imagine that there could be some temperature high enough that we give entropy so much
weight that the reaction might become non-spontaneous. That’s the topic of the section below.

10.10.4 An important note about the meaning of “spontaneous”


Let’s say for some reaction X ➞ Y, you calculate ∆Gº > 0. You conclude that this reaction, as
written, is non-spontaneous. Do NOT interpret this as meaning the reaction X ➞ Y can never
happen. If you had calculated that ∆Gº < 0, you’d conclude the reaction as written is spontaneous.
Do not interpret this as meaning the opposite reaction (Y turning into X) can never happen. That’s
not what spontaneity means. When we conclude a reaction is spontaneous, we mean only that the
forward reaction is the preferred direction. That is, the reaction will continue to the right (towards
making more product Y). When we conclude a reaction is non-spontaneous, we mean only that the
forward reaction is not preferred: rather the reverse reaction is preferred. This means that the
reaction would actually want to go the other way to make more X. So spontaneity isn’t about which
reactions are allowed and which are forbidden. Spontaneity tells us only which way the reaction
would like to go. If X ➞ Y is spontaneous, X will want to turn into Y. If X ➞ Y is non-spontaneous, Y
will want to turn into X I’ll go into this a little more in §10.11.2.

10.10.5 Finding the critical temperature at which a reaction becomes spontaneous


We can use the Gibbs Free Energy equation to determine what temperature a reaction will go from
being non-spontaneous to spontaneous. It is non-spontaneous when ∆Grxn > 0 and then becomes
spontaneous when ∆Grxn < 0. So the critical point is right when ∆Grxn = 0. The temperature that
gives a ∆Grxn = 0 is the “magic” temperature at which a reaction will become either spontaneous or
non-spontaneous. So let’s set ∆G = 0 and solve
∆G = 0 = ∆H – T∆S ...thus... T = ∆H/∆S
Let’s see an example. Take the endothermic process of water boiling at sea level

Chapter 10 Thermodynamics 391


:

fi
:


fi
)

H2O(l) ➞ H2O(g)
∆H = +44 kJ/mol ∆S = +0.119 kJ/mol
At what temperature should this process become spontaneous? Plug in the values and solve
T = ∆H/∆S = (44kJ/mol)/(0.119kJ/molK) = 370
Hmm. Does that temperature sound familiar? That is about 100ºC, where we do observe the
spontaneous boiling of water at sea level. Because entropy favors spontaneity, we want the
temperature to be at least this high; so this is the minimum temperature needed
Now let’s look at an exothermic reaction
N2(g) + 3 H2(g) ➞ 2 NH3(g)
∆H = –46.3 kJ/mol ∆S = –0.199 kJ/mol
T = ∆H/∆S = (–46.3kJ/mol)/(–0.199kJ/molK) = 233
This is the critical temperature, but will it become spontaneous above or below this temperature?
Because entropy favors non-spontaneity, we want to minimize the effects of entropy, so the
reaction will become spontaneous below 233K.

Example 22:
Some reaction occurs that has the following thermodynamic data. At what temperatures will the reaction
be spontaneous
∆H = +135 kJ/mol ∆S = +250 J/molK
Solution:
Using the equation for the critical temperature (converting ∆S to kJ
T = ∆H/∆S = (135 kJ/mol)/(0.250kJ/molK) = 540
Okay, that’s the critical temperature. Will it be spontaneous above or below that temperature? Notice that
entropy favors spontaneity. If we want the reaction to be spontaneous, we have to help entropy “win.” We
do this by using high temperatures. Therefore, our answer is that it will be spontaneous at temperatures
above 540K.

Section Overview

The Forest:
The Gibbs Free Energy equation is to our thermodynamics chapter what the ideal gas law is to the
gas chapter. It puts together all the factors that in uence spontaneity – enthalpy, entropy, and
temperature – into one equation that de nitively tells us if a reaction will occur as written

The Trees:
A reaction will be spontaneous if the Gibbs Free Energy change is less than zero (exergonic). If the
reaction is occurring at 25ºC, we can simply use the ∆Gfº values listed in the table. At other
temperatures, we must calculate ∆H and ∆S separately and plug them into the ∆G equation. We
can also use Tcrit = ∆H/∆S to nd the temperature at which a reaction becomes spontaneous.

Test Yourself:
By the way, IUPAC and NIST have 1. Use the information in your thermodynamic data table to nd ∆Gº at 100ºC and determine if the following
recommended writing the subscripts reaction is spontaneous at that temperature. (At this T, can you use the ∆Gfº values in the table?) Watch the units!
after the delta for thermodynamic
data. Thus, for example, we would 2 SO3(g) ➞ 2 SO2(g) + O2(g)
write ∆fHº instead of ∆Hfº and
∆rGº instead of ∆Gºrxn (“rxn” is
abbreviated to “r”).

2. Use the ∆Hº and ∆Sº you (hopefully) already calculated in #1 to nd the temperature range at which the reaction
would be spontaneous. (Make sure you state it is “above” or “below” the temperature you calculate.

392 Chapter 10 Thermodynamics

fi

fi
K

fl
fi
fi
)

10.11 Clari cations and Teasers


In this “postlude,” I want to clarify a couple of important points. In the short run, these won’t affect
your ability to do the following problems; but in the long run, understanding them will help you get
how thermo works in the greater scope of chemistry. But rst, I’ll bring up an interesting (and
exceedingly important) idea as a teaser to entice you to read the following Interchapter.

10.11.1 Pssst! Here’s a secret: spontaneity really depends only on entropy!


“Whaaa?! I thought you told us it depended on entropy and enthalpy.” Yes, this is the way chemists
handle it. In a chemical reaction, two things are happening. For one, bond energy is absorbed or
released, which is given by the enthalpy change, ∆H. For another, the identities (and thus entropy
states) of the chemicals are changing (number of gas particles, solid vs. liquid, solids vs. ions, etc.),
which is given by the entropy change, ∆S. Chemists treat those ideas separately

However, the “real” rule is the following: a reaction will be spontaneous if the entropy of the
universe increases. That is, if ∆Suniv > 0, the reaction will be spontaneous. It is one of those
fundamental and crucial laws of physics, like ∑F = ma or conservation of mass-energy. In fact, it is
the 2nd law of thermodynamics. But where is our ol’ buddy enthalpy? I don’t want to get into it too
much here: I’ll leave it for the curious-minded who want to delve into the Interchapter. But here’s
the basic idea. All this time, we’ve been ignoring the change in entropy of the surroundings! All of
our examples were about what’s happening to the system (the reactants and products). We never
stopped to consider what’s happening to the surroundings. Let’s see how that changes things.

Notice how I said that the entropy of the universe must increase for a spontaneous process. Recall
from looooong ago at the beginning of this massive chapter that we divided the universe into the
system (stuff of interest) and the surroundings (everything else). Thus, ∆Suniv = ∆Ssys + ∆Ssurr, and
we can rewrite the 2nd law a
If ∆Ssys + ∆Ssurr > 0 the reaction will be spontaneous.
“Okay, I get it, but where is ∆H? Tell me. I demand to know!” Remember that we said exothermic
reactions tend to be spontaneous. What happens with heat during exothermic reactions? It goes
from the system to the surroundings, right? Hmmm...heat being dispersed into the surroundings.
What does that imply about the entropy of the surroundings? It increases! The very fact that
exothermic reactions disperse energy to the surroundings means that ∆Ssurr > 0 for exothermic
reactions. Conversely, we say that endothermic reactions tend to be non-spontaneous. Why?
Because they steal thermal energy from the surroundings, which means ∆Ssurr < 0. That’s why
enthalpy is said to affect spontaneity. But from another perspective, enthalpy is important only in
how it affects the entropy of the surroundings. So we recast our rules of spontaneity only on the
basis of entropy, which is the real rule.

In the following interchapter, I’ll revisit this idea, eshing it out a bit. Even more fun and revelatory is
that we’ll see that the Gibbs Free Energy Equation is really just the 2nd law of thermodynamics in a
chemistry disguise. So awesome! Additionally, we’ll use this new concept of ∆Suniv to understand
why heat must always ow from the hotter object to the colder object. And nally, I’ll explain
something that the most perceptive of you might have noticed. Earlier, I said that a hot pan cooling
off is a spontaneous process. I used that to help de ne entropy. But later, I said that heating up a
pan involves an increase in entropy, which favors spontaneity. Huh? Which one is it? Again, it all
has to do with ∆Suniv needing to be > 0

10.11.2 Does “non-spontaneous” mean that the forward reaction can’t happen?
In §10.10, we used thermodynamic data (∆H’s and ∆S’s) to determine if a reaction is spontaneous. Don’t let this section confuse you.
We found out that if ∆Gº < 0, the forward reaction is spontaneous, and if ∆Gº > 0, the forward You can ignore it and will be just fine
reaction is non-spontaneous, meaning the reverse reaction is the direction that is preferred. But to solve all our questions. However,
if you go on to higher levels of
let’s understand this a little better. chemistry, it’s nice to understand it.
Take the simple reaction A ➞ B. Pretend that we calculate the ∆Hº and ∆Sº using the thermo table
and nd out ∆Gº > 0. To make it more extreme, say that it is endothermic and entropy is decreasing
(∆Hº > 0, ∆Sº < 0). From what we learned earlier, we’d say the reaction A ➞ B is non-spontaneous
at all temperatures. That sounds so nal and absolute. But does it really mean that reaction will
NEVER happen? Similarly, take some reaction C ➞ D. You use the table to nd that ∆Hº < 0 and
∆Sº > 0. We’d say that the reaction is spontaneous at all temperatures. Does that mean that this
reaction will ALWAYS happen no matter what, and the reverse reaction could never happen? The

Chapter 10 Thermodynamics 393


fi
fi
fl
s

fi

fl

fi
fi

fi
.

fi

answer to both questions is “no.” I already discussed this in §10.10.4. If A ➞ B is non-spontaneous,


it only means that the reaction currently will try to go backwards to make more A, not forwards to
make more B. It does not mean A ➞ B can never happen. I hope you understand that. But let me
continue with a question you might have

“If A ➞ B is non-spontaneous, meaning that B will want to turn into A and not vice versa, doesn’t
that very fact mean that the reaction A ➞ B can’t/won’t happen? After all, if we ever want to claim A
will turn into B, we need to calculate that ∆G < 0. But using the values from the table, we’ll never
get that, will we? So under what conditions could we ever get A ➞ B to be spontaneous?

I’ll give an abbreviated answer here, although you’ll understand it more when you get to Ch. 8. You
can “force” a non-spontaneous reaction to happen by “overloading” the concentrations/pressures of
the chemicals. Take A(g) ➞ B(g). Let’s say we have some A and B sitting in the container and
calculate ∆Gº > 0. That tells us the forward reaction is non-spontaneous, meaning B wants to turn
into A. But if we add gobs of A into the container, we may be able to force the forward reaction to
actually be the spontaneous direction. Similarly for C(g) ➞ D(g): if we put some C and D in a
container and calculate ∆Gº < 0, we say the forward reaction is spontaneous and C will turn into D.
But if we add gobs fo D into the container, we may be able to force the reverse reaction to happen
Changing temperature and volume instead. Thus, you can actually cause a non-spontaneous reaction to become spontaneous if you
could also affect spontaneity. See Le adjust the amounts of the reactants and products
Châtelier’s Principle in Ch. 8.
Now you might object here: “Hey! Then what good is calculating ∆Gº if its sign (+ or –) isn’t going to
tell us de nitively whether a reaction is spontaneous or not because we have to take into account
the amounts of reactants and products?” Good question. Although I won’t get into it here (it’s
beyond the scope of this intro course), there’s an equation that gives you the real ∆G when taking
into account the amounts of reactants and products. Note that it doesn’t have the º superscript. ∆Gº
(calculated by using the values in the thermo table) assumes you have standard amounts of every
reactant or product. When you use the handy-dandy equation to make the corrections for non-
standard amounts, you go from ∆Gº to ∆G. Once you do that, the sign of this new ∆G will tell you
de nitively if the reaction will be spontaneous or not, taking into account the amounts you have

I didn’t bring it up earlier, because I didn’t want you to get confused. I wanted to make it black &
white for you initially. Don’t let it confuse you. For our intro into thermo, assume we always have
standard amounts, and if you calculate a ∆Gº < 0, we’ll say the forward reaction is spontaneous
and the reactants will want to turn into products. If ∆Gº > 0, the opposite will happen.

10.11.3 Why doesn’t propane readily combust in the presence of oxygen?


Consider the combustion of propane (the fuel in the tanks used for gas grills)
C3H8(g) + 5O2(g) ➞ 3CO2(g) + 4H2O(g) ∆Hrxn = –2044 k
Notice that it is exothermic and we end up with more gas particles than we started with. We’d say
that this is spontaneous, right? But if you release propane into the air, it doesn’t combust with the
oxygen in the air. Why not, if the forward reaction is de nitely spontaneous? Don’t attempt to invoke
some kind of reasoning based on the last subsection: even if we overload our container with C3H8 +
O2, that forward reaction will still not occur, even though we say the reaction is spontaneous

The solution goes back to what I’ve been saying all along: “spontaneous” only means the direction
that would naturally want to occur, not how fast or easy the reaction is. The problem is that it isn’t
easy to get that reaction started. Mix propane and oxygen and nothing will happen. But if you light a
match in their presence, kaboom! It takes a little energy input to get that reaction going. It’s like how
the burning of a match is spontaneous, but you need to strike it so the friction will create enough
heat to get the process going. Once the propane and oxygen start reacting or the match starts
burning, the reactions provide enough energy on their own to sustain the reaction and no longer
need that initial “spark.” Then they’ll very happily react quite spontaneously and completely.

The next chapter deals with kinetics: how fast reactions occur. We’ll learn why some reactions need
a little “spark” to get them going, even if they are spontaneous. There’s a little barrier that the
reactants have to overcome in order to react to form the products. So it turns out that while the
combustion of fuels like propane is very thermodynamically favored, it is kinetically unfavored due
to the high barrier they must overcome. Thermodynamics says, “Oooo...I would love for this
reaction to happen. I get to release energy and increase entropy.” Kinetics says, “Sure, but I’m
going to make it very hard for you.”


394 Chapter 10 Thermodynamics


fi
fi
:

fi
J

Problems
10.1 Chemical Energy and Heat 12. TRUE/FALSE Every liquid has a density of 1g/cm3. That
is, 1cm3 (1mL) of every liquid has a mass of 1g
1. What is the First Law of Thermodynamics
a. You do NOT talk about thermodynamics
See §10.1.6 if you need help with these next problems
b. Thou shalt not kill thermodynamics
13. How much heat must be removed from 100 mL of liquid
c. Thermodynamics always gets shottie
water to decrease its temperature from 20ºC to 0ºC? Recall
d. The total energy of the universe is constant
that the speci c heat of liquid water is 1cal/gK = 4.184J/gK.
You can answer either in calories or joules
2. A system gains 100J of energy. That means that the
surroundings lost/gained (circle one) _______J of energy

3. What is 473ºC in kelvin

4. TRUE/FALSE The temperature 20ºC is equivalent to the


14. You have 30mL of water initially at 25ºC. If you add 360
temperature 20 K
calories of heat to it, what will be its nal temperature?
5. TRUE/FALSE A temperature change of 20ºC is equivalent
to a temperature change of 20 K

6. Fill in the blanks with a word from this word bank


temperature, heat, thermal energy, speci c heat
a. _____________ describes how much energy is required 15. A piece of nickel requires 100 J of heat to raise its
to raise the temperature of 1 gram of a substance by 1ºC temperature from 15ºC to 20ºC. If the speci c heat of nickel
b. _____________ describes the total random energy of all is 0.444 J/gK, what is the mass of the piece of nickel
the atoms, ions, or molecules of a substance
c. _____________ describes the ow of energy from one
substance to another
d. _____________ describes the average kinetic energy of a
single atom, ion, or molecule of a substance
16. You take 29g of iron, initially at 100ºC, and place it in 40g
7. Consider the following data for some mystery substance of water initially at 18ºC. As a result, the water increases to
undergoing a change in temperature due to the the addition 24ºC while the iron cools to 24ºC. (We assume the water
of heat and iron – being mixed – will end up with the same nal T.
25 1.6 J/g 3.3 a. Use the mass of water, its temperature change, and its
a. Which value represents the temperature, T speci c heat to calculate the heat the water absorbed from
b. Which value represents the heat absorbed, q the piece of iron. That is, nd qwater.
c. Which value represents the speci c heat, s

8. Convert 1000 J into calories.

b. Assume that all the heat lost by the iron was completely
absorbed by the water (what you calculated in part a). So
9. Consider 100mL of water at 10ºC and 50mL of water at what is the heat lost by the iron? That is, what is qiron
10ºC.
a. Which has a higher average kinetic energy c. Now you know qiron. Use that, along with the mass of iron,
b. Which has more thermal energy and the change in temperature of the iron, to calculate the
speci c heat of iron
10. A piece of iron is at 20ºC. You increase its temperature to
25ºC. Circle all the things that have increased
i. the average KE of the particle
ii. the total KE of the particle
iii. the thermal energy of the iro
iv. the speci c heat of the iro
v. the average speed of the particle 10.2 Enthalpy
17. Circle the correct choices: A positive ∆H means that the
11. You have 50mL of liquid water reaction is endothermic/exothermic and that heat will be
a. How many cm3 of liquid water do you have transferred from the system/surroundings to the system/
b. How many grams of liquid water do you have surroundings. As a result, the temperature of the
surroundings increases/decreases.

Chapter 10 Thermodynamics 395


fi
fi

fi
J

fi

fi
?

K
?

fl
fi
fi

fi
.

fi
.

fi
?

18. Identify each as an endothermic or exothermic process 23. Given the following information
a. ENDO/EXO The sign of ∆H is positive. 4C(gr) + 2O2(g) ➞ 4CO(g) ∆H = –440k
b. ENDO/EXO As a reaction occurs in a beaker, it gets hot
What would be the ∆Hrxn of the following reaction
c. ENDO/EXO Wood burning in a replace
d. ENDO/EXO The melting of an ice cube 2CO(g) ➞ 2C(gr) + O2(g)
e. ENDO/EXO A reaction occurs in which heat is transferred
from the surroundings to the system

19. The gure below represents the enthalpy changes for the 24. Given the following information
chemical reaction: SO2(g) ➞ S(s) + O2(g) A + B + 100kJ ➞ A
What would be the ∆Hrxn of the following reaction
4AB ➞ 4A + 4
Enthalpy (Energy)

kJ

10.3 Finding ΔH by Calorimetry


297

25. You perform the reaction below and measure that the
system absorbs 150 kJ of heat when you react 0.20 mole of
X. What is the heat of reaction, ∆Hrxn?
a. Label the two levels with the correct atoms and molecules
b. Is the reaction endothermic or exothermic X+Y ➞ X
c. Which has more energy: the SO2 or the S+O2
d. What is the change in enthalpy of reaction, ∆Hrxn

20. Write a thermochemical equation based on this diagram: 26. You perform the reaction below, using 14g of N2 and an
excess of O2. You measure a heat ow to the surroundings
of 126 kJ. Find ∆Hrxn for the reaction as written. Make sure
2H2(g) + O2(g)
Enthalpy (Energy)

you include the correct sign


N2(g) + O2(g) ➞ 2NO(g)
483.6 k
J

2H2O(g)
27. You perform the reaction below. When you use 96 grams
of C, you measure a heat change of 208 kJ, and you nd
that the surroundings got colder. What is ∆Hrxn?
21. Draw an energy diagram (such as in #19/#20) for the
reaction: CO2 ➞ C + O2 ∆H = +400k 2C(gr) + 2H2(g) ➞ C2H4(g)

28. You perform the reaction shown below in a calorimeter.


You measure that the 50 grams of solution (water + LiCl) in
the calorimeter changed from 10.0ºC to 27.7ºC.
22. Which of these represent the endothermic reactio LiCl(s) ➞ Li+(aq) + Cl–(aq)
A + B ➞ AB? (There are three correct answers. a. Was this an endothermic or an exothermic reaction
b. Find qrxn, the heat change of the system, remembering to
AB
Enthalpy (Energy)

Enthalpy (Energy)

A+B use the correct sign.

AB A+B
a. b.
c. If you got this q by using 0.1 mole of LiCl, nd ∆Hrxn
c. A + B + 200kJ ➞ A
d. A + B ➞ AB + 200k
e. A + B ➞ AB ∆H = –200k
f. A + B ➞ AB ∆H = +200k

396 Chapter 10 Thermodynamics


fi
Y

fi
:

fl

fi
)

fi
.

29. A perfectly-insulated calorimeter contains 100 mL of 31. Consider the reactio


water initially at 19.0ºC. After dissolving 15 grams of KCl, the C(gr) + 2H2(g) ➞ CH4(g) ∆Hrxn = –75 k
solution drops to 10.7ºC.
If 48 grams of graphite react, how much energy is given off
KCl(s) ➞ K+(aq) + Cl–(aq) or absorbed (state which one)? (See Example 8 for help.
a. Was the reaction endothermic or exothermic
b. What was the total mass of solution

c. Find qrxn, remembering to include the correct sign

32. Consider the reactio


2H2O(l) ➞ 2H2(g) + O2(g) ∆Hrxn = +571 k
If 100 grams of H2O decompose, how much energy is given
d. What is ∆Hrxn off or absorbed (state which one)?

30. You mix 50 mL of 0.2-M HCl and 50 mL of 0.2-M NaOH


in a calorimeter. You measure a temperature increase of 33. Consider the combustion of methane
1.1ºC. Assuming that the speci c heat and the density of the CH4(g) + 2O2(g) ➞ CO2(g) + 2H2O(l) ∆Hrxn = –891 k
solution is the same as those of water (4.184J/gK, 1g/mL),
If we want to produce 1 MJ (megajoule = 106J) of heat, how
let’s nd ∆Hrxn
many grams of methane do we need to react? (See Example
HCl(aq) + NaOH(aq) ➞ H2O(l) + NaCl(aq) 9 for help.
a. First, what is the mass of the solution that results

b. Now nd qrxn

34. Now consider the dissolution of NaCl


c. How many moles of HCl and NaOH did we use NaCl(s) ➞ Na+(aq) + Cl–(aq) ∆Hrxn = +4 k
If 117 g of NaCl dissolve in 1000 mL of water, what will be
the change in temperature? (See Example 10 for help.

d. Find what qrxn would have been if we had used 1 mole (as
indicated by the balanced equation)

e. So, what is ∆Hrxn

Chapter 10 Thermodynamics 397


fi
fi
)

fi

10.4 Finding ΔH by Hess’s Law 38. Determine the enthalpy of reaction for
35. Determine the enthalpy of reaction fo 2C4H10(g) + 13O2(g) ➞ 8CO2(g) + 10H2O(l)
2AC + 2B ➞ 2AB + C2 given the following information
given the following information C(gr) + O2(g) ➞ CO2(g) ∆H = –393.5 k
A+B➞A ∆H1 = +100 k H2(g) + ½O2(g) ➞ H2O(l) ∆H = –285.5 k
2A + C2 ➞ 2A ∆H2 = –200 k 4C(gr) + 5H2(g) ➞ C4H10(g) ∆H = +20.6 k

36. Determine the ∆Hrxn fo


P4 + 6Cl2 ➞ 4PCl3
given the following information
4PCl5 ➞ P4 + 10Cl2 ∆H = +1774 k
PCl5 ➞ PCl3 + Cl2 ∆H = +124 k 39. (Tougher one!) Determine the heat of reaction for
NO + O → NO2
given the following information:
NO + O3 → NO2 + O2 ∆H = –198.9 kJ 

O3 → 3/2 O2 ∆H = –142.3 k
O2 → 2O ∆H = +495.0 kJ

So if you react 50 grams of P4, how much energy will be


released (or gained)

10.5 Finding ΔH by Standard Enthalpies of Formation


40. Which of these have ∆Hfº=0? In other words, which of
these would you write on the reactant side of a balanced
37. Determine the heat of reaction for equation for the standard enthalpy of formation of a
2C(gr) + 3H2(g) ➞ C2H6(g) compound? (Try not to use your table.
a. F2(g) b. Mg(s) c. CH4(g) d. O2(l) e. He(g)
given the following information
f. H2(g) g. Fe(g) h. H2O(l) i. Na+(aq) j. Fe(s)
C(gr) + O2(g) ➞ CO2(g) ∆H = –393.5 k
H2(g) + ½O2(g) ➞ H2O(l) ∆H = –285.8 k 41. Write balanced equations for the standard enthalpy of
2C2H6(g) + 7O2(g) ➞ 4CO2(g) + 6H2O(l) ∆H = –3119.6 k formation of
a. CBr4(s)

b. Mg(OH)2(s)

c. IBr5(l)

d. C6H7N(l)

42. Use your table to nd the ∆Hºrxn of the following at 25ºC


4FeO(s) + O2(g) ➞ 2Fe2O3(s)

398 Chapter 10 Thermodynamics


B
?

fi
C

r

43. Find the ∆Hºrxn of the following reaction at 25ºC 47. Use the table of bond energies to nd ΔHrxn for this
2CH3OH(l) + 3O2(g) ➞ 2CO2(g) + 4H2O(l reaction
CH4(g) + 3 F2(g) ➞ CHF3(g) + 3 HF(g)

H H

H C H + 3 F F ➞ F C F + 3 H F

H F
44. Calculate ∆Hºrxn for the combustion of C2H4(g). (Stuck?
Write out a chemical equation for the combustion of C2H4(g).

45. Here’s a tough one. We want to calculate the standard


enthalpy of formation of CS2(l)
10.7 Spontaneity and Entropy
C(gr) + 2S(s) ➞ CS2(l) ∆Hfº =
48. Predict whether the entropy increases or decreases
given the following information a. a beaker of water evaporate
a. CS2(l) + 3O2(g) ➞ CO2(g) + 2SO2(g) ∆Hºrxn = –1072 k INCREASES DECREASE
b. ∆Hfº (CO2) = –393.5 kJ/mo b. sugar dissolves in wate
c. ∆Hfº (SO2) = –296.1 kJ/mo INCREASES DECREASE
Here’s some help c. gas escaping from a test tube into the classroo
If you were going to nd the ∆Hºrxn for reaction (a) above, what INCREASES DECREASE
would be the equation – in terms of the ∆Hfº‘s of the reactants and d. rain turning into snow ake
products – that you would write to nd it? The only one you’re INCREASES DECREASE
missing is ∆Hfº (CS2). But you know that products minus reactants
equals –1072 kJ. Now solve 49. Predict the sign of ∆S
a. dry ice sublimes (turns into gas ∆S>0 ∆S<0
b. water in the freezer solidi e ∆S>0 ∆S<
c. air escaping from a leak in a tir ∆S>0 ∆S<

50. Predict the sign of ∆S.


a. NH4Cl(s) ➞ NH3(g) + HCl(g) ∆S>0 ∆S<
b. 2 H2(g) + O2(g) ➞ 2 H2O(g) ∆S>0 ∆S<
c. NaCl(s) ➞ Na+(aq) + Cl–(aq) ∆S>0 ∆S<
d. H2(g) + Br2(g) ➞ 2 HBr(l) ∆S>0 ∆S<

10.6 Finding ΔH by Bond Energies 51. Predict the sign of ∆H and ∆S


46. Use the table of bond energies to nd ΔHrxn for this a. An exothermic reaction where energy is dispersed
reaction ∆H>0 ∆H<0 ∆S>0 ∆S<
CO2(g) + 2 H2(g) ➞ 2H2O(g) + C(gr) b. An endothermic reaction where energy is dispersed.
To help, here are the types of bonds. Note some are double ∆H>0 ∆H<0 ∆S>0 ∆S<
c. An exothermic reaction where energy becomes localized
O=C=O + 2 H–H ➞ 2 H–O–H +
∆H>0 ∆H<0 ∆S>0 ∆S<
d. An endothermic reaction where energy becomes localize
∆H>0 ∆H<0 ∆S>0 ∆S<

52. Predict whether ∆S is positive, negative, or you can’t tell


a. CaO(s) + CO2(g) ➞ CaCO3(s) ∆S>0 ∆S<0 can’t tel
b. NH3(g) ➞ NH3(aq) ∆S>0 ∆S<0 can’t tell
c. 2 NH3(g) ➞ N2(g) + 3 H2(g) ∆S>0 ∆S<0 can’t tell
d. N2(g) + O2(g) ➞ 2 NO(g) ∆S>0 ∆S<0 can’t tel

Chapter 10 Thermodynamics 399









:

fi
fl
.

fi
s

s
0

e
.

)
fi

fi
fi
0

10.8 Spontaneity and Temperature 10.9 Calculating Entropy Changes Numerically


53. Given the following information, determine if the reaction 59. Consider the following reaction
is going to be (1) spontaneous at all temperatures, (2) non- Ti(s) + 2Cl2(g) ➞ TiCl4(g)
spontaneous at all temperatures, (3) spontaneous only at
a. Calculate the ∆Sºrxn
high temperatures, or (4) spontaneous only at low
temperatures
a. An exothermic reaction where entropy increases
Spont at All T Non-Spont at All T
Spont at Hi T Spont at Low
b. An endothermic reaction where entropy increases
Spont at All T Non-Spont at All T
Spont at Hi T Spont at Low
c. An exothermic reaction where entropy decreases
Spont at All T Non-Spont at All T
Spont at Hi T Spont at Low b. Does entropy favor spontaneity or non-spontaneity

d. An endothermic reaction where entropy decreases


Spont at All T Non-Spont at All T
Spont at Hi T Spont at Low c. Let’s say that the reaction is endothermic. The reaction wil
i. be spontaneous at all temperature
54. Same directions as Problem 53 ii. be non-spontaneous at all temperature
a. 2 HgO(s) ➞ 2 Hg(l) + O2(g) ∆H = +90.7 k iii. be spontaneous at high temperature
Spont at All T Non-Spont at All T iv. be spontaneous at low temperature
Spont at Hi T Spont at Low
b. 2 CO(g) + O2(g) ➞ 2 CO2(g) ∆H = –566 k
Spont at All T Non-Spont at All T
10.10 Gibbs Free Energy
Spont at Hi T Spont at Low 60. A reaction has a ∆Gº = +173 kJ.
c. 2 C(gr) + O2(g) ➞ 2 CO(g) ∆H = –221 k a. Is it endergonic or exergonic
b. Is it spontaneous or non-spontaneous
Spont at All T Non-Spont at All T
Spont at Hi T Spont at Low
d. 2 C(gr) + 2 H2(g) ➞ C2H4(g) ∆H = +52.3 k
61. If a reaction is ENDERGONIC/EXERGONIC, the forward
Spont at All T Non-Spont at All T
reaction is spontaneous and ∆Gº is POSITIVE/NEGATIVE
Spont at Hi T Spont at Low

55. The following reaction is currently non-spontaneous.


62. Given the following information, determine if each of the
Should we raise or should we lower the temperature to make
following will be spontaneous or non-spontaneous
it spontaneous
a. ∆Hº = –27.6 kJ, ∆Sº = –55.2 J/K, T = 262ºC
AB(g) + 2 C(g) ➞ AC(s) + BC(s) ∆Hrxn = –100 k

56. The following reaction is not spontaneous


A2(s) + B2(s) ➞ 2 AB(g)
Do you think it is endothermic or exothermic

57. At a certain temperature, this reaction is spontaneous. b. ∆Hº = +145 kJ, ∆Sº = +322 J/K, T = 482
What physical state do you think product X must be in
Z(g) + X2Z(s) ➞ 2 X(?) + Z2(s) ∆H = +100 k

58. At a certain temperature, this reaction is spontaneous. Is


it endothermic or exothermic
DM(g) + X(g) ➞ DMX(s)

400 Chapter 10 Thermodynamics


63. Determine if the reaction is spontaneous by calculating 66. Determine for each of the following whether the forward
∆Gº at 473K (Hint: are you allowed to use ∆Gfº’s from the or reverse reaction is spontaneous by calculating ∆Gº at
table at this temperature?) 25ºC. (Hint: does this temperature give you a clue to how
2 N2O(g) + 3 O2(g) ➞ 4 NO2(g) you can solve this more quickly than in #63 and #65?
a. CH4(g) + 2 O2(g) ➞ CO2(g) + 2 H2O(l)

b. 2 CO2(g) ➞ 2 CO(g) + O2(g)

64. Using the ∆Hº and ∆Sº you calculated in #63, determine
the temperature at which that reaction will be spontaneous
(make sure you say “below” or “above” that temperature)

67. Given the following thermodynamic data for a reaction,


determine the temperature at which it will be spontaneous.
Make sure you specify if it is over or under that temperature
∆Hº = +284 kJ ∆Sº = +344 J/K

65. Determine if the reaction is spontaneous by calculating


∆Gº at 1000K
6 Fe2O3(s) ➞ 4 Fe3O4(s) + O2(g)

Will it be spontaneous at 2000K? (Note: you don’t need to


nd ∆Hº and ∆Sº again. Just reuse them at this new T.

Chapter 10 Thermodynamics 401


fi
.

You might also like