Unit - 9 - CH 123456

Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

9.

1 Introduction to Entropy

What is Entropy?

In unit 5 you were introduced to the world of thermochemistry. That is, applying the rules of thermodynamics to our
study of chemistry! During this unit, you learned about one key measure of energy: heat or enthalpy. In unit 9, our
view of thermodynamics will be expanded past simply heat to measure reactions in terms of two other measures:
entropy and Gibbs’ Free Energy. In doing so, we’ll explore a reaction’s spontaneity. In simple terms, a spontaneous
process will occur without an outside intervention like adding energy to the system. For example, a ball rolling down a
hill is a spontaneous process because it occurs without needing anything to happen. However, the reverse process,
rolling the ball back up, is non-spontaneous. It takes work to roll a ball up a hill. In this section, we’ll dive into the idea
of entropy.

Entropy can be thought of as a measure of “chaos” in a system.


Essentially, entropy answers the question of “how ordered is this
system?” Entropy is also described as the number of possible
arrangements in a system. Essentially, the more spread out and
chaotic the system is, the higher the entropy.

Entropy is important to chemists because it helps them


understand energy flows in terms of creating order or disorder,
especially when discussing reversible processes. To understand
this, you can think about a practical example: your bedroom can
get messy pretty easily. You can ruin the sheets, punch holes in
the wall, throw clothes everywhere, get crazy. This process is a
process in which entropy increases. The room went
from ordered to disordered meaning entropy increased. You can
also use your energy to clean your room. This would mean your
room would go from disordered to ordered and the entropy of
your room would decrease. This situation explains how systems
tend towards chaos—it didn’t take energy to ruin your room but it did take energy to clean it. This is part of the Second
Law of Thermodynamics which we’ll discuss in more detail later in this guide.

Entropy and States of Matter

There are meaningful connections between entropy and states of matter as well. In general, solids have the least amount
of disorder, followed by liquids, and finally, gases have the most entropy. This makes sense when you think about the
properties of solids, liquids, and gasses. Solids are tightly packed in and have little movement between the pieces of the
solid (either atoms or molecules). Therefore there is the least amount of entropy. Moving to liquids, we see more
movement in the molecules and more disorder. Finally, gases are extremely chaotic. Remember that in a gas the
molecules theoretically are wildly flying around. Compare this to a solid where there is not much movement at all.

We can represent state changes in the following way:

X (s) ⇌ X (l) ⇌ X (g)

Moving forward along this chain increases entropy and moving backward decreases entropy.
The Three Laws of Thermodynamics

Law #1: The Law of Conservation of Energy

You’re probably already familiar with the First Law of Thermodynamics. This law is also known by a more common
name: the Law of Conservation of Energy. This tells us that energy can never be created or destroyed but rather can
only change form. For example, energy can change from potential energy to kinetic energy, but the energy between the
system and the surroundings is constant. Anything lost to the system is gained by the surroundings and vice versa.
From this we also know that the total energy of an isolated system remains constant.

Law #2: Energy Quality and Entropy Increases In Isolated Systems

As far as our discussion of entropy goes, the second law of thermodynamics is perhaps the most important. There are
two main parts of this law.

The first part of the second law of thermodynamics refers to energy quality. It tells us that as energy is converted from
form to form and/or transfers from body to body, some of it is lost to the surroundings as heat. For example, suppose a
power plant runs a turbine to create electricity. In that case, some of the mechanical energy of the turbine is converted
to electrical energy while some of it is lost as heat energy via friction. This same idea is applicable to any energy usage.
The second part of the second law of thermodynamics refers to
entropy changes in an isolated system. The second law of
thermodynamics tells us that in an isolated system, entropy will tend
to either increase or stay constant. The second law of
thermodynamics states that for any spontaneous process, the overall
ΔS must be greater than or equal to zero.

Law #3: Absolute Zero


The third law of thermodynamics is the least important for unit 9 but is still incredibly important. This law tells us that
at absolute zero (0K = -273.15°C), entropy is a constant zero. This is because at absolute zero, all molecular motion
stops and there is no disorder. The formal definition of this law is that “the entropy of a system approaches a constant
value as its temperature approaches absolute zero.”

9.2 Absolute Entropy and Entropy Change

In the last section, we introduced the idea of entropy as a measure of disorder in a system. Unlike enthalpy, which was discussed
in unit 5, entropy can be measured both absolutely and as a change. This differs from enthalpy, which can only be represented as a
change (we can never find H° for a reaction but we can find ΔH°).

Comparing S and ΔS
Entropy can be thought of both in its absolute form (S°) and as representing a change in entropy (ΔS°) (note that the ° symbol in
each simply means that we are at standard conditions, that is 1atm of pressure and 273K). Absolute entropies, also known as
standard entropies, describe the number of possible states a molecule can take, a measure of its disorder. Calculating these values is
incredibly complex, but you will be given any absolute entropies you need on the exam. Similarly, most if not all chemistry
textbooks have tables of thermodynamic data in the back of the book with a chart of standard entropies. These tables may also
contain other data such as enthalpies of formation like the one below. For now, lets look at the far right column that says S°.

Image From Mr. Bigler

Change in entropy (ΔS°) can be thought of the same way we thought about changes in enthalpy (ΔH°) in unit 6 just thinking about
entropy instead of heat changes. In Unit 6, we learned about Hess’s Law, which told us that enthalpy was a state function, meaning
enthalpy changes are pathway independent. The same is true about entropy.
State Functions
Let’s take a quick side note to explain what a state function means and why it’s important. In essence, a state function is a function
that has the property of pathway independence. This means that whatever “path” you take to get to the end result, the end result
will be exactly the same. An example of a state function is the change in altitude when climbing a mountain, whereas an example
of a non-state function is the distance traveled. Since no matter which way you go, you will end up at the top of the mountain, your
change in altitude will always be the altitude. Whether you went straight up, zigzagged, went curvily, or any other pathway, your
change in altitude will be the same at the end. This makes this function pathway independent. On the other hand, the distance you
travel up the mountain does depend on the path you take. If you go straight up, the distance will differ from if you zigzag and for
any other path you take. Therefore, the distance traveled is not a state function and is not pathway independent.

Applying the idea of a state function to entropy, we can logically find the following equation (note that it is strikingly similar to an
equivalent statement about enthalpy change):

This equation tells us that for a reaction the overall entropy change is the sum of the changes in the entropies of the products minus
the sum of the entropies in the reactants. This can be applied the same way we did in unit 5 with enthalpy, so you can simply use
the table and plug in numbers for each product/reactant.

� Pro Tip: Remember your stoichiometric coefficients! The real equation has you multiplying by the stoichiometric coefficient of
each product/reactant but the CollegeBoard omits it from the
formula for some unknown reason.

Interpreting The Sign of ΔS


Another important way of looking at entropy is by observing the
sign of ΔS° for a reaction and drawing conclusions as to whether
or not the reaction became more or less ordered and vice versa—
predicting the sign of ΔS° by looking at properties of a reaction.
When ΔS° is positive, a reaction creates disorder and thus
increases the entropy of the system. Conversely, when ΔS° is
negative, the reaction reduces disorder or in other words creates
order. Therefore the entropy of the system decreases.

By using these rules we can also predict the opposite: the sign of ΔS° of a reaction given the reaction. We’ll take a look at a few of
these in the practice problems.

Practice Problems
Image From Abigail Giordano

In this problem, we’re given a reaction along with the entropies of various substances at 298K. For this problem we can simply use
our equation for S° to find the ΔS° for the reaction:

= ((175) + 2(150)) - (250 + 2(125)) = -25 J/molrxnK ⇒ A.

Note that we multiplied 150 and 125 by 2 because the stoichiometric coefficients on KI and KNO3 are 2.

2. Consider the reaction: 2Na (s) + Cl2 (g) → 2NaCl (s). Predict the sign of ΔS° for this reaction and explain

Looking at this reaction, we see two reactants, sodium and chlorine, forming one solid product, sodium chloride. Let’s think about
the disorder of the reactants versus the disorder of the products. On the reactants side we have 3 total moles (2 + 1 = 3) of reactants,
one of which is a gas. These react to form fewer moles of a solid (less chaotic) product. Therefore, our reaction is more chaotic at
the beginning and becomes less chaotic. This means in the process we “lose” entropy (we don’t actually destroy entropy because
of the first law of thermodynamics, but we can think of it this way). Therefore, we can predict that ΔS° for this reaction is negative

9.3 Gibbs Free Energy and Thermodynamic Favorability

Explaining Thermodynamic Favorability


Thermodynamic favorability is used to predict whether a reaction (or any process for that matter, but for our purposes, a chemical
reaction) is either spontaneous or nonspontaneous. As we discussed earlier in this unit, a spontaneous process is one that occurs
without external inputs, whereas a nonspontaneous reaction requires such external inputs in order to occur.

Spontaneous reactions are considered thermodynamically favorable and non-spontaneous reactions are
considered thermodynamically unfavorable.

As we’ll see in section 9.5, spontaneity has intimate connections with the equilibrium constant for the process in question. By
measuring a reaction’s spontaneity, we can draw predictions about whether the reaction is product-favored or reactant-favored.
Enthalpy and Entropy
There are two main measures that play into spontaneity:

First, we have enthalpy change (ΔH°), the change in the heat of the system relative to the surroundings due to a reaction. If ΔH° is
negative, we know our system is losing heat to the surroundings and is thus exothermic. If ΔH° is positive, we know the
opposite—the system is gaining heat from the surroundings and is therefore endothermic. Knowing if a reaction is exothermic or
endothermic is the first piece of the puzzle.

Next, we have entropy change (ΔS°), the change in the disorder in the system. If ΔS° is positive, our entropy is increasing meaning
the disorder of our system is also increasing. Conversely, if ΔS° is negative, we are “losing” disorder and gaining order meaning
entropy is decreasing. We call reactions that increase entropy exentropic and reactions that decrease entropy endentropic. These
terms are pretty uncommon and most likely will not show up on your exam, but they’re nice to know for your future chemistry
career.

By looking at the change in enthalpy and the change in entropy as a result of a reaction, we can quantitatively find out whether or
not the reaction is spontaneous by introducing a new term:
Gibbs Free Energy.

Gibbs Free Energy


Gibbs Free Energy is a new thermodynamic value that is
used to determine the spontaneity of a reaction. The formula
for Gibbs Free Energy is written in terms of ΔH° and ΔS°
along with temperature: ΔG° = ΔH° - TΔS°

The equation for Gibbs Free Energy is given to you during the
examination on the AP Chemistry reference sheet.

In simple terms, Gibbs Free Energy calculates the amount of


energy available in a system to do work. Therefore,
calculating ΔG° tells us the amount of energy either released from a reaction (if it’s negative) or absorbed (positive). From here, we
can conclude whether a reaction is exergonic (energy releasing) or endergonic (energy absorbing) and therefore
either spontaneous or nonspontaneous.

The formula for ΔG° can be seen as a combination of our two measures of spontaneity. We see enthalpy and entropy combined in
one equation meaning both factor into thermodynamic favorability. We’ll see examples of reactions that are endothermic but
spontaneous and vice versa because of this reason along with the equivalent statements regarding entropy. The following rules can
be applied when calculating ΔG°, but remember to logically understand the rules as well:

 If ΔG° is negative, the reaction is exergonic and therefore spontaneous and thermodynamically favorable.

 If ΔG° is positive, the reaction is endergonic and therefore non-spontaneous and thermodynamically unfavorable.

Practice Problem
Let’s take a look at a practice problem:

Consider the following reaction: 2H2 + N2 ⇌ N2H4. If at 25°C ΔH° = 50.6 kJ/mol and ΔS° = -0.332 kJ/(molK), determine the
value of ΔG° for the reaction and conclude whether it is thermodynamically favorable.

In this question, we are asked to apply the formula from above to calculate Gibbs Free Energy change and then use the sign of ΔG°
to find the spontaneity of the reaction. Plugging into ΔG° = ΔH° - TΔS°, we find the following:
ΔG° = (50.6) - T(-0.332)

Remember that T has to be in Kelvin, so we can add 273 to 25 to find that T= 298 (this is simply converting from C to K, you
should memorize this conversion but it will be on your formula sheet).

Therefore ΔG° = (50.6) - 298(-0.332) = 149.5 kJ/mol.

Because ΔG° is positive, we know that this reaction is nonspontaneous and therefore thermodynamically unfavorable.

We can also calculate ΔG° similarly to the ways we’ve calculated ΔH° and ΔS°, by using standard free energies of formation for
various reactants and products as follows:

Image Courtesy of College Board

Conditions For ΔG To Be Positive or Negative


For a reaction to be spontaneous, it must do one of the following: it must be exothermic or increase the entropy of the system. A
reaction can have one or more of these qualities and be nonspontaneous, but a reaction that has neither will 100% be
nonspontaneous. We can logically see this by looking at the equation for ΔG°: ΔG° = ΔH° - TΔS°

In order for a reaction to be spontaneous, ΔG° must be less than zero. If a reaction is both endothermic (ΔH° > 0) and decreases
entropy (ΔS° < 0) there is no chance of this happening because a positive minus a negative will always be positive.

However, if one of the two is true, the reaction could be spontaneous or nonspontaneous depending on the temperature. If the
reaction is exothermic (ΔH° < 0) but does not increase entropy (ΔS° < 0), it will be spontaneous at low temperatures. This is
because at high temperatures -TΔS° (which in this case is a positive number) will overtake ΔH° and cause ΔG° to be positive.

Conversely, if the reaction is endothermic (ΔH° > 0) but increases entropy (ΔS° > 0), it will be spontaneous at high temperatures
because -TΔS° must be large enough to overtake the positive ΔH° to make ΔG° negative.

Finally, if both conditions are met, that is if ΔH° < 0 and ΔS° > 0, ΔG° will always be negative regardless of temperature. This is
because a negative minus a negative will always be negative.

The following chart shows these rules more neatly:


Entropy and Enthalpy-Driven Reactions
These rules help us predict how a reaction will act without going through
the calculations of ΔG°. It also may enable us to predict whether a
reaction is entropy-driven or enthalpy driven.

For a spontaneous reaction, we call it entropy-driven if the spontaneity


is mainly driven by an increase in disorder. For example, the dissolution
of NaCl is a spontaneous process but is endothermic. Therefore, the spontaneity must be driven by the entropy increase from the
dissolution of NaCl(s) into Na+ (aq) and Cl- (aq).

The opposite can be said for an enthalpy-driven reaction. This is a reaction that shows a decrease in disorder but is exothermic.

9.4 Thermodynamic and Kinetic Control

This section examines the relationship between thermodynamic favorability and kinetics, specifically the shortcomings of Gibbs
Free Energy in predicting reaction behavior. Despite being spontaneous, many reactions do not occur at an observable rate because
they are under kinetic control. In this guide, we will explore what kinetic control means and why it occurs.

Brief Review of Kinetics


Before getting into the thermodynamics of kinetic control, we need to briefly review topics from Unit 5. If you feel confident with
kinetics, feel free to move past this section.

Recall that kinetics is the study of the rate of a reaction—essentially, how fast a reaction occurs. We measure the rate of a reaction
by measuring the change in the concentration of reactants over time: R = Δ[A]/Δt or R = -d[A]/dt (for those familiar with
calculus). Higher R values indicate a quicker loss of reactants and formation of products.

Rates can also be described using rate laws, where initial reactant concentration is directly tied to the rate of a reaction as follows:

R = k[A]^n[B]^m

where [A]^n, [B]^m, etc., are various reactants raised to their reaction order (how much of an impact their concentration has on the
rate).

Take a look at the following rate law as an example:

rate = k[A]

These laws help us describe how quickly a reaction will occur, with a higher rate implying a faster reaction overall.

Most important to our study of kinetic control is understanding activation energy. Due to the kinetic molecular theory, chemical
reactions occur when molecules hit each other at the right angle and speed/energy. This activation energy is the energy required for
a chemical reaction to actually occur. The higher the activation energy, the harder it is for the reaction to occur at an observable
rate.

The following diagram visually shows the concept of activation energy:


A Shortcoming of Gibbs Free Energy
A common misconception when looking at thermodynamic favorability is that a thermodynamically favorable reaction occurs
quickly. Many reactions that are spontaneous occur incredibly slowly. A good example of this concept that can be applied to the
real world is the conversion from diamonds into graphite (represented as Cdiamond(s) → Cgraphite(s)).

For this reaction, ΔG° = -3 kJ. This tells us that the reaction for graphite formation occurs spontaneously and does not require the
input of any external energy to occur. However, take a look at the nearest diamond (because people have those around the
house…right?). Is it suddenly morphing into graphite? No! (I hope not. Fiveable is not responsible for diamond graphitification…).

The conversion of diamond into graphite is incredibly slow, as in thousands of years slow. We say that this reaction is in kinetic
control because it is driven by the slowness of the reaction. These types of reactions are often slow because they have a high
activation energy.

A spontaneous process may take either the thermodynamically controlled or the kinetic controlled pathway. A kinetically
controlled path like the one above is driven by a high activation energy. A thermodynamically controlled reaction is driven by the
difference in free energy between the products and reactants, the type of reaction we saw in the last section.

Reasons For Kinetic Control


As we mentioned before, the primary reason for a reaction to be under kinetic control is because of a high activation energy.
Because of this, even if the reaction is thermodynamically favorable, it may not continue at a measurable rate. There is a way
around this, however, and that is through the use of a catalyst! Catalysts change the mechanism behind a reaction in order to
decrease the activation energy and make reactions quicker. By reducing the activation energy, the reaction can proceed at a
measurable rate.

For example, the decomposition of hydrogen peroxide usually occurs at an unmeasurable rate. However, when iodide ions are used
as a catalyst, it creates the famous "Elephant’s Toothpaste" reaction, which you may be familiar with.

This example shows us how catalysts can transform a reaction from one at an immeasurable rate to one at a measurable
rate. The concept of catalysis can be applied to kinetic control by noticing that catalysts can help get a reaction out of
kinetic control by lowering the activation energy. For example, if there was a catalyst for the reaction: Cdiamond(s) →
Cgraphite(s), the reaction would be able to proceed at a measurable rate. However, without one, it cannot because the
rate of the reaction is incredibly slow.

9.5 Free Energy and Equilibrium

Kinetic and Thermodynamic


Definitions of Equilibrium
In Unit 7, we spent a ton of time discussing the concept of equilibrium.
Equilibrium helped us determine how a reversible reaction may behave
under certain circumstances. In this section, we will look at how equilibrium connects to spontaneity and ΔG° by observing various
definitions of equilibrium along with relationships between ΔG° (ΔG at standard conditions) and ΔG (ΔG at any other conditions).

When learning about equilibrium, the typical definition is the kinetic definition of equilibrium. This definition concerns the rates
of the forward and reverse reactions as the reaction approaches equilibrium. Recall that we defined equilibrium as the point at
which the forward reaction and reverse reaction proceeded at the same rate, meaning the concentrations of products and reactants
stay the same. It is important to note that equilibrium does not mean
nothing is happening, just that both reactions are happening at the
same rate, so they “cancel” each other out.

However, there is another important definition of equilibrium, which


concerns thermodynamics (the thermodynamic definition). The
thermodynamic definition defines equilibrium as the point of
minimum free energy. While the reaction occurs spontaneously, ΔG
(note the lack of the naught symbol: the ° next to G) will be less than
zero, meaning the reaction will be releasing free energy. After it
reaches equilibrium concentrations, ΔG will be positive and, thus,
requires external sources of energy to occur.

We can see the point of minimum free energy visually in the following graphs, in which the y-axis is free energy (G), and the x-
axis is the extent of reaction from 100% reactants to 100% products:

First, note that ΔG is not the y position of the graph but rather the rate of change of the graph. When ΔG < 0, the graph decreases,
and when ΔG > 0, the graph increases.

We begin on the far left with 100% reactants, so ΔG = ΔG° of the reactants. Similarly, on the far right, we have 100% products, so
ΔG = ΔG° of the products. We can see that the ΔG° of the products is lower than the ΔG° of the reactants, indicating that ΔG° for
this particular reaction is negative and defines the reaction as spontaneous: ΔG° = ΣnΔG°f (products) - ΣnΔG°f (reactants).

For this reaction, we begin with a negative ΔG. The reaction is still in a spontaneous “state." Therefore, we will continue producing
products. We move forward until we hit the point at which ΔG=0—the minimum point of free energy known as the equilibrium
point.

From the equilibrium point forward, ΔG > 0. Thus, we need to add energy to the system to produce more products because the
reaction is now nonspontaneous.

The above graph is similar to the graph from before. Except, note that our ΔG° (products) is now greater than ΔG° (reactants).
Essentially, ΔG° is positive for this reaction and nonspontaneous. We are also told that the mole fraction of reactants is large
compared to products at equilibrium, suggesting that we have many more reactants than products. A nonspontaneous reaction is
linked to a reactant factored reaction.
Relationship Between ΔG, ΔG, and K
After observing a qualitative relationship between ΔG°, ΔG, and K, we can look at some of the mathematical equations that are
used to describe this relationship. Remember that ΔG is a measure of free energy change at nonstandard conditions, so a
relationship can be drawn between ΔG° and ΔG by connecting it to Q, the reaction quotient:

To calculate Q, we plug non-equilibrium concentrations or pressures in the law of mass action (our equilibrium formula). R is the
gas constant (8.314 J/molK or equivalent units), and T is the temperature in Kelvin.

Next, we examine the relationship between ΔG° and K. Using our above equation, we can derive a relationship directly with only
ΔG°. At equilibrium, the two main conditions are ΔG = 0 and Q = K. Substituting into our equation, we find the following:

0 = ΔG° + RTln(K)

ΔG° = -RTln(K)

Solving for K:

-ΔG° = RTln(K)

-ΔG°/RT = ln(K)

K = e^(-ΔG°/RT)

These two equations show a direct relationship between ΔG° and K. Qualitatively, we see that a higher ΔG° means a lower K and
vice versa. With a positive ΔG°, we see that -ΔG°/RT is negative. Therefore, K is e^(negative number), making it less than 1.
Similarly, with a negative ΔG°, -ΔG°/RT is positive, so K = e^(positive number) and greater than 1.

You might also like