Vous êtes sur la page 1sur 277

CHEMISTRY 102/105

INTRODUCTORY UNIVERSITY
CHEMISTRY II
¾ This course focuses upon many of the
essential concepts in physical
chemistry, which is the branch of
chemistry concerned with the application of
the laws of physics to chemical phenomena.

¾ We will begin this term by examining the


fundamental underpinnings of chemical
equilibrium.
Chemical Equilibrium

¾Chemical equilibrium is the state of


chemical balance. By definition, equilibrium
is reached when the forward and reverse
reactions proceed at the same rate.

¾One of the consequences of attaining


equilibrium is that the quantities of reactants
and products remain fixed with time.
¾ Consider, for instance, the reaction
H2(g) + I2(g) 2HI(g)
beyond the time te on the graph below, the concentrations for all three
species in the equilibrium remain constant. Thus, we can say that
equilibrium was achieved at time te
¾ It is very important to stress at this point that, contrary to what some
of you may have been told in the past, NO REACTION GOES 100%
TO COMPLETION!

¾ ALL REACTIONS should have two half arrows pointing in opposite


directions separating the reactants from the products to indicate that
the equilibrium mixture invariably contains both reactants and
products, and that the equilibrium is dynamic.

¾ The ratio involving the product of the equilibrium concentrations for


the products raised to their respective coefficients divided by the
product of the equilibrium concentrations for the reactants raised to
their respective coefficients has a single value, at a fixed temperature.

¾ This ratio is called the equilibrium constant expression. It is based


upon the Law of Mass Action and its form is determined by the
stoichiometry for the reaction.
¾ i.e. for the generic reaction…

aA + bB cC + dD

the equilibrium constant expression is written as

Kc = [C]c×[D]d
[A]a ×[B]b

where the c in Kc represents concentration in


moles per liter and the square brackets are equilibrium
concentrations in moles per liter
¾If partial pressures are used instead of
concentrations, as is the case for gaseous
systems, the expression becomes

PCc × PD d
KP = ,
PA a
× PB b

where P represents equilibrium partial


pressure
¾ To sum up, when reactants are mixed, they
immediately begin to form products, which leads
to a decrease in the concentration of the
reactants and an increase in the concentration
of the products.

¾ Beyond a certain period of time, the


concentrations for the reactants and products no
longer change-equilibrium has be reached,
meaning the rates for the forward and reverse
reactions have become equal. The value unique
to the temperature at which the reaction
occurred can then be found by inserting the
equilibrium concentrations into the equilibrium
constant expression and solving the mass action
ratio specific to the reaction.
Movement to Equilibrium and Equilibrium
Positions

Consider, for instance, the NO2 - N2O4


equilibrium:


N2O4(g) 2NO2(g)
2
where Kc = [NO 2 ]
[N 2O 4 ]
Calculate Kc for the two following experiments (assume T is the same
for both experiments)

Experiment 1
Initial Concentrations Equilibrium Concentrations
[N2O4] = 1.0 M [N2O4] = 0.80 M
[NO2] = 0.0 M [NO2] = 0.40 M

Kc = [NO2 ]2 = [0.40 M]2 = 0.20 M


[N 2O4 ] [0.80 M]
Experiment 2
Initial Concentrations Equilibrium Concentrations
[N2O4] = 0.0 M [N2O4] = 0.36 M
[NO2] = 1.0 M [NO2] = 0.27 M

[NO2 ]2 [0.27 M]2


Kc = = = 0.20 M
[N 2O4 ] [0.36 M]
¾ Therefore, the final equilibrium positions for Experiments 1 and 2 are
quite different but the Kc is the same!(as it should be)

Gaseous Equilibria
¾ When the system under study involves only gases, it is customary to
use equilibrium partial pressures, which are proportional to
concentrations, in place of equilibrium concentrations (pressures of
gaseous mixtures are easily measured).

e.g. for the reaction

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

[NH3 ] 2 ( PNH3 ) 2
Kc = and KP = 3
(preferred)
[N 2 ][H 2 ]3
( PN2 )(PH 2 )
¾ The numerical values for KP and Kc are related by the expression

Δn g
K P = K c (RT)

where ∆ng = ∑ no. of moles of gaseous products - ∑ no. of moles of


gaseous reactants ;

L • atm
R = 0.08206 ; T = Temperature in Kelvins
moles • K
Note: When ∆ng = 0, KP = Kc

Upshot of this section: If the reaction involves only gases, the


equilibrium constant should be expressed in terms of equilibrium partial
pressures (KP).
Relationship of the Equilibrium Constant to the
Balanced Equation

1. If the equation is reversed,


1
K reversed =
K orginal

2. If the entire balanced equation is multiplied through by the


value n, the K value for the new equation is given by

n
K new = ( K original )
Heterogeneous Equilibria

¾ For equilibria that involve pure liquids or solids, the


concentration of the pure liquid or solid is excluded from
the equilibrium constant expression. To be more precise,
pure liquids or solids are assigned a value of one with no
units in the K expression.

¾ The reason this is done is because pure liquids and


solids have constant concentrations (i.e. constant
densities) at a fixed temperatures and thus they do not
change during the course of the reaction. These constant
concentration values don’t actually “disappear” but rather
are submerged mathematically into the equilibrium
constant expression. In other words, the concentrations
for pure liquids and solids actually form part of the K
value itself.
e.g. For the reaction

CaCO3(s)
CaO(s) + CO2(g)

Kprototype = [CO 2 ][CaO] = K proto × [CaCO3 ] = [CO2 ] × 1 = Kc


[CaCO3 ] [CaO] 1

Kc = [CO 2 ] × 1 ; Therefore, Kc = [CO2]


1

and KP = PCO2
¾ Note that the equilibrium concentration of CO2 is not
affected by the amount of CaO and CaCO3 present.
However, at least a small quantity of both of these solids
must be present. Without both solids being in contact with
gaseous CO2, the reaction cannot occur and, clearly, the
system, as a result, will not be at equilibrium.

Quantitatively Interpreting the Equilibrium Constant

¾ The size of the equilibrium constant indicates the extent of


reaction.
¾ If the equilibrium constant is large (much greater than
one), one can immediately conclude that products are
strongly favoured at equilibrium.
e.g. N2(g) + 3H2(g)
2NH3(g) has
Kc = 4.1108 × 108 M-2 at 25oC.

¾ Thus, the number of product molecules is vastly greater


than the number of reactant molecules at equilibrium. (i.e.
the reaction strongly favours the formation of ammonia at
25oC.)

¾ If the equilibrium constant is small (much smaller than


one), the reactants are strongly favoured at equilibrium.

e.g. N2(g) + O2(g)


2NO(g) has
Kc = 4.610 × 10-31 at 25oC.
¾ Since the equilibrium constant is exceedingly small, the
concentration for the product is essentially undetectable,
but not equal to zero!(i.e. at 25oC, the reaction very
strongly favours reactants at equilibrium).

¾ When the equilibrium constant is neither large nor small


(between ca. 0.1 and 10), the equilibrium mixture contains
significant amounts of all substances in the reaction (i.e.
neither reactants nor products predominate).


e.g. CH3COOH(aq) + CH3OH(aq) CH3COOCH3(aq) + H2O(l)
has Kc = 2.0 M-1 at 25oC.
Thus, the equilibrium mixture at 25oC contains a “dog’s
breakfast” of roughly equal amounts of CH3COOH(aq) ,
CH3OH(aq) and CH3COOCH3(aq).
Effect of a Catalyst

¾ Catalysts, as we saw last term, are substances


that increase the rate of a reaction by providing
an alternate pathway that possesses a lower
activation energy.

¾ Catalysts have no effect upon the equilibrium


composition for a reaction mixture. They merely
speed up the attainment of equilibrium.
Predicting the Direction of a Reaction to
Equilibrium

¾ A given reaction mixture is not necessarily at


equilibrium.

¾ The direction of the reaction to equilibrium can


be determined by calculating the reaction
quotient (Q) and comparing it to the K for the
reaction.

¾ The reaction quotient, Q, has the same format as


the K, but its calculation involves the use of initial
concentrations or partial pressures. Therefore, Q
is not a constant, unless of course, the Q is equal
to K and the system is already at equilibrium.
[Product C]cinital × [Product D]d initial
Reaction quotient: Q =
[Reactant A]a initial × [Reactant B]b initial

For the reaction aA + bB


cC + dD

Three Scenarios When Q is compared to K


1. If Q < K, the reaction will proceed to the right, in the direction of
products, to reach equilibrium. In other words, the ratio of
products to reactants is too small and, to establish equilibrium,
reactants must be converted into products to push the Q up to K.

2. If Q > K, the reaction will proceed to the left, in the direction of


reactants, to reach equilibrium. In other words, the ratio of
products to reactants is too large and, to establish equilibrium,
products must be converted into reactants to push the Q down to
K.

3. If Q = K, the reaction mixture is at equilibrium and thus no net


reaction in either direction will occur.
Sample Problem:
Gaseous methanol can be produced by the reaction

CO(g) + 2H2(g)

CH3OH(g) KP = 10.5 atm-2
at 500 K

A gaseous mixture at 500K is 0.020 atm in CH3OH,


0.10 atm in CO and 0.10 atm in H2. Determine whether
the system is at equilibrium and, if not, in which
direction the net reaction will proceed to reach
equilibrium.
Le Chatelier’s Principle

¾ This important equilibrium concept can be stated


as follows:

“ If a system at equilibrium is subjected to a


stress (e.g. a change in temperature, changes in
the pressures of gaseous reactants or products,
a change in the concentration of reactants or
products), the position of the equilibrium will shift
in the direction that reduces the imposed stress.”
Table of Stresses and the Resulting Equilibrium Shift

Stress Direction of Equilibrium Shift

Addition of reactant or To the right (more product(s) is/are formed)


removal of product

Addition of product or To the left (more reactant(s) is/are formed)


removal of reactant

Increase total gas pressure To the side of the equation with fewer moles
by decreasing volume of gas (to decrease the pressure)

Decrease total gas pressure To the side of the equation with more moles
by increasing volume of gas (to increase the pressure)

Increase the temperature Endothermic reaction is favoured (K value


for the endothermic reaction gets bigger)

Decrease the temperature Exothermic reaction is favoured (K value


for the exothermic reaction gets bigger)
Note: Of all of the above stresses, only a change in temperature
changes the value of K.

K is larger at higher T if the forward reaction is endothermic while


K is larger at lower T if the forward reaction is exothermic

Sample Problem:
For the reaction
N2(g) + 3H2(g) 2NH (g)
3 ∆Ho = -92 kJ/mol rxn
at equilibrium in a closed vessel at 773 K and 350 atm pressure, in
which direction will the equilibrium shift in response to each of the
following changes in conditions ?
a) Remove some NH3
b) Add some more N2
c) Decrease the volume of the system while keeping T constant
d) Add a catalyst
e) Increase the total pressure by adding some He gas
f) Increase the temperature
How to Solve Equilibrium Problems: The ICE Table Method

1. Determine whether the reaction is at equilibrium by calculating Q


(the reaction quotient) and comparing it to K (the equilibrium
constant).

2. If the system is not at equilibrium (i.e. Q ≠ K ), set up a table


of initial (I), change (C) and final equilibrium (E) concentrations (or
partial pressures) called an ICE table.

3. Choose an unknown (x) to represent the change in concentration


or partial pressure that a reactant or product must undergo in order
for equilibrium to be achieved. If at all possible, let x = necessary
change in concentration/ partial pressure for a species in the
balanced equation that has a coefficient of one (this will simplify
the solution).

4. Express the changes in concentrations or partial pressures for the


various reactants and products in terms of x and their
stoichiometric coefficients. Signs in front of the terms involving x
indicate the direction of the net reaction to get to equilibrium.
5. For those species that have an initial concentration or partial
pressure greater than zero, in each case, you must divide the
concentration or partial pressure by the equilibrium constant for the
reaction to equilibrium that is taking place. If all (Ci/K) results falls
between 1000 and 0.001, then the complete polynomial must be
solved. In other words, no approximations can be made in these
cases and thus the Intermediate K algorithm must be used to find
the equilibrium concentrations.

6. If all of the (Ci/K) ratios for those species with non-zero initial
concentrations/partial pressures are greater than 1000, then the
reaction is a Small K problem. In situations where K is small, we
can employ what is called an heuristic strategy, which involves
making certain legitimate approximations that make the solution to
the problem more simple. One can avoid solving quadratic
equations or higher order polynomials by ignoring the amount nx
that is added to or subtracted from the concentration or partial
pressure of a reactant or product if the amount nx is less than 5%
of the reactant/ product concentration or partial pressure to which it
is to be added or from which it is to subtracted (This is called the
5% rule). The amount nx will be less than 5% of the initial reactant/
product concentration or partial pressure when the ratio Ci/K
>1000.
7. If all of the (Ci/K) ratios for those species with non-zero initial
concentrations/partial pressures are less than 0.001, then the reaction is
a Large K problem. In situations where K is large, the algebraic
determination of the equilibrium concentrations is made much simpler if
we first take the reaction 100% to completion and then back it up a little
bit to get to the state of equilibrium. It is during the backing-up step that
we introduce the unknown x. Effectively, what we accomplish by first
taking the reaction 100 % to completion in the direction for which K is
large is to convert the system into one that can be solved via the Small K
approach. By introducing x in the back-up step, we ensure that it will be
negligible with respect to the concentration or partial pressure of at least
one species in the balanced equation, which will cause the order of
the polynomial to decrease when we apply the 5% rule. (Final note:
whenever the Ci/K ratio for a reactant or product is greater than 1000,
any associated nx term can be ignored).

8. Check your results by substituting them into the equilibrium


constant expression and solving for K.

9. Finally, if you invoke the 5% rule, you must make certain that the
assumption you made does not introduce more than 5 % error. If the
error created by the assumption exceeds 5 %, the approximation cannot
be made and the complete polynomial must be solved.
The Method of Successive Approximations

¾ In addition to the heuristic strategies described above,


higher order polynomials can be solved by using the
method of successive approximations. This method can
be used whenever the 5% rule fails and the order for the
polynomial is three or higher.

¾ This iterative procedure is a systematic method of trial


and error; the estimated value gets closer and closer to
the correct answer with each successive iteration
(repeated calculation).

¾ A reasonable value for x is guessed and substituted


everywhere x appears, except for one place.
e.g. consider the cubic equation (2 x) 2 (0.25 + x)
= 4.0
(0.25 − x)

Rearranging into standard cubic equation format we get


4x3 + x2 + 4.0x = 1.0

Start with x = 0.20 (x must be less than 0.25)


xo = 0.20 ; 4(0.20)3 + (0.20)2 + 4.0x = 1.0
x1 = 0.232; x2 = 0.224; x3 = 0.226 ; x4 = 0.226

Therefore, x = 0.23 M (reported to two significant figures)

Since we obtained the same answer for two successive


iterations, the solution has converged and we have thus
found the correct value for x.
Intermediate K Problem

The system described by the reaction


CO(g) + Cl2(g)
COCl2(g) ∆Ho = - 75 kJ/mol rxn
is at equilibrium at 100oC when
PCO = 0.30 atm, PCl2 = 0.10 atm and PCOCl2 = 0.60 atm .
An additional 0.40 atm of Cl2 gas is added.

a) Find the final pressure for CO after the system


returns to equilibrium.
b) What effect will an increase in volume have upon the
mole fraction of COCl2 in the mixture?
c) What effect will a decrease in temperature have upon
the mole fraction of COCl2 in the mixture?
Small K Problem
At high temperatures, sulfuryl chloride gas (SO2Cl2(g))
decomposes to a small extent to give molecular chlorine
gas and sulfur dioxide gas. The equilibrium constant for this
reaction is KP = 1.34 × 10-5 atm at 500 K. What is the
equilibrium partial pressure for each gas if initially 1.00 atm
of SO2Cl2 gas is present in a closed container at 500 K?
Large K Problem

At 295 K, the reaction

2NO(g) + Cl2(g) has KP = 1.0 × 10-5 atm



2NOCl(g)

Calculate the partial pressure for each gas at equilibrium if


initially 0.60 atm of NO(g) and 0.70 atm Cl2(g) are present
in a sealed container at 295 K.
Acids and Bases

¾ Acids and bases have widespread use in our society and they play essential
roles in the chemistry of living systems.

Arrhenius Definition

An Acid is a substance that release H+(aq) (or H3O+(aq)) when dissolved in


water while a

Base is a substance that releases OH-(aq) when dissolved in water.

This is a somewhat limited definition as it allows for only one kind of base
(viz. OH-(aq)) and it is restricted to aqueous solution.

¾ A much more versatile model was proposed by Johannes Bronsted and


Thomas Lowry.

¾ This approach, aptly named the Bronsted-Lowry model, an acid is a proton


donor, while a base is a proton acceptor.

¾ Thus, Bronsted-Lowry model emphasizes proton transfer and thus applies to


reactions in other solvents as well as water.
¾ Bronsted bases have at least one pair of electrons that can be used to
form a dative covalent bond with the proton from the Bronsted acid
(Note: a dative covalent bond is one in which just one of the two atoms
supplies both of the bonding electrons)

¾ The products of Bronsted-Lowry acid-base reactions are also acids


and bases. In fact, in these reactions, the acid is converted into its
conjugate base, while the base is converted into its conjugate acid.
Thus, conjugate acid-base pairs differ by one proton (vide infra).
¾ Amphoteric species can behave as either an acid or a base.

¾ These molecules contain a relatively easily ionizable hydrogen atom and at


least one pair of electrons that can be used to form a dative covalent bond.

¾ They act as Bronsted bases towards stronger acids and Bronsted acids
towards stronger bases.
e.g. H2O is an amphoteric substance

Above, water is acting as a Bronsted base towards the stronger acid HA


¾ This time, the water acts as a Bronsted acid towards the stronger
base NH3(aq):
Acid Strength

¾ The strengths of acids and bases are defined by the positions of their
dissociation reactions with water (i.e. through their Ka and Kb values,
respectively).

¾ The Ka is the equilibrium constant for the hydrolysis reaction between


water and the acid, while the Kb is the equilibrium constant for the
hydrolysis reaction between water and the base (vide infra):

Acid Hydrolysis

HA(aq) + H2O(l) H3O+(aq) + A-(aq)

[H 3O + ][A − ]
Ka = ; [H 2 O](pure liquid) = 55.5 M( Blended into K a )
[HA]
Base Hydrolysis

B‫( ׃‬aq) + H2O(l) BH+(aq) + OH-(aq)

[BH + ][OH − ]
Kb = ; [H 2 O](pure liquid) = 55.5 M( Blended into K b )
[B :]

For strong acids, the hydrolysis equilibrium lies far to the right
(Ka >>1), while
For weak acids, the hydrolysis equilibrium lies far to the left (Ka<< 1).
Similarly,
For strong bases, the hydrolysis equilibrium lies far to the right
(Kb >>1), while
For weak bases, the hydrolysis equilibrium lies far to the left (Kb<< 1).
¾ e.g. CH3COO-(aq) is weak base (meaning much weaker
than OH-(aq)), as indicated by its small Kb value
(Kb CH3COO-(aq) = 5.68 × 10-10 M).

¾ NaNH2, on the other hand, is a very strong base (meaning


much stronger than OH-(aq)), as indicated by its large Kb
value (Kb NH2- >>100).
¾ The Ka values for a selection of acids, from weak to strong, is given in the table below:
¾ If an acid is strong and hence easily loses protons to
water, then its conjugate base does not readily accept
protons. Hence, the strength of an acid is inversely
related to the strength of its conjugate base. Similarly,
strong bases produce relatively weak acids while weak
bases produce relatively strong acids.

¾ This behaviour can be summarized in the following


manner: “Strong acids yield weak conjugate bases while
weak acids yield relatively strong conjugate bases”.
Mechanism of Acid Hydrolysis

¾ In acids, the ionizible hydrogen atom is attached to an


electronegative atom. The resulting polar bond can be cleaved by
water, thereby facilitating departure of H+ from the neutral acid
molecule as hydronium ion.

¾ For instance, the attack of a water molecule on the highly polarized


H-Cl bond of a hydrochloric acid molecule results in the release of a
hydronium ion into solution (vide infra):
Mechanism of Base Hydrolysis

¾ When they are dissolved in water, bases produce hydroxide ion via
deprotonation of water molecules.

¾ The weak base ammonia, for instance, can produce hydroxide ion
via nucleophilic attack of its lone pair on one the polarized O-H
bonds in an adjacent water molecule (vide infra):
Acid-Base Behaviour and Chemical Structure
¾ Next we will consider what factors influence the strength of acids.

Binary Acid
¾ Binary acids are the simplest acids known. They contain hydrogen along with
a non-metal element; H2Se and HCl are examples.

¾ Binary acid strength increases from left to right across a period and from top
to bottom within a vertical group (vide infra):
¾ In general, the H-X bond strength is the most important determinant of acid strength.

¾ The more polar the H-X bond, the stronger the acid and the weaker the H-X bond, the
stronger the acid.

¾ As X becomes more electronegative, the H-X bond becomes more polar and hence
more susceptible to cleavage.

¾ Thus, the acidity of binary acids formed from elements in the same period increases
from left to right.

¾ For instance the binary acids formed from elements in the latter half of the second row
follow trend
¾ The strength of an H-X bond tends to decrease as the element X
increases in size.

¾ Thus, H-X bond strength decreases and the acidity of the binary acid
increases down a given group.

e.g. H2S (Ka1 = 9.80 × 10-8 M) is a stronger acid than H2O


(Ka1 = 1.80 × 10-16 M), due to the fact that H-S bonds are weaker
than O-H bonds, and hence easier to break.

Oxyacids

¾ Acids in which OH groups and additional oxygen atoms are attached


to a central atom are called “oxyacids”.
General Structural Formula
Where Z = non-metal atom ; W, X , Y = O, OH or
electron pair W
Z
Y OH
X

e.g. sulfuric acid is an oxyacid with Z = S, X = OH, Y= O and W = O :


¾ As the electronegativity for the central atom Z increases, the acidity
for the oxyacid increases.

Consider, for instance, the trend for the hypohalous acids below:

¾ As Z becomes more electronegative, 1) the O-H bond becomes more


highly polarized, thereby making heterolytic cleavage of the O-H bond
to give H+ and -O- more facile.
¾ In addition, when Z is more electronegative, the stability of the
conjugate is increased, and this leads to the hydrolysis products
being even more highly preferred.
¾ The strength of the oxyacid is also increases as additional
electronegative elements are bonded to the central atom Z.
Increasing the number of electronegative elements attached to Z
further stabilizes the anionic conjugate base by giving it the ability to
delocalize its negative charge over a greater number of
electronegative atoms.

e.g. Placing more and more oxygens on the Cl atom of HOCl leads to
a steady increase in acidity
Levelling Effect
¾ All acids stronger than H3O+(aq)
(Ka = 1.8 × 10-3 M) ionize to ca. 100% in water and appear equally
strong (i.e. the strongest acid that can exist in water is H3O+(aq)
because all stronger acids protonate water virtually quantitatively).

¾ Similarly, all bases stronger than OH-(aq) virtually quantitatively


deprotonate water to give OH-(aq).

¾ Therefore, strong bases are leveled to the strength of OH- and thus
the strongest base that can exist in water is hydroxide ion.

e.g. CH3Li(s) + H2O(l) LiOH(aq) + CH4(g) (≈ 100 % rxn)


(v. strong base)

HClO4(aq) + H2O(l) H3O+(aq) + ClO4-(aq) (≈ 100 % rxn)


(v. strong acid)
Practice Problem:
For the set of bases, Cl-, F-, CN- , O2- and OH-, arrange in
order of decreasing base strength and classify each as
strong, weak or extremely weak by considering the
strength of their conjugate acids.
The Self-Ionization of Water: pKa for H2O and pH/pOH

Water, by virtue of being amphoteric, undergoes a self-ionization


reaction (vide infra):

H2O(l) + H2O(l) H3O+(aq) + OH-(aq)


(acid) (base) (acid) (base)

Archetypal equilibrium Constant Expression (Karchetype):


[H 3O + ][OH - ]
K archetype × [H 2 O] =
2

K archetype × [H 2 O]2 = K w

K w = [H3O + ][OH - ] = 1.00 × 10−14 M 2


(at 25 o C)
¾ The Kw is the equilibrium constant for the self-dissociation of water.

¾ Its small size indicates that undissociated water molecules very


strongly predominate at 25oC in pure water.(hydroxide/proton
concentrations are equal and very tiny).

¾ The concentrations of OH-(aq) and H3O+(aq) in water are frequently


expressed in terms of the pH and pOH (logarithmic system for
expressing concentrations):

pH = -log10[H3 O+]; Therefore [H3O+ ] = 10-pH

pOH = -log10[OH-]; Therefore [OH - ] = 10-pOH


¾ By taking the log base ten of both sides of the Kw
expression we get the logarithmic form of the expression,
viz.

pH + pOH = 14.000

¾ The Kw and pKw are constants, but the proton and


hydroxide ion concentrations can vary, as long as their
product is 1.00 × 10-14 M2 at equilibrium.
¾ The great advantage of the logarithmic pH/pOH scale is
that it gives more manageable values that range from ca.
-1 to 14 (vide infra):
Neutral Solutions: −14 −7
[H3O ] = [OH ] =
+ - 1.00 × 10 M 2
= 1.00 × 10 M
Therefore, pH = pOH = 7.000
for neutral solutions in which # protons = # hydroxide ions (neither
predominates)

Acidic Solutions:
−7 −7
[H3O+] > 1.00 × 10 M ; [OH-] < 1.00 × 10 M ;
There are more protons than hydroxide ions in these solutions
Thus, pH < 7.000 while pOH > 7.000

Basic Solutions:
−7 −7
[H3O+] < 1.00 × 10 M ; [OH-] > 1.00 × 10 M ;
There are more hydroxide ions than protons in these solutions
Thus, pH > 7.000 while pOH < 7.000
pH of a Solution of a Strong Acid or Strong Base

¾ Because strong acids and strong bases are virtually completely


ionized in water their pH/pOH can be calculated directly from the
concentration of the acid or base provided the concentration of the
acid or base is greater than 10-6 M; this allows the proton and
hydroxide ion contributions from the self-ionization of water to be
ignored.

Sample Problem:
Calculate the [OH-], [H3O+], pH and pOH for a solution prepared by
dissolving 0.60 g of Ca(OH)2 in pure water and making it up to
the1.00 L mark in a volumetric flask at 25oC.
Relationship between Ka and Kb for a Conjugate Acid-Base Pair in H2O

Let’s imagine that we are given an aqueous solution containing equimolar


amounts of HA and A-.
We start by adding together the base hydrolysis reaction for the base and the
acid hydrolysis for the conjugate acid:

Acid Hydrolysis:
[H 3O + ][A - ]
HA(aq) + H2O(l) H3O+(aq) + A-(aq) Ka =
[HA]
+
Base Hydrolysis
[OH - ][HA]
A-(aq) + H2O(l) HA(aq) + OH-(aq) Kb =
[A - ]

___________________________________________________________
Net: + - -
[H O ][A ] [OH ][HA]
2H2O(l) H3O+(aq) + OH-(aq) K a × K b = 3
×
[HA] [A - ]
Or

2H2O(l) H3O+(aq) + OH-(aq) Ka × Kb = K w

Ka × K b = K w

This equation quantifies the principle that a strong acid has a weak
conjugate base and vice versa (i.e. If the Kb for the base is large, then
the Ka for the conjugate acid will be small). Moreover, the fact that the
product of the two hydrolysis equilibrium constants is equal to the self-
ionization constant for water tells us that all acid-base reactions that
occur in water are intimately linked to the self ionization reaction for
water itself.
Taking log10 of both sides of the expression above we
obtain

-log10 Ka + -log10 Kb = -log10Kw

-log10 Ka = pKa ; -log10 Kb = pKb ; -log10Kw = pKw

Thus, pKa + pK b = pK w = − log10 (1.00 × 10−14 M 2 ) = 14.000

Important Note: The conjugate base of a weak acid is a still a weak


base and the conjugate acid of a weak base is still a weak acid.

The conjugate base of a strong acid is, however, exceedingly weak.


Equilibrium Position in Acid-Base Reactions

¾ For any acid-base reaction, the equilibrium will always favour the side
with the weaker acid and the weaker base.

Stronger Acid + Stronger Base Weaker Acid + Weaker Base

Weaker Acid + Weaker Base Stronger Acid + Stronger Base

Sample Problem:
Write the equation for the reaction of HSO4-(aq) with HS-(aq) and
determine if the equilibrium position favours reactants or products.
(Note: If the reactants are amphoteric, the species with the larger Ka
will act as an acid).
Acid-Base Properties of Salts: Neutral Solutions, Cation Hydrolysis and
Anion Hydrolysis

¾ A salt is another name for an ionic compound.

¾ When a salt dissolves in water, it breaks up into its constituent cations


and anions.

¾ Certain ions, when released, will react with water (i.e. hydrolyze) to give
acidic or basic solutions, depending on the natures of the anion and
cation.

Ions that Give Neutral (pH = 7.00) Solutions

¾ Conjugate bases of strong acids (e.g. Cl-, Br-, I-, NO3-, ClO4-) do not
hydrolyze to a measurable extent.Thus, these ions have no detectable
effect on the pH.

¾ Monoatomic cations with positive charges of +1 and +2 also effectively


do not hydrolyze. (i.e. cations such as Mg2+, K+, Na+, Sn2+ do not have
any detectable impact upon solution pH).
Ions that Give Basic (pH > 7.00) Solutions

¾ Basic solutions are formed by anionic, conjugate bases of weak


acids.
¾ The solution is made basic (i.e. an excess of hydroxide ions over
protons results) via a reaction between the anionic conjugate base
and water, which produces hydroxide ion and the conjugate acid (vide
infra):

[HA][OH − ]
A-(aq) + H2O(l) HA(aq) + OH-(aq) Kb =
[A - ]

Hydroxide ions outnumber protons in the solution at equilibrium; thus


pH > 7.00. To find the Kb for the base, use the equation
Kw
Kb =
Ka (for the conjugate acid)
Ions that Produce Acidic Solutions (pH < 7.00)

¾ Type I: Conjugate acids of weak bases produce acidic solutions (i.e.


those in which protons outnumber hydroxide ions) via the acid
hydrolysis reaction between water and the weak acid (vide infra):
+
[B:][H O ]
BH+(aq) + H2O(l) B:(aq) + H3O (aq) K a =
+ 3

[BH + ]
Protons outnumber hydroxide ions in the solution at equilibrium; thus
pH < 7.00

¾ Type II: Highly Charged, Hydrated Metal Ions undergo an internal


rearrangement in water that results in the release of protons into
solution, which pushes the pH below 7.00.
e.g. AlCl3, when dissolved in water, is converted into the hexahydrate
trication [Al(H2O)6]3+(aq) (vide infra):
¾ This complex ion releases protons via the abstraction of an hydroxide
group from a coordinated water molecule:

[Al(H2O)6]3+(aq) [Al(H2O)5(OH)]2+(aq) + H+(aq)

[Al(H 2O)5 (OH)2+ ][H + ] −6


Ka = 3+
= 7.9 × 10 M
[Al(H 2O)6 ]
(protons outnumber hydroxide ions in solution)

Sample Problem:
For aqueous solutions of the following two salts, classify each of the ions
as acidic, basic or neutral. Write a net ionic equation for any hydrolysis
reaction, indicate the value of the Ka or Kb as appropriate and state
whether the resulting solution will be acidic, basic or neutral.
a) KBr
b) NH4CH3COO
Calculating Equilibrium Constants for Acid-Base Reactions in Aqueous
Solution

¾ To correctly determine the equilibrium constant for the reaction


between an acid and a base in water, you must not only take into
account the two relevant hydrolysis reactions, but also the reverse of
the self-ionization reaction for water itself.

e.g. Calculate Koverall for the reaction

CN-(aq) + CH3COOH(aq) CH3COO-(aq) + HCN(aq)


Weak Acids/Weak Bases: Equilibrium Problems

¾ The vast majority of weak acid or weak base hydrolysis


questions can be solved by using the small K method.
Nonetheless, some starting mixtures do not obey the 5%
rule and must be solved as intermediate K problems.

¾ To simplify matters we will restrict our discussion here at


the start to hydrolysis reactions for weak monoprotic
acids. Similar arguments will apply, of course, to the
hydrolysis reactions of monobasic compounds, and these
will be covered later in this section.
General Approach for Finding [H3O+] and pH at
Equilibrium for a Weak Acid HA

e.g. Calculate the [H3O+] and pH at equilibrium for a 0.010 M


aqueous solution of HOCl (hypochlorous acid).
Note: Ka for HOCl = 3.5 × 10-8 M

Solution:
First step: Write down the major species in the mixture
In this case, the two major species are H2O(l) and HOCl(aq)

Ka for HOCl = 3.5 × 10-8 M


Ka1 for H2O = 1.8 × 10-16 M
¾ By comparing the two Ka values, we can clearly see that HOCl is a
much stronger acid than water and, since the [HOCl] is much greater
than 10-6 M, we can ignore the tiny contribution of H+ ions from water
itself.

(Please note, however, that if the Ka for the dissolved acid is smaller
than 10-11 or if its starting concentration is less than 10-6 M, then the
initial 10-7 M contribution of protons from water must be included in
the “I” line of the ICE table).

¾ Consequently, we need only consider the dissociation of HOCl to


come up with a very good estimate of the equilibrium concentration of
[H3O+] and thus the pH.

¾ To answer this question we must set up and solve an ICE in much the
same manner as we did in the preceding fundamentals of chemical
equilibrium section. The extent of hydrolysis will depend upon the size
of the Ka and the starting concentration for the acid. When Ka is very
small, as is the case here, and the starting concentration of
appreciable size (greater than 0.001 M), the reaction will proceed to
only a tiny extent, and thus the 5 % will apply, thereby simplifying the
algebraic solution.
ICE Table:

HOCl(aq) + H2O(l) H3O+(aq) + OCl-(aq) Ka = 3.5 × 10-8 M

I(M) 0.010 - ≈0 0 Qa = 0; Qa < Ka


C(M) -x - +x +x Thus, → Dir of rxn

E(M) 0.010 – x - x x Ca
= 2.8 × 10 > 1000
5

Invoke 5% rule Ka
∴ Small K problem
E(M) (≈ 0.010 ) x x

−8[H 3O + ][OCl- ] x×x


3.5 × 10 M = =
[HOCl] ≈ 0.010

x2 = 3.5 × 10-10 M2 ; x = 1.87 × 10-5 M ; Thus [H3O+] [HA]initial × K a

[H3O+]equil ≈ 1.87 × 10-5 M ; pH = -log10(1.87 × 10-5 M) = 4.73(2 SF)


5% error check: 1.87 × 10-5 M(x)
× 100% = 0.19% error
0.010 M

The error is clearly far less than the 5% max. allowed and thus our
approximation was valid.

Percent Dissociation or Percent Ionization


¾ The percent dissociation or percent ionization is an important
calculation in acid-base chemistry that describes the extent to which a
particular acid-base reaction occurs.

¾ The dissociation of a weak acid, for example, is given by

[H3O + ]equil
Percent Dissociation = × 100%
[HA]initial
For a weak base, the corresponding equation would be

[OH - ]equil
Percent Dissociation = - × 100%
[A ]initial

¾ For a given weak acid or weak base, the percent


dissociation increases as the acid or base becomes more
dilute.
e.g. 1.0 M CH3COOH(aq) is 0.42 % dissociated while a 0.010 M
CH3COOH(aq) is 4.2 % dissociated (i.e. 100-fold dilution results in a 10-fold
increase in percent ionization).

This plot shows that the extent is reaction is determined not only by the size
of the K value, but also upon the starting concentration for the acid or base.
¾ Bear in mind that although dilution does indeed increase the extent of
hydrolysis, the net effect is, nonetheless, a decrease in either the
proton concentration (for an acid hydrolysis equilibrium) or the
hydroxide ion concentration(for a base hydrolysis equilibrium).

¾ Also, it should be pointed out here that all acids and bases(weak and
strong) at infinite dilution are 100% dissociated and the solution
pH = 7.00.

Dilution increases the ratio [OH - ]equil [H3O + ]equil


-
or
[A ]initial [HA]initial

(Ka/Kb of course, are not affected by dilution)


Why does percent ionization increase with decreasing starting
acid/base concentration?”

Answer:
Consider the weak acid hydrolysis equilibrium

HA(aq) + H2O(l) H3O+(aq) + A-(aq)

If we let [H3O+]equil = [A-]equil = x , and assume [HA]initial ≈ [HA]equil ,

then Ka = x 2
; Now we dilute the solution by a factor of 10:
[HA] 2
⎛ x⎞
⎜ ⎟
Qa = ⎝ ⎠
10
⎛ [HA] ⎞
⎜ ⎟
⎝ 10 ⎠
(At this stage, the initial concentrations of all species are decreased by a
factor of 10)
Simplification gives

⎛ x2 ⎞ ⎛ x2 ⎞
⎜ ⎟ ⎜ ⎟
⎝ 100 ⎠ 100[HA] 2
= ⎝ ⎠ =
Qa = x
⎛ [HA] ⎞ (10 ) 10 [HA]
⎜ ⎟
⎝ 10 ⎠

1 ⎛ x2 ⎞
So, Qa = ⎜ ⎟ Qa = 1/10 Ka
10 ⎝ [HA] ⎠
Since Q is less than K, the equilibrium will shift to the right to reestablish
equilibrium. (i.e. the system must bump up [H3O+] and [A-] above 1/10
original and push down [HA] below 1/10 original to raise Q back up to K).
This shift will cause the percent dissociation to increase. Thus, dilution is a
Le Chatelier stress upon the system that the system alleviates by shifting the
equilibrium to the side of the equation with ions.
Weak Base Hydrolysis Equilibrium Problems
¾ Weak bases (such as NH3) react with water by abstracting protons
from water molecules, thereby increasing the [OH-] concentration.

B:(aq) + H2O(l) BH+(aq) + OH-(aq)


(base) (conjugate acid)

[BH + ][OH − ] ; [OH-] > [H3O+] ; basic solution


Kb =
[B:]

¾ The pH and percent ionization calculations for solutions of weak


bases in water are very similar to those for weak acids outlined
earlier. The degree of ionization is determined by the Kb and the
starting concentration for the base.

¾ In those cases where the 5% rule is obeyed


[OH-] [B]initial × Kb (i.e. x must be less than 5% of [B]init for this
calculation to be valid)
Practice Problem:
Calculate the pH and percent ionization for a 0.50 M
solution of methylamine (CH3NH2) in water.
Alternate Algorithm for Solving Acid-Base Equilibria Problems

¾ A system of equations is a group of two or more equations that


possess the same variables. A solution to a system of equations is a
set of values for the variables that simultaneously satisfy all of the
equations.

¾ All acid-base equilibria questions can be solved by using the systems


of equations method outlined below.

¾ To simplify matters, we will only consider acid hydrolysis here. Of


course, the same methodology we are about to explore can be
applied equally as well to any base hydrolysis reaction.

First we must define the variables x and y in terms of certain acid-


base relationships:
For instance, we can
Let x = [H3O+] at equilibrium = [A-] at equilibrium and
Let y = [HA] before any acid hydrolysis occurs
¾ Thus, the acid-hydrolysis reaction ICE table would be
HA(aq) + H2O(l)
H3O+(aq) + A-(aq)
I(M) y - 0 0
C(M) - x - +x +x
E(M) y-x - x x

¾ By using these definitions, we end up generating at equilibrium the


generic Ka expression

( x) × ( x) x2 (Equation1)
Ka = =
( y − x) y−x
Note that [HA] at equilibrium is equal to y-x in the Ka expression and
that not only is x equal to the equilibrium concentrations for
hydronium ion and conjugate base, it also represents the amount of
HA that ionizes.
¾ At this stage we have two unknowns (assuming the Ka is available) but
only one equation. To find both x and y and solve the equation, we
need to develop a second equation involving x and y. Luckily, there is
another acid-base equation involving both x and y that we can use for
our purposes.

The equation in question is the percent ionization equation, which has


the general format

Percent Ionization = Concentration of acid that ionizes


× 100%
Acid concentration prior to hydrolysis

¾ When our particular variables, as defined above, are inserted into this
expression we end up with a second equation containing x and y, viz.

[H 3O + ]equil (or[A - ]equil ) x


Percent Ionization = × 100% = x 100%
[HA]initial y
(Equation 2)
¾ With two equations containing the same two variables in hand, we
can now simultaneously solve both equations through the process of
substitution.

Percent Ionization
e.g. We can replace x in Equation 1 with ×y
100%

which will produce an equation that has only the variable y. Once we
have found y using algebra, x is readily found via substitution back
into either Equation 1 or Equation 2.

¾ The great strength of the systems of equations method is that it can


be used to find any or all of the following quantities for any acid
hydrolysis problem you will encounter in this course:
[HA] initial, [HA]equilibrium, [H3O+]equilibrium, [A-]equilibrium, Percent Ionization
and Ka
Buffers and the Common Ion Effect

¾ As a prelude to buffers, we will focus our attention on the common ion


effect, which is an integral part of the buffering phenomenon.

Common Ion Effect

¾ The common ion effect is simply a shift in equilibrium caused by the


addition of a salt that has an ion in common with one of the dissolved
species.

¾ Consider, for instance, an aqueous solution of hydrofluoric acid. Since


HF is a weak acid, for a solution of moderate concentration, only a
small fraction of the HF molecules will undergo hydrolysis:

HF(aq) + H2O(l) H3O+(aq) + F-(aq) Ka = 6.63 × 10-4 M


¾ If some solid NaF is added to a solution of the acid at equilibrium,
according to Le Chatelier’s principle, the equilibrium will shift to the
left to offset the increase in F- concentration
(i.e. adding F- makes Q > K).

HF(aq) + H2O(l) H3O+(aq) + ↑ F-(aq)


← Dir. of equilibrium shift

¾ This shift will, of course, also drive down the [H3O+]. Thus, adding the
common F- ion to the solution causes a shift towards the
undissociated acid. The net result of the addition of a common ion,
therefore, is a repression of the hydrolysis reaction, in this case acid
hydrolysis.
Common Ion Effect Practice Problem:

a) What is the pH of a 0.10 M CH3COOH(aq) solution?


b) What is the pH after adding enough sodium acetate to
make the solution 0.10 M in CH3COO-(aq)?
Buffers

¾ A buffer is a solution that resists changes in pH when either H3O+ or


OH− ions are added. Buffering solutions contain roughly equal
amounts of a weak acid and its conjugate base (vide infra):
¾ Some examples of buffer pairs are NH4+/NH3, HCN/CN−,
CH3COOH/CH3COO−.

¾ The concentrations of HA and A− in a buffer solution are


much greater than the hydroxide and proton
concentrations.

¾ The pH of a buffer solution is determined by the Ka (or


Kb) and the buffer ratio.

[HA]
- is the buffer ratio
[A ]
e.g. For an acidic buffer solution (one with a pH < 7):

HA(aq) + H2O(l)
H3O+(aq) + A−(aq)
(acid) (conjugate base)
I(M) Ca - ≈0 Cb
C(M) – x - +x +x
E(M) Ca – x - x Cb + x
≈ Ca ≈ Cb
[H 3O + ][A - ]
( Ca and Cb >> [H3O+ ], [ OH−] ); Ka =
[HA]
[HA] [HA]
[H 3O ] = K a
+
- ; - is the buffer ratio
[A ] [A ]

Note : The concentrations for both HA and A− tend to be large in


buffer solutions.
¾ The concentration of [H3O+] in a buffered solution is small
because:

a) The acid has a small Ka


b) The acid ionization is repressed by the A− in solution
(common ion effect).
c) High concentrations of HA and A− are used

¾ Now, let’s covert the Ka expression into an equivalent logarithmic


form:
Taking − log10 of both sides of the expression, followed by
rearrangement to give pH by itself on one side affords

⎛ [HA] ⎞ ⎛ [A - ] ⎞
pH = pK a + - log ⎜ - ⎟ or pH = pK a + log ⎜ ⎟
⎝ [A ] ⎠ ⎝ [HA ] ⎠

This is known as the Henderson-Hasselbalch equation. It can be


used to estimate the pH of buffer solutions.
¾ The H-H equation assumes that the initial
concentrations for the acid and its conjugate base differ
from the corresponding equilibrium concentrations by no
more than 5%. Thus, the H-H equation gives results
with acceptable error (i.e. < 5%) when:

1. [HA] and [A−] are greater than 0.01 M


2. The Ka and Kb values for the conjugate acid-base
pair fall between 10-3 and 10-11
3. The buffer ratio is between 0.1 and 10.

¾ In other words, if a quadratic equation is needed to


calculate the equilibrium concentration for either the
acid or the conjugate base, then the standard H-H
equation will fail (i.e. terms involving x are not usually
not inserted into this equation!)
¾ For basic buffer solutions it is sometimes more convenient to use
the basic form of the H-H equation(vide infra):

Viewed from the perspective of base hydrolysis, the relevant


equations are

B(aq) + H2O(l)
BH+(aq) + OH−(aq)
(base) (conjugate acid)

[OH - ][BH + ] [B] [B]


Kb = ; -
[OH ] = Kb ; is the buffer ratio
[B] [BH ]
+
[BH ]
+

⎛ [B] ⎞ ⎛ [BH + ] ⎞
pOH = pKb + - log ⎜⎜ + ⎟⎟ or pOH = pKb + log ⎜ [B] ⎟
⎝ [BH ] ⎠ ⎝ ⎠
Buffering Action
¾ When a strong acid is added to an HA/A− buffer solution, the added H+ ions
react essentially to completion with the conjugate base A− (via OH− and the
self-ionization reaction for water) because the Kneutralization for this process is
large:
H3O+(aq) + A−(aq)
HA(aq) + H2O(l) Kneut >103
(products strongly favoured)

Note: For the neutralization of added protons Kb A −


K neut =
Kw
¾ Similarly, when a strong base is added to the buffer, the added OH− reacts
essentially to completion with the weak acid HA (via H3O+ and the self-
ionization reaction for water) again, due to the large size of the Kneut for the
process:
OH−(aq) + HA(aq) A−(aq) + H2O(l) Kneut >103


(products strongly favoured)

Note: For the neutralization of added hydroxide ions K a HA


K neut =
Kw
¾ In general:

For the system…

HA + H2O
A− + H3O+

Add OH−: → Equilibrium shifts to the right to offset the stress

Add H+: ← Equilibrium shifts to the left to offset the stress

If [HA] and [A-] are large compared to the amount of H3O+ or OH− that
is added, then the concentrations of [HA] and [A−] will change by only
by small amounts and thus buffer ratio will also change very little, and
thus the pH will increase or decrease only very slightly.

The buffering agents in the water, viz. HA and A−, react with the
added hydroxide ions or protons, respectively, and thereby severely
curtail the change in pH. Essentially, the A− in the buffer converts the
added strong acid H3O+ in to the much weaker acid HA, while the HA
present converts the added strong base OH− into the much weaker
base A− and, in both instances, the solution experiences, as a result,
only a minor alteration in the pH.
Most Effective Buffers

¾ The most effective buffers have a buffer ratio of one


(i.e. [HA] = 1 ) and the highest possible concentrations
[A - ]
for [HA] and [A−] (greater than 1M, if possible).

¾ As long as the buffer ratio remains essentially


unchanged at one, the pH of the solution will remain
essentially fixed at the pKa for the weak acid.
[HA]
[HA] = [A−]; - = 1; pH = pK a + log[1] ;
[A ]

Thus, pH = pKa
¾Ideal buffers are effective up to a pH change
of ± 1 unit. Once the buffer ratio drops below
0.1 or climbs above 10, the solution can no
longer buffer to any appreciable extent. In other
words, the buffer capacity is almost completely
exhausted once the pH has been changed by 1
unit.

¾Clearly, then, there is a limit to how much


strong acid or base can be added to a buffer
solution before one of the active components is
used up. This limit, called the buffer capacity
is defined as the number of moles of either
strong acid or strong base that must be added
to change the pH of 1.0 of the buffer solution
by one unit.
Buffer Capacity = (number of moles of OH- or H3O+ added)
(one unit pH change)x (vol. of buffer soln’in L)

¾ The greater the concentrations for the acid and conjugate


base, the greater will be the buffer capacity for the solution

Final Note: One can construct a buffer solution at virtually any


pH by
1) selecting the appropriate acid-base pair and
2) manipulating the buffer ratio.
Buffer Practice Problem
a) Calculate the pH of a buffer solution
prepared by adding 15.0 mL of
2.5 M HCl(aq) to 300. mL of 0.25 M
NH3(aq). Kb NH3= 1.80 × 10−5 M
(Buffer Construction Problem)

b) Let’s imagine that 10. mL of 0.15 M


HClO4(aq) were added to the buffer
solution prepared in part a). What would
be the pH of the resulting solution after the
buffering reaction goes to completion?
(Buffering Action Problem)
Titration Curves

¾ An acid-base titration curve is a plot of solution pH


(y-axis) against volume of titrant added, usually in mL, added during
the course of the titration(x-axis).

¾ pH is measured with a pH meter equipped with a H+ ion-sensitive


electrode.

¾ The theoretical pH of the solution at each stage of the titration can be


calculated.

Titration Curve for the Titration of a Strong Acid with a Strong Base

¾ For the reaction of a strong acid with a strong base in aqueous


solution, the net ionic equation for the neutralization reaction is
1
H3 O+(aq) + OH-(aq)
2H2O(l) Kneut = = 1.00 × 1014 M −2
Kw
¾ Thus, at each point in the titration, the equilibrium for the
titration reaction strongly favours products (i.e. the titration
reaction at all points is essentially stoichiometrically
complete).

¾ This quality combined with the fact that the initial pH for a
strong acid is small causes the strong acid titration curve
to have a large pH change near the equivalence point.

¾ Near the equivalence point, the pH changes rapidly


because only small concentration changes are needed to
produce big pH unit changes.
e.g. Consider the titration curve for the reaction of 0.100 M HCl(aq) with
0.100 M NaOH(aq) shown below:
Strong Acid-Strong Base Titration Curves have Three Distinct Regions

1) Before equivalence, [H3O+(aq)] and the pH of the solution are


determined by the concentration of untitrated acid. Between the
start of the titration and the equivalence point, there are more
protons than hydroxide ions in solution.

2) At the equivalence point (inflection point of the curve), all of the


H3O+ released by the acid has reacted with an equimolar quantity
of OH-(aq). Thus the pH at this point, viz. 7.00, is determined
solely by the self-ionization reaction for water (pH = -log10
(1.00 × 10-7 M)). The anions and cations of the neutral salt formed in the
titration do not hydrolyze to an appreciable degree and thus do not effect
the pH of the solution.

3) Beyond the equivalence point, the pH of the solution is determined by the


concentration of excess OH- ions from the strong base added.
Sample Problem: Strong Base-Strong Acid Titration

For the titration of 25.0 mL of 0.200 M KOH(aq) with


0.15 M HCl(aq), calculate the pH at the following points:
a) Initial pH; b) Equivalence point ; c) 1.0 mL past
equivalence
Titration Curve for the Titration of a Weak Acid with a
Strong Base

¾ When solutions of a weak monoprotic acid and a strong


base are mixed, the neutralization reaction outlined
below takes place:
HA(aq) + OH-(aq) H2O(l) + A-(aq) Koverall = ?

The overall the neutralization reaction is the sum of the


acid hydrolysis reaction for HA and the reverse of the
self-ionization reaction for water:
H2O(l) + HA(aq) H3O+(aq) + A-(aq) ( Ka )
+
⎛ 1 ⎞
H3O+(aq) + OH-(aq) 2H2O(l) ⎜ ⎟
⎝ w⎠
K
¾ Therefore, the overall equilibrium constant for the
neutralization reaction is the product of the equilibrium
constants for the two component processes:

1 K
K overall = K a × = a
Kw Kw

1
(Note : = 1.00 × 1014 M −2 )
Kw

The weak acid is a much stronger acid than water


(Ka >> Kw) and consequently the neutralization will be
virtually quantitative (i.e. it will proceed to ca. 100 %
completion as long as Ka >10-11).
e.g. Consider CH3COOH(aq) being titrated with OH-(aq):

−5
CH3COOH(aq) + H2O(l)
CH3COO (aq) + H3O (aq) Ka = 1.76 × 10 M
- +

+
1
H3O+(aq) + OH-(aq) 2H 2 O(l) = 1.00 × 1014 M −2
Kw

________________________________________________________
Net Rxn:
CH3COOH(aq) + OH-(aq) H2O(l) + CH3COO-(aq) Ka
K overall =
Kw

1.76 × 10−5 M −1
K overall = = 1.76 × 10 9
M (v.large)
1.00 × 10 M
−14 2

We can conclude, therefore, that the reaction will proceed 100 % to the right
and the neutralization will be essentially quantitative.
¾ A generic weak acid-strong base titration curve is shown
below.
¾ Weak Acid-Strong Base Titration Curves Have
Six Distinct Points/Regions, as indicated in the
previous figure:

1. Initial Point – Before any base has been


added we have just an aqueous solution of
the weak acid HA. To find the pH of the solution,
we need only solve the ICE table for acid
hydrolysis.

2. Pre-Buffer Region – This is the short interval


between the initial point and the start of the
buffer region. In this region, the [HA] > 10[A-].
3. Buffer Region

¾ In this region, some OH- has been added, but not sufficient to
neutralize all of the acid present. More specifically, the ratio in this
region falls between 10 and 1/10.

The added OH- reacts essentially quantitatively according to the


reaction…

CH3COOH(aq) + OH-(aq)
CH3COO-(aq) + H2O(l)

Kneut = 1.76 × 109 M-1

Thus, this reaction converts HA to the corresponding conjugate base


A-.

The unreacted HA along with the A- produced by the neutralization


reaction, buffers the solution so that the pH is determined by the
position of the acid dissociation equilibrium.
CH3COOH(aq) + H2O(l)
CH3COO-(aq) + H3O+(aq)
(weak acid) (conjugate base)

The pH does not change drastically in this region due to


buffering action.

The pH in this region can be calculated with the


Henderson-Hasselbalch equation (vide infra), as long as
the 5 % rule is obeyed:

⎛ [A- ] ⎞
pH = pK a + log ⎜ ⎟
⎝ [HA] ⎠
4. Mid-Point of the Titration (Middle of the Buffer
Region)

At this point in the titration, half of the original number of


moles of HA has been converted into A-.

Thus at the halfway mark of the titration, [HA] = [A-]


(i.e. the buffer ratio is equal to one). This is the point of
maximum buffering in the titration.

[H3O+] = Ka , so solution pH = pKa for HA


at ½ equiv. point volume
5. Equivalence Point

¾ The point at which the number of moles of added OH-


equals the number of moles of HA originally present.
Thus, at this point in the titration all of the HA has been
converted to A-.

¾ The pH of the solution is determined by the hydrolysis


reaction between the conjugate base A- and water, which
releases OH- ions into solution, thereby pushing
the pH above 7:

A-(aq) + H2O(l)
HA(aq) + OH-(aq)

In other words, the pH at equivalence for the titration of a


weak acid with a strong base is always above 7 because
the salt produced by the neutralization reaction is basic.
6. Excess Base Region Beyond Equivalence

¾ The solution in this region consists of the excess OH-


and the weak base A- produced by the neutralization
reaction.

¾ To calculate the pH in this region only the concentration


of excess OH- need be considered. The OH- contribution
from the A- hydrolysis reaction can be ignored because
the base hydrolysis reaction is repressed by the high
OH- concentration.

Important Note: Weak base-strong acid titrations are


essentially the inverse of weak acid-strong base
titrations described above, with the titration starting
at high pH, ending at low pH, and the pH at equivalence
being less than 7.
Sample Titration Curve Question: Weak Acid-Strong
Base Titration

For the titration of 20.0 mL of 0.250 M HF(aq) with


0.200 M NaOH(aq), calculate the pH at the following
points in the titration:
a) the initial point; b) the half-neutralization point; c) the
equivalence point; d) 1.00 mL past equivalence.
(Note: Ka HF = 6.63 × 10-4 M)
Solubility Equilibria

¾ In this section of the course we will consider equilibria arising from sparingly
soluble solids dissolving to form aqueous solutions of ions.

¾ Consider, for instance, the dissolution of solid calcium phosphate in water:


H2O
Ca3(PO4)2(s) ⎯⎯⎯
→ 3Ca2+(aq) + 2PO43-(aq)

¾ As dissolution proceeds, the concentrations of PO43-(aq) and Ca2+(aq)


steadily increase, making the probability of recombination to form the solid
more likely:

3Ca2+(aq) + 2PO43-(aq) → Ca3(PO4)2(s) ↓


⎯⎯

Eventually, a dynamic equilibrium is reached:

Ca3(PO4)2(s)
3Ca2+(aq) + 2PO43-(aq)

At this point, no more solid will dissolve


(Rate of Dissolution = Rate of Precipitation)
¾ Conversely, when solutions of Ca2+ and PO43- are mixed,
this equilibrium will be established in reverse.

¾ Increasing the concentration of Ca2+ or PO43- will favour


precipitation while lowering the concentration of either or
both of the dissolved ions by adding pure water will
cause more of the solid to dissolve.

¾ The equilibrium expression for solid calcium phosphate


in equilibrium with its dissolved ions is given by:
[Ca 2+ ]3 ×[PO 43− ]2
Keq[Ca3(PO4)2] = Ksp = = 1.3 ×10−32 M 5
1
for the dissolution reaction…
Ca3(PO4)2(s) 3Ca2+(aq) + 2PO43-(aq)
(some solid must be present)
VERY IMPORTANT:

¾ Do not confuse the solubility of a compound with its solubility


product!

Solubility Product (Ksp ) is an equilibrium constant, and as such, has


just one value for a given ionic compound at a given temperature.

Solubility, in contrast, is the extent to which the compound in


questions dissolves in a given solution.

In pure water at a specified temperature, a given salt has a particular


solubility, expressed usually in g/L or moles/L of saturated solution.

If a common ion is present, however, the solubility will vary with the
concentration of the common ion (i.e. the common ion represses
dissolution).
Relative Solubilities

¾ For salts that belong to the same formula class(i.e. those


salts that produce the same number of ions in the
solubility reaction), the solubility decreases with
decreasing Ksp.

e.g. Each of the three carbonate salts below produce one


cation and one anion when they dissolve in water. Thus,
they all belong to the same salt class. The salts are
arranged in order of decreasing solubility from Left to
Right. Notice that the smaller the Ksp, the lower the
solubility for the salt:

NiCO3(Ksp = 1.4 × 10-7 M2) > CaCO3(Ksp = 8.7 × 10-9 M2)


> CuCO3(Ksp = 2.5 × 10-10 M2)
¾ For ionic compounds that belong to different
formula classes (e.g. MgF2 vs. Ba3(PO4)2), the
relative solubilities cannot be assessed simply by
comparing their Ksp values because they
produce different numbers of ions when they
dissolve.

¾ In these instances, their solubilities must be


derived from their Ksp values.
e.g. Consider, for example, the following three
salts:
CuS(s) (Ksp = 8.5 × 10-45 M2), MX salt
Ag2S(s) (Ksp = 1.6 × 10-49 M3), M2X salt
Bi2S3(s) (Ksp = 1.1 × 10-73 M5), M2X3 salt
We can calculate the solubility for each salt at 25oC as
follows:

1) CuS(s)
Cu2+(aq) + S2-(aq);
x moles/L CuS ⎯⎯ → x moles/L Cu2+(aq) + x moles/L S2-(aq)
(molar solubility)
1 part CuS gives 1 part Cu2+ and 1 part S2-

Thus, Ksp = [Cu2+][S2-] = 8.5 × 10-45 M2 = x2


(equilibrium saturated solution)

x = 9.2 × 10-23 moles/L for CuS(s) to form a saturated solution at


25oC.(Molar Solubility for CuS)
2) Ag2S(s)
2 Ag+(aq) + S2-(aq)

x moles/L Ag2S ⎯⎯
→ 2x moles/L Ag+(aq) + x moles/L S2-(aq)
(molar solubility)

1 part Ag2S gives 2 parts Ag+ and 1 part S2-

Thus, Ksp = [Ag+]2[S2-] = 1.6 × 10-49 M3 = [2x]2[x]


(equilibrium saturated solution)

4x3 = 1.6 × 10-49 M3

x = 3.4 × 10-17 moles/L for Ag2S(s) to form a saturated solution at


25oC.
(Molar Solubility for Ag2S)
3) Bi2S3(s)
2 Bi3+(aq) + 3 S2-(aq)

x moles/L Bi2S3 ⎯⎯→ 2x moles/L Bi3+(aq) + 3x moles/L S2-(aq)


(molar solubility)
1 part Bi2S3 gives 2 parts Bi3+ and 3 parts S2-

Ksp = [Bi3+]2[S2-]3 = 1.1 × 10-73 M5= [2x]2[3x]3


(equilibrium saturated solution)

108x5 = 1.1 × 10-73 M5

x = 1.0 × 10-15 moles/L for Bi2S3(s) to form a saturated solution at


25oC.
(Molar Solubility for Bi2S3)

Therefore, the order of solubilities is :


Bi2S3 > Ag2S > CuS
which is precisely opposite to the order of the Ksp values!
Determining Ksp from Solubility Measurements

¾ The method by which Ksp values are obtained is the


opposite of the solubility calculations procedure we have
just considered.

e.g. The solubility of PbI2 is 0.060 grams per 100. mL of


pure water at 25oC.Use this information to calculate the
Ksp for PbI2 at 25oC.

Solution:
First we have to determine the equilibrium solubility for
PbI2 in moles per liter (i.e. the amount that must dissolve
to form a saturated solution).
⎛ 0.060 g PbI 2 ⎞ ⎛ 1mol PbI 2 ⎞ ⎛ 1000 mL ⎞ −3
⎜ ⎟ × ⎜ ⎟ ×
s (molar solubility) = ⎝ 100. mL ⎠ ⎝ 461 g PbI 2 ⎠ ⎝ 1 L ⎠⎜ ⎟ = 1.3 × 10 moles/L

Therefore, for the dissolution reaction

PbI2(s)
Pb2+(aq) + 2I-(aq)

s moles/L → s moles/L + 2s moles/L


(to form a saturated solution)

Ksp = [Pb2+][I-]2 = [s][2s]2 ;

4s3 = Ksp = 4 ×(1.3 × 10-3 M)3 = 8.8 × 10-9 M3.


Common Ion Effect on the Solubility of Precipitates

¾ Up until now, we have only considered the dissociation of ionic solids in


pure water.

¾ Now let’s see what happens when the water contains one of the ions
present in the dissolving salt.

¾ Using Le Chatelier’s principle, we would predict that the solubility of a solid


in a solution containing an ion in common with the solid is less than its
solubility in pure water. In other words, the solubility of the salt is lowered if
the dissolving solution already contains one or both of the ions that make
up the salt.

e.g. Consider the silver(I) acetate dissolution reaction:

AgCH3COO(s)
Ag+(aq) + CH3COO-(aq)

If we add Ag+ (as AgNO3), the equilibrium will shift to the left, forming more
solid :
AgCH3COO(s) ← Ag+(aq) ↑ (Conc’n Increases) + CH3COO-(aq)

The net result is that the solid has a lower solubility if external Ag+ ions (or
acetate ions) are added to the solution.
Sample Problem:
Calculate the solubility of PbI2 at 25oC in a 0.0100 M
solution of NaI(aq).
Precipitation Reactions

¾ Precipitation is the reverse of dissolution.

¾ In order to determine if a precipitate will form when two


solutions of soluble ionic salts are mixed, we need to
calculate the inverse ion product (Qppt) at the instant of
mixing and compare it to the Kppt.

¾ The Qppt differs from the Kppt in that the concentrations


used in the Qppt calculation are the initial values and not
necessarily those at equilibria. (c.f. the reaction quotient Q
covered in the Fundamentals of Chemical equilibrium
chapter).
¾ If the Qppt = Kppt, the system is at equilibrium (i.e.
the solution is saturated).

¾ If the Qppt > Kppt, the system is not at equilibrium,


no precipitate will form (i.e. the solution is
unsaturated).

¾ If the Qppt < Kppt, the system is not at equilibrium, a


precipitate will form immediately after the instant of
mixing (i.e. the solution is supersaturated).
Sample Precipitation Problem

a) Will a precipitate form when 200. mL of 0.0040 M


BaCl2(aq) are added to 600. mL of 0.0080 M
K2SO4(aq)? (Note: Ksp BaSO4 = 1.0 × 10−10 M 2 )

b) Calculate the Ba2+(aq) concentration at equilibrium.


Complex-Ion Equilibria

¾ A complex ion is a stable ion composed of many atoms that has been
formed by coordination (attachment) of anions or neutral molecules to
a central metal atom or ion.

¾ The anions or neutral molecules bound to the metal centre are called
ligands.

¾ Each ligand donates at least one pair of electrons to the central metal
atom.
e.g.

NiSO4(s) + 6H2O(l)
[Ni(H2O)6]2+(aq) + SO42-(aq)

Hexaaquo nickel (II) sulphate


Hexaaquo nickel (II) cation

Each water ligand in the complex ion is attached to the


central Ni atom via a lone pair of electrons.
Lewis Acid-Base Theory: Basis for the
Formation of Coordination Complexes

¾ Lewis Acid: Species that is the acceptor of a lone pair of


electrons in the formation of a coordinate covalent bond.

¾ Lewis Base: Species that is the donor of a lone pair of


electrons in the formation of a coordinate covalent bond.
A Typical Lewis Acid-Base Reaction:

(A) (B) (C)

(Ammonia adduct of BF3 (Molecule (C)): the BF3 has been stuck on the
lone pair of the NH3 molecule)
The B-N bond H3N−BF3 in the adduct is called a coordinate covalent or
dative covalent bond.
¾ Similarly, the formation of a transition-metal complex can
be viewed as a Lewis acid-Lewis base reaction.

e.g. CuSO4(aq) + 4 NH3(aq)


[Cu(NH3)4]2+(aq) + SO42-(aq)
Lewis Acid: Cu2+ Lewis Base: :NH3 Transition-Metal Complex

Microscopic view
Complex-Ion Equilibria: Formation Constants for
Complex Ions

¾ The formation of complexes involving monodentate


ligands (groups attached via one atom) is a stepwise
process, with each ligand being added successively to
the central metal ion.

e.g. Consider the formation of [Ag(NH3)2]+(aq):

Step 1
[Ag(NH 3 )]+ −1
[Ag(NH3)]+(aq) K1 = = × 3
Ag+(aq) + NH3(aq) [Ag + ][NH 3 ]
7.1 10 M
Step 2

[Ag(NH3)]+(aq) + NH3(aq)
[Ag(NH3)2]+(aq) [Ag(NH 3 ) 2 ]+
K2 = = 2.3 × 103 M −1
[Ag(NH 3 )]+ [NH 3 ]

The product K1 x K2 is the overall stability or formation constant for the


generation of the bis amine adduct [Ag(NH3)2]+(aq):

Ag+(aq) + 2NH3(aq)
[Ag(NH3)2]+(aq)
[Ag(NH 3 ) 2 ]
+
7 −2
Kf = + 2
= 1.6 × 10 M
[Ag] [NH 3 ]

¾ The large value for the Kf indicates that the complex ion product is
strongly favoured.

¾ Most formation constants for complex ions are quite large. This suggests
that these species are very stable compared to the free ligand and
uncomplexed metal ion.
Practice Problem: Complex Ion Formation

Estimate the equilibrium concentrations of Cl-(aq) and


Hg2+(aq) after 0.010 moles of Hg(NO3)2 have been
dissolved in 1.00 L of 0.20 M NaCl(aq).
Dissolving Precipitates via Complex-Ion
Formation
¾ Complex-ion formation results in an increase in the
solubility of precipitates.
e.g. AgCl, which is only sparingly soluble in water, readily
dissolves in 6 M NH3(aq).
AgCl(s)
Ag+(aq) + Cl-(aq) K = 1.8 x 10-10 M2
sp

+ +
[Ag(NH ) ] 7
Ag (aq) + 2NH3(aq)
+

[Ag(NH3)2] (aq) f
+ K = =
3 2
1.6 × 10 M
−2
+ 2
[Ag] [NH ] 3
_________________________________________________________
AgCl(s) + 2NH3(aq)
[Ag(NH ) ]+(aq) + Cl-(aq)
3 2
Ko = Ksp x Kf = 2.9 x 10-3 M-1

Although Ko is small, AgCl will dissolve if the concentration of NH3 is


large.
Sample Problem:

Calculate the equilibrium constant for the reaction

CuS(s) + 4NH3(aq)
[Cu(NH3)4]2+(aq) + S2-(aq)
Thermodynamics

¾ Thermodynamics is the study of the movement of heat


(energy) into or out of the chemical or physical system
under investigation.

¾ In this section of the course we will discover that the heat


of reaction at constant pressure (qp) is frequently different
from its heat at constant volume (qv).

¾ We will investigate why this is so, and study the


fundamental principles of thermodynamics in this
module.
First Law of Thermodynamics

¾ This law states that the energy of the universe is constant


(a.k.a. The Law of Conservation of Energy).

¾ Energy can be converted from one form into another, but it


can neither be created nor destroyed.

In equation form this law can be expressed as

∆UUniverse = ∆Usystem + ∆Usurroundings = 0 ,

Where U = Total Internal Energy(kinetic/potential


energies) for the system or surroundings
Important Definitions:

System: Chemical or physical system under study (e.g. the reacting


molecules or atoms themselves).

Surroundings: Everything else in the universe.

Boundary: Imaginary line that separates the system from


the surroundings.

Kinetic Energy: The energy present in a moving object described by


(Kinetic Energy) K.E. = 1/2mv2 , m = mass, v = velocity.

Potential Energy: Latent energy present in a body/system due to


position or composition (e.g. attractive repulsive forces lead to potential
energy).

Internal Energy(U): Sum of all of the kinetic and potential energies for
all of the particles of the chemical or physical system
What is Chemical Energy?
Consider the reaction

CH4(g) + 2O2(g)
CO2(g) + 2H2O(g) + “Energy”

¾ Since the reaction results in the evolution of heat, it is said


to be exothermic (exo- outside of the system)(i.e. energy is
flowing out of the chemical system and into the
surroundings, therefore ∆H , the enthalpy change for the
system, is negative).

¾ Reactions that absorb energy from the surroundings are


said to be endothermic (endo- inside of the system)
(i.e. energy is flowing from the surroundings into the
chemical system, therefore ∆H , the enthalpy change for the
system, is positive).
e.g. the reaction below requires the absorption of energy from
the surroundings to proceed, and thus is endothermic:

N2(g) + O2(g) + “Energy”


2NO(g)

State Properties and Path Functions

¾ State properties depend only on the state of the system and not how it
was reached.
e.g. State Properties: A man is standing upright on the summit of
Mount Everest. How he arrived at the summit is a path function :
a) Perhaps he parachuted on to the summit from a Boeing 747 jet.
b) Maybe he was carried up to the top of Everest by some Sherpas.
c) It could be that he was beamed to the pinnacle using the transporter
system of the Starship Enterprise.
a, b and c are possible pathways which might have been used to end
up in the “state” of standing erect on Everest.
e.g. 1.0 mol of a gas at 300 K and 2.0 atm is in a definite,
fixed energy state.

If the gas is heated/compressed by a particular pathway to T2 = 440 K


and P2 = 4.00 atm, the state and hence the internal energy for the
system, has been changed.

Change in Internal Energy for a System is given by:

ΔUsys = U final − Uinitial = q + w (Internal Energy Change Equation)

∆U = internal energy change for the system


q = heat flow ; w = work
The internal energy depends upon the state of the system and not
how the system has achieved its internal energy state.

q and w are not state functions, but rather pathways by which the
state achieves a certain internal energy.
Sign Conventions for the Internal Energy Change Equation

∆U = positive (+) : internal energy of the system increases

∆U = negative (-) : internal energy of the system decreases

q = positive (+) : system absorbs heat from the surroundings


(endothermic process)

q = negative (-) : system loses heat to the surroundings


(exothermic process)

w = positive (+) : surroundings do work on the system

w = negative (-) : system does work on the surroundings


¾ Here, the work being referred to is P∆V work, or work at constant
pressure arising from a change in volume for the system:
w sys = − PΔVsys

To recapitulate, q(heat) and w(work) are the means (pathways) by


which a particular energy state for the system is reached.

Pressure-Volume Expansion/Compression Work

¾ The most common type of work encountered in chemical processes is


the work done by a gas (expansion) or to a gas (compression).

¾ Many chemical reactions occur in the open air and thus are subjected to
a constant pressure (i.e. atmospheric pressure)
e.g.
2C8H18(l) + 25O2(g)
16CO2(g) + 18H2O(g) + “Heat” (Fire Bomb)

¾ In this case the energy exchange between the system and the
surroundings is not restricted just to heat flow (qp).

¾ Since 34 moles of gaseous products are formed for every 25 moles of


gaseous reactants, the products occupy a greater volume (under the
constant pressure) than the reactants and thus the combustion
products, the new organizational form for the system, will push out
against the pressure of the surrounding atmosphere.

¾ Because this expansion is done against a constant pressure, the


system is forced to give up some of its internal energy in the form of
work. Thus, the work done by the system to the surroundings will
decrease the amount of energy transferred to the surroundings in the
form of heat (i.e. some of the heat of reaction is converted into work
needed to push the atmosphere out of the way).
Reversible Expansion Work is Given by:

wsys = - Pexternal × ∆V = - Pexternal × (Vf – Vi)

Note: The system does work on the surroundings when ∆V is positive


and thus w is negative.
In the example below, upon reaction, the system does work on the
surroundings due to an increase in volume, and as a result, its internal
energy drops:
Remember: Work = Force × Distance moved = PΔV

Work units: L • atm or


1L • atm = 101 J
for more conventional energy units

Energy versus Enthalpy

¾ For reactions that occur in a closed container, ∆V = 0 and thus wsys = 0


(no P∆V work is possible if the volume of the system remains constant)
So, under constant volume conditions, P∆V = 0, w = 0 and thus
∆Usys = qv (heat flow under constant volume)

¾ Therefore, for any process occurring at constant volume, the internal


energy can be changed only by heat flow (i.e. the work pathway is not
available under constant volume conditions).
¾ By contrast, for reactions that occur in an open container at a constant
pressure there is the possibility of P∆V work if gases are involved and
the system changes volume upon reaction.

¾ Thus, ΔU = q p + − PΔV (under constant P); qp = ∆U + P∆V

¾ By definition, qp = ∆H, so the enthalpy change for a reaction, or the


heat flow at constant pressure is given by

ΔH = ΔU + PΔV
(Enthalpy Change Equation)

Note: ∆H = ∆U when ∆ng = 0 (i.e. when the volume of the system


doesn’t change when reactants are converted to products).
Remember as well that ΔH sys = ΔU sys − w sys

For gaseous reactions:

P∆V = ∆ngRT
(at constant P and constant T, according to the ideal gas law)

∆ng = Moles of gaseous products – Moles of gaseous reactants

¾ These facts lead to an alternate form of the enthalpy equation, which


is actually more useful to us is this course:

ΔH = ΔU + Δn g RT

“Thus, work will be done either to or by the system if the reaction


occurs at constant P (and constant T) and the volume of the system
changes as a result of reaction.”
To sum up:

∆H = ∆U when

1) The reaction does not involve gases


(e.g. reactions in solution that do not form gases,
reactions between solids/liquids).

2) ∆ng = 0 (i.e. number moles of gaseous products =


number of moles of gaseous reactants).
Sample Problem: Calculating Heats of Reaction

An experimental determination of the heat of combustion of


liquid isopropyl alcohol, C3H7OH(l), in a constant volume
calorimeter gave a value of -33.41 kJ/g of C3H7OH(l).
Calculate ∆Ho and ∆Uo for the combustion reaction at
standard conditions (o) per mol C3H7OH(l) burned.
Calorimetry

¾ The science concerned with the energy changes that accompany


chemical or physical transformations is called calorimetry.

¾ Reactions which release heat into the surroundings are called


exothermic reactions (exo-outside of the system).

¾ Reactions which absorb heat from the surroundings are termed


endothermic reactions (endo-inside the system).

¾ Heat can be transferred between the system and the surroundings by


radiation, conduction and convection.

¾ The device used to determine the heat changes (read energy changes)
associated with chemical reactions is called a calorimeter (from the
old energy unit the calorie, which is the energy needed to raise one
gram of water by one degree centrigrade).
¾ Calorimeters measure heat flow by recording the temperature change
that occurs when the calorimeter and its contents absorbs or discharges
energy as heat.

¾ The calorimeter, by virtue of being well insulated, is effectively the


surroundings, while the reacting molecules/atoms/ions themselves
constitute the system.

Heat Capacity Definitions:


General Heat Capacity: The general heat storage capacity of a body
denoted by capital C, is the ratio of heat absorbed divided by the
increase in temperature, or less commonly, the heat discharged divided
by the decrease in temperature:

heat absorbed (J) J J


Cgeneral = increase in temperature (K or o C) ; Units : K or o
C
Specific Heat Capacity: The amount of energy needed to
raise one gram of a substance by one K (or one degree
centigrade), which has been given the symbol small c, is
called the specific heat capacity.

J J
Units : or o
K•g C•g

Molar Heat Capacity: The amount of energy needed to


raise one mole of a substance by one degree centrigrade or
one Kelvin is called the molar heat capacity.

J J
Units : or o
K • mol C • mol
e.g. H2O has a relatively high specific heat capacity of
4.184 J
K•g
while metals have much smaller specific heat capacities.
e.g. Fe(s) = 0.45 J
K•g
(Note 1.000 calorie = 4.184 J)
Measuring the Heat of Solution

¾ The heats of reactions in solution can be measured using a coffee-cup


calorimeter (vide infra):

Coffee-cup calorimeters, since they operate


under constant pressure conditions, provide ∆H.

The procedure involves mixing together


measured amounts of reactants and record the
temperature change of the mixture.
¾ The heat released by an exothermic reaction is absorbed primarily by
the solution, and to a tiny extent, by the calorimeter (also, a small
amount of heat will leak out of the top of the calorimeter).

Heat Lost by the System = Heat gained by the surroundings

(the solution in the cup, excluding the reacting molecules and the cup
itself, constitute the surroundings).

Therefore, qsys = - qsurr

qsurr = (Ctotal) x ∆t ; ∆t = tfinal - tinitial and

Ctotal = Ccalorimeter + Csolution ;

Thus,

qsys = - ( Ccal + Csol’n) x ∆t

(assume that the heat absorbed by the cup itself is negligible)


qsys = - (C sol’n) x ∆t; (only for reactions in cup calorimeters)

⎛J⎞ ⎛ J ⎞
Ccalorimeter = Heat capacity for the empty calorimeter ⎜ ⎟ or ⎜ o ⎟
⎝K⎠ ⎝ C⎠

Csolution = mass of soln’ (g) x specific heat capacity for the


soln’ ⎛⎜ J ⎞⎟ or ⎛⎜ J ⎞⎟
o
⎝K⎠ ⎝ C⎠

¾ Also, it is important to note that in calorimetry problems, it is


customary to report the heat of reaction in kJ/mol of a
particular reactant/product or mol rxn

The equation to calculate the enthalpy change per mol of


limiting reagent is q
ΔH rxn = rxn
# moles of limiting reagent
Practice Cup Calorimetry Problem:

In an experiment designed to measure the heat of


neutralization of a strong acid with a strong base, 33.0 mL
of 1.20 M HCl(aq) were added to 42.0 mL of 1.50 M
NaOH(aq) in a coffee cup calorimeter. The solution
temperatures were originally at 25.0oC. After mixing, the
temperature rose to 31.8oC. Calculate ∆Hneutralization per mol
H2O(l) formed. Assume that the heat absorbed by the
coffee cup itself is negligible, that the specific heat for the
solution formed in the reaction is 4.18 J
o
C• g
g
the density for the final solution is 1.00
mL
and that the total volume is the sum of the component
volumes.
Bomb Calorimetry
¾ Bomb calorimetry involves using a constant volume device to estimate
the internal energy change (∆U) caused by a reaction.

¾ The procedure involves completely combusting a sample under high


pressure molecular oxygen inside a sealed steel bomb. A typical bomb
calorimeter is shown below:

By measuring the temperature change


for the bomb calorimeter (water
and the hardware) and by using
the total heat capacity for the bomb
and its
contents, the heat of reaction is
readily determined.
Sample Bomb Calorimetry Problem:

A bomb calorimeter was heated for 568 seconds with a


1J
15.00 watt heater (1 watt = ). The temperature of the
s
calorimeter rose by 2.00 C. A 0.455-g sample of solid
o

sucrose (C12H22O11) was then burned in excess molecular


oxygen gas in the same calorimeter. The temperature
increased from 24.49oC to 26.25oC. Calculate:
a) the heat capacity for the calorimeter
b) the molar heat of combustion for sucrose
(both ∆Ho and ∆Uo)
c) the number of calories in a gram of sucrose
(Note: 1 calorie = 4.184 J)
Hess’s Law

¾ The enthalpy change for a chemical reaction (∆H) is not dependant upon the
pathway (enthalpy change is a state function).

¾ Thus, the enthalpy change for a given reaction is the same, regardless of
whether the transformation occurs in one step or several steps.

¾ Consequently, it is possible to estimate the enthalpy change for a particular


reaction by adding together other reactions and their associated enthalpies
(the unwanted species must, of course, be cancelled out in the process).

¾ This important principle is known as Hess’s Law

Two Fundamental Properties of Enthalpy Changes

In order to use Hess’s law properly, we must keep in mind the following rules:

1.Reversing the direction of a reaction reverses the sign of ∆H.

2.The magnitude of ∆H is proportional to the quantity of reactants converted to


products. In other words, if you multiply the coefficients in a balanced
equation by an integer, the ∆H must be multiplied by the same integer.
Practice Problem: Hess’s Law:

The heat of combustion of propene gas (C3H6(g)) is


kJ
-2058 mol C3H 6 burned .

Use this value and other tabulated thermodynamics data to


determine the ∆H for the hydrogenation of propene gas to
propane gas (C3H8(g)).
Standard Enthalpies of Formation (another method for finding ∆Ho)

Definition of Std. Enthalpy of Formation (∆Hof): Change in enthalpy


that accompanies the formation of one mole of a compound in its
standard state from its elements in their standard states.

e.g. The reaction for the formation of water from its


elements, with all species in their std. states is

H2(g) + 1/2O2(g)
H2O(l) ∆Hof H2O(l)= -285.8 kJ/mol

Standard State: 1 atmosphere pressure and usually 25oC (273.15 K).

Reactions performed under standard state conditions have a


superscripted (o) after the ∆H.

For gases, the standard state is 1 atmosphere pressure and usually


25oC.

For substances dissolved in a solvent, the standard state is a


concentration of 1 molar.
kJ
¾ Elements in their standard states have ∆Hof = 0
mol
kJ
e.g. Na(s), Hg(l), O2(g) all have ∆Hof = 0 mol .

Standard enthalpies of formation for simple compounds commonly


encountered in first-year chemistry courses are given below:
Sample Problem: Using Std. Enthalpies of Formation to
Determine ∆H

By using the appropriate ∆Hof values from the


thermochemistry tables, calculate ∆Ho for the following
reaction:
SiCl4(l) + 2H2O(l)
SiO2(s) + 4HCl(aq)

Remember that

ΔH o rxn = ∑ ΔH o
f (products) − ∑ ΔH o f (reactants)
Covalent Bond Energies: Use of Bond Energies to
Estimate ∆H Values

¾ Energy changes that accompany chemical reactions arise primarily


from bond energy changes.

¾ For reactions in the gas phase, the heat of reaction can be estimated
by considering which bonds are broken and which are formed during
the course of the reaction. In other words, the use of covalent bond
energies is yet another way in which ∆H can be calculated.

Covalent bonds have two fundamental intrinsic properties:

1. Bond Length: equilibrium distance between the nuclei of the two


bonded atoms.

2. Bond Energy: the energy required to homolytically cleave a


particular bond in the gas phase.

e.g. Cl2(g) + hv
2Cl (g)
¾ Bond energies are positive since energy is needed to
break covalent bonds.

Bond Cleavage is ENDOTHERMIC while


Bond Formation is EXOTHERMIC

¾ Bonds form if the energy of the aggregate of atoms is less


than that of the separated atoms. (Lower energy → more
stable)

¾ Thus the driving force behind the formation of chemical


bonds is the attainment of a lower energy state.
e.g. Consider the formation of the dihydrogen
molecule:

2H(g)
H2(g) + energy

¾ The attraction of each nucleus for the other electron holds


the molecule together.

¾ One mole of H2 molecules is more stable than two moles of


hydrogen atoms by 436 kJ. Thus, the energy required to
break one mole of H-H bonds is + 436 kJ:

H2(g)
2H(g) ∆Ho = + 436 kJ/mol rxn
¾ For all diatomic molecules, the bond energy is the energy
required to split the gaseous molecule into its atoms:

X2(g)
2X(g)

∆Ho = Bond Dissociation Energy

¾ For all polyatomic molecules, each bond has its own


unique bond energy.
¾ For all polyatomic molecules, each bond has its own
unique bond energy.

e.g. H2O : Two O-H bonds

¾ At the start, each O-H bond has the same strength, but
once the first bond is broken, the second becomes easier
to break (energized OH is activated for bond scission):

H2O(g)
H(g) + OH (g) ∆Ho = + 498 kJ/mol rxn
OH(g)
H(g) + O(g) ∆Ho = + 428 kJ/mol rxn
kJ kJ
498 + 428
Thus, Av. OH bond energy for H2O = mol mol kJ
= 463
2 mol
Notes on Bond Energies

¾ Bond energy is an indicator of bond strength: strong


bonds have large bond energies and relatively short
lengths.
e.g. C-C: 346 kJ/mol, 154 pm;
C=C: 615 kJ/mol, 134 pm;
C ≡C: 812 kJ/mol, 120 pm

¾ Bond energies also give an indication of the reactivity of a


molecule.
e.g. N2 molecule is relatively unreactive owing to the great
strength of the N ≡ N triple bond (946 kJ/mol)
To a first approximation, the heat of reaction is the overall
enthalpy change arising from the breaking of bonds (+ve)
in the reactants and the making of bonds in the products
(-ve).
Calculation of Bond Energies

¾ Bond energies can be obtained from thermochemical data


using Hess’s law (only works for gas phase reactions).

e.g. Estimate the energy of an S-F bond


by using the following thermochemical data:

S(s)
S(g) ∆Ho = + 278.8 kJ

F2(g)
2F(g) ∆Ho = + 158 kJ

S(s) + 3F2(g)
SF6(g) ∆Ho = -1209 kJ
Solution:

We want to find ∆Ho for the reaction

S-F(g)
S(g) + F(g)

(enthalpy change is the energy present in one mole of


S-F bonds)

Note: All species must end up in the gas phase for this
procedure to be valid.
This is simply another application of Hess’s Law:

1. S(s)
S(g) ∆Ho1 = + 278.8 kJ/mol
+
2. 3 x (F2(g)
2F(g) ) ∆Ho2 = 3 x 158 kJ/mol
+
3. S(g) + 6F(g)
SF6(g) ∆Ho3 = negative
_____________________________________________
4. S(s) + 3F2(g) SF6(g) ∆Ho4 = -1209 kJ/mol

∆Ho4 = ∆Ho1 + ∆Ho2 + ∆Ho3 ;


278.8 kJ + 474 kJ + ∆Ho3 = -1209 kJ
∆Ho3 = −1961.8 kJ ; + 1961.8 kJ kJ
= 327
Thus, average S-F bond energy = 6 moles of S-F bonds mol
Calculating the Heat of Reaction from Bond Energies

¾ The ∆H for a chemical reaction can also be calculated via the bond
energies for the constituent bonds in the reactants and products. This
procedure only works with all species in the gas phase and when all
reactant molecules are broken apart into their constituent atoms. Product
molecules are then assembled from the appropriate free atoms in the gas
phase (vide infra):

Atoms
Bond Scission Bond Formation

(positive delta H) (negative delta H)

Reactant Product
Molecules ∆Ho rxn Molecules
ΔH o rxn = ∑ Bond Energies for Bonds Broken in the Reactants −

∑ Bond Energies for Bonds Formed in the Products

¾ If bond breakage consumes more energy than is liberated


from bond formation, ∆Ho rxn is positive.

¾ If bond breakage consumes less energy than is liberated


from bond formation, ∆Ho rxn is negative.
Sample Problem: Bond Energies

a) Calculate the enthalpy of atomization for cyclopropane


gas (vide infra):
H H
C
H
H C C
H H
(Note ∆Hof cyclopropane gas = 52.7 kJ/mol)

b) Use your answer to part a) and the C-H bond energy


value on your data sheet to calculate the average C-C
bond energy for cyclopropane.
Spontaneous Change: Entropy and Free Energy

¾ In this section we will investigate the Second Law of Thermodynamics,


which gives us a criterion for predicting if a given process is
thermodynamically spontaneous.

¾ A thermodynamically spontaneous process has a natural tendency to


proceed to equilibrium without being driven by an external force (e.g.
heat, pressure).

¾ All spontaneous processes have a preferred direction. e.g. gas flows


from high to low pressure, water freezes naturally below 0oC. As we
shall see, all spontaneous processes lead to a decrease in the “free
energy” of the system.

¾ Spontaneous processes can be fast or slow. The Second Law of


Thermodynamics (pursuit of minimum energy) tells us the direction in
which a reaction will occur, but not the rate of the process (e.g. diamond
spontaneously converts to graphite at standard conditions but at an
imperceptibly slow rate).
¾ Central question: “What drives a chemical reaction/physical
change in a particular direction?”
Most spontaneous reactions are exothermic.

¾ So, one of the driving forces that makes a reaction


spontaneous is the tendency of the system to give off heat.

¾ There are, however, spontaneous processes that absorb


energy form the surroundings (i.e. some processes have
∆H = + ve)

e.g. H2O(l) boils spontaneously at 100oC under 1 atm


pressure even though the process is endothermic:

H2O(l)
H2O(g) ∆H = + 41 kJ/mol
¾ Thus, the evolution of heat cannot be the only driving
force. There must be another factor that determines
whether or not a process is spontaneous.

¾ This factor is called entropy, which is a measure of the


dilution of energy (randomness) in the system.

¾ Entropy, which has the symbol capital S, is a qualitative


measure of the amount of dilution of energy in the system
or the surroundings. The dilution of energy is manifested
as an increase in chaos for the system or the
surroundings.

VERY IMPORTANT: ALL SPONTANEOUS PROCESSES


RESULT IN AN INCREASE IN THE ENTROPY OF THE
UNIVERSE.
Entropy

¾ The entropy of a system depends upon the number of


energy states available to it. Energy is more effectively
diluted as the number of energy states available to the
system increases. (i.e. as the number of occupied energy
states goes up, the entropy gets larger)

¾ The number of available energy states increases when a)


the temperature is raised, b) the reaction results in an
increase in the number of particles and c) the volume of
the system increases.

¾ At absolute zero (0 K), the entropy of a crystalline solid is


zero (absolutely no energy has been diluted and the
particles in the lattice are perfectly ordered).
¾ The steady increase in entropy with heating is caused by the random jiggling of
the particles in the solid. The abrupt increases in entropy that occur when the
solid melts and the liquid vapourizes are a result of an increase in the volume
of the system, which gives the particles more freedom of motion. Because the
particles are spread out in a greater volume, their energies are more effectively
diluted.(vide infra):

Note that ∆Svap > ∆Smelt .


This is because movement
from a liquid to a gas results
in a much greater increase in
the volume for the system,
and hence, much more
effective energy dilution.
¾
Processes That Have Positive ∆S(∆Ssys) Values

1. Increase in the volume of a gas at constant P

2. Increase in the number of moles of gas in a reaction

3. Increase in temperature at constant V

4. Dissolving of a liquid or solid in a solvent

5. Melting, vapourization, sublimation

6. Any process that increases the number of particles


Entropy Changes for Phase Transitions
¾ Melting and vapourization are equilibrium processes (at the melting
point and boiling point temperatures, respectively) and thus the
entropy change for these processes at constant P is
q rev
ΔSsys = (General Equation)
T

¾ More specifically,
ΔH fus
ΔSfus =
Tm.p.
(ΔH is the molar enthalpy of fusion; Tm.p. is the melting point (K))
and
ΔH vap
ΔSvap =
Tb.p.
(ΔH is the molar enthalpy of vapourization; Tb.p. is the boiling point (K))

¾In both cases, energy must be added to bring about the phase change
and this energy comes from the surroundings.
Since these are equilibrium processes,

∆Ssys = -∆Ssurr and ∆Suniverse = 0

Consequently,

the equation q rev


ΔSsys =
T

holds true only for reversible, isothermal processes


(i.e. at transition temperatures for phase changes, there is
no net change in the entropy for the entire universe)
Practice Problem: Phase Transitions-Entropy Changes

Calculate the entropy change that occurs when 3.00 mol of


benzene vapourizes at its normal boiling point of 80.1oC.
The molar enthalpy of vapourization for benzene at 80.1oC
is 30.8 kJ/mol.
Standard Molar Entropies

¾ So is the symbol for the amount of entropy present in a mole of a pure


substance at 298.2 K and at 1.00 atmosphere pressure.
J
Units:
K• mol

¾ All standard molar entropies are positive because the temperature


for the system is 298.2 K, and to get to this temperature the system
must absorb (read dilute) energy starting from 0 K , which makes the
entropy greater than zero.

¾ Standard molar entropies increase with increasing molar mass


and increasing numbers of atoms in the constituent molecules.

¾ Also, the more ordered the substance is, the less effectively it
dilutes energy, and hence the smaller is its standard molar
entropy.
e.g. Consider the three allotropes of carbon, viz.
J J
Graphite So = 5.74 ; Diamond So = 2.42
K• mol K• mol
J
Buckyball (C60) So = 7.30
K• mol

Of the three, diamond is the poorest at diluting energy owing to its rigid,
highly-ordered internal structure;

Diamond Buckyball Graphite


Standard Reaction Entropies

¾ The standard reaction entropy is the entropy change for


the reaction in which reactants in their standard states are
converted to products in their standard states.

Entropy Change Equation:

ΔSo rxn = ∑ So
(products) − ∑ (reactants)
So

¾ This is the equation that we must use to calculate entropy


changes for chemical processes.
Sample Problem: Entropy Changes for Chemical Reactions

Calculate the entropy changes for the following chemical


reactions:

a) F2(g)
2F(g)

b) CO(g) + Cl2(g)
COCl2(g)
The Second Law of Thermodynamics

¾ Spontaneous processes are those for which there is an increase in the


total entropy for the universe.

¾ For a spontaneous irreversible process:

∆Suniverse = ∆Ssystem + ∆Ssurroundings > 0

¾ This does not mean that the change in the entropy for the system or
its surroundings cannot be negative. To be spontaneous, these two
terms simply must sum to a value greater than zero.

¾ For reversible processes, the total entropy change for a system and
its surroundings is zero.

e.g. A water ice-liquid water mixture at 273.15 K and 1 atm is a


reversible system and hence the ∆Suniverse for the process at
273.15 K is zero.
Entropy Change for the Surroundings

q surr − q sys −ΔH sys


ΔSsurr = = = ( at constant P)
T T T
¾ If heat flows out of the system, it must flow into the surroundings and
vice versa. i.e. qsys = - qsurr

¾ So, heat flow into the surroundings increases the entropy of the
surroundings and vice versa.

General Rules for Predicting Spontaneity

1. Exothermic reactions (∆H < 0) are often, but not always


spontaneous.
2. Reactions that result in an increase in the entropy of the system
(∆S > 0) are often, but not always spontaneous.
¾ Thus, to decide in which direction a chemical reaction will
naturally proceed, we need to consider both the enthalpy
and entropy changes for the system.

Gibbs Free Energy Change (∆G)

¾ Next we want to develop an equation that will allow us to


focus on the system alone to decide if a process is
spontaneous:

∆Suniverse (or ∆Stotal ) = ∆Ssystem + ∆Ssurroundings

−ΔH sys ΔH sys


and ΔSsurr = , thus ΔStotal = ΔSsys −
T T
Rearrangement and multiplication of both sides by –T gives

−TΔStotal = ΔH - TΔS ; − TΔStotal = ΔG

ΔG = ΔH − T Δ S
(Gibbs-Helmholtz Equation)

¾ The beauty of this equation is that it allows us to predict the direction of


the spontaneous reaction by considering only the entropy and
enthalpy changes for the system (enthalpy and entropy changes to
the surroundings can be ignored). In other words, the Gibbs-Helmholtz
equation takes into account temperature, enthalpy and entropy
contributions and allows for a quick prediction of spontaneity via the
∆G value (i.e. do the enthalpy and entropy changes for a reaction at a
given temperature give a negative or positive ∆G)
¾ When ∆G is negative, ∆Suniverse is positive, and thus the
process is spontaneous in the forward direction (from
L→R). In other words, the direction of a spontaneous
process is in the direction of decreasing free energy for the
system.

¾ Note that ∆G cannot be measured directly, but rather is


calculated from the ∆S and ∆H values for the reaction.

¾ ∆G represents the total amount of free energy available


from the reaction that can do work on the surroundings or
the minimum amount of work needed to drive a non-
spontaneous reaction to completion.
Remember:

∆G < 0 : Reaction is spontaneous in the


forward direction

∆G > 0 : Reaction is spontaneous in the


reverse direction

∆G = 0 : The system is at equilibrium; no


net change in free energy occurs
¾ It is important to note that while ∆G is clearly temperature dependent
(see plot directly below), ∆H and ∆S are not affected significantly by
changes in temperature (i.e. to a first approximation, ∆H and ∆S are
temperature insensitive).
CONDITIONS THAT FAVOUR SPONTANEOUS PROCESSES

ENTHALPY ENTROPY SPONTANEOUS?


CONTRIBUTION CONTRIBUTION
(∆H) (-T∆S) (∆G)
__________________________________________________________
Exothermic YES AT ALL TEMPS
(∆H < 0) ∆S > 0 (∆G IS ALWAYS
NEGATIVE)

Exothermic YES AT LOW TEMPS


(∆H < 0) ∆S < 0 WHEN TΔS < ΔH

Endothermic YES AT HIGH TEMPS


(∆H > 0) ∆S > 0 WHEN TΔS > ΔH

Endothermic NO AT ALL TEMPS


(∆H > 0) ∆S < 0 (∆G IS ALWAYS POSITIVE)
Note: (∆Go = negative) is the free energy released when the system
moves from standard state ( Q = 1, all gases 1 atm ; all dissolved
species 1M concentration) to equilibrium. The temperature does not
have to be
298.2 K!

Significance of the Sign and Magnitude of ∆Go

¾ Since ∆Go is calculated for products/reactants in their standard


states, viz. 1 M for any solute and gases at 1atm, the reaction
quotient Q must equal one.

¾ Thus, ∆G = ∆Go if standard state is the equilibrium state

¾ The sign of ∆Go indicates the direction in which the reaction


must proceed to reach equilibrium, starting at std. state ( Q = 1).
¾ If ∆Go < 0, then Goproducts < Goreactants

Thus, the reaction will proceed to the right (forward direction) to bump
up the concentration of products at the expense of reactants to make

∑ G o
Products = ∑ Reactants
G o

At equilibrium K will be greater than one (product concentrations are


larger than reactant concentrations)

¾ If ∆Go > 0, then Goproducts > Goreactants

Thus, the reaction will proceed to the left (reverse direction) to bump
up the concentration of reactants at the expense of products to make

∑ G o
Products = ∑ Reactants
G o

At equilibrium K will be less than one (product concentrations are


smaller than reactant concentrations)
¾ If ∆Go = 0, then Goproducts = Goreactants

At equilibrium K = 1. Special equilibrium position where


products and reactants are preferred equally. This is the
case where standard state is equilibrium.

¾ The smaller the absolute value of ∆Go, the closer


standard state is to equilibrium.

¾ The larger the absolute value of ∆Go, the further the


reaction has to go from standard state to reach
equilibrium.
¾ The relationship between the free energy change and the direction of the
spontaneous reaction to equilibrium is outlined in the diagram below:
Sample Problem: Free Energy Changes and Spontaneity

Calculate the sign of the entropy and enthalpy


changes for the following reactions and use them to
predict which are

a) Spontaneous at all temperatures


b) Non-spontaneous at all temperatures
c) Spontaneous at high temperatures (entropy driven)
d) Spontaneous at low temperatures (enthalpy driven)

2NO2(g)
N O (g)
2 4

2NaF(s)
2Na(s) + F2(g)

C5H12(l) + 8O2(g)
5CO2(g) + 6H2O(g)
Estimating the Temperature at Which an Enthalpy/ Entropy-
Driven Reactions Becomes Thermodynamically Possible

¾ A reaction is thermodynamically possible if ∆Go is negative.

¾ If ∆Ho and ∆So for the reaction are known, we can estimate the transition
temperature, at which the reaction reverses spontaneity.

¾ At the transition temperature , the point of transition from non-


spontaneous to spontaneous, equilibrium is standard state.

Thus, at this point ∆Go = 0 , So,


∆Go = ∆Ho - T∆So = 0 at Ttrans ; Rearrangement gives

ΔH o
Ttrans =
ΔSo
VERY IMPORTANT: ∆Go = ∆G = 0 ONLY AT Ttrans for entropy or
enthalpy-driven reactions.
At Ttrans , K =1 and reactants/products are preferred equally
Estimating ∆Go at Temperatures Other than 25oC

¾ Simply assume that ∆Ho and ∆So do not change with


temperature and continue to use the Gibbs-Helmholtz
equation
∆Go = ∆Ho - T∆So

Unlike ∆Ho and ∆So, ∆Go is strongly dependent upon


temperature
(i.e. ∆Go is not just for 298.2 K!)
Sample Problem: Estimating Ttrans and ∆Go
at a Temperature Other Than 25oC

For the reaction

CH4(g) + H2O(g)
3H2(g) + CO(g)

A) Calculate ∆So, ∆Ho at 25oC and ∆Go at 600 K


B) Estimate Ttrans if it exists
Standard Free Energy of Formation and Reaction

¾ The standard free energy of reaction is the difference


between the free energies of formation for the reactants in
their standard states and the free energies of formation for
the products in their standard states. (i.e. ∆Go is the free
energy change that occurs when reactants in their standard
states are converted to products in their standard states):

ΔG o rxn = ∑ f
ΔG o
(products) − ∑ f (reactants)
ΔG o

and
ΔG o f = ΔH o f − T Δ So f
Definition: ∆Gof is the change in standard free energy
that accompanies the formation of one mole of a
substance in its standard state from its elements in their
standard states (usually at 25oC). Tabulated values on the
data sheet are for 25oC.

Note: ∆Gof , ∆Hof and ∆Sof for a pure element in its


standard state is zero!

Important: “A compound is thermodynamically stable


with respect to its elements if its free energy of
formation is negative”
Practice Problem: Calculating Standard
Free Energies of Formation
kJ
1. For PH3(g), ∆Ho f is + 5.4
mol
J
and So = 210 K• mol .

Calculate ∆Gof for PH3(g) at 25oC.

2. Is PH3(g) thermodynamically stable at 25oC/ 1 atm?


Free Energy and Concentration: ∆G at Non-Standard
Conditions

¾ Definition: ∆Gnon-standard is the free energy change that occurs when a


non-standard state combination of reactants and products is converted
into an equilibrium mixture. Non-standard mixtures have a least one
reactant or product with a concentration or partial pressure not equal to
1M/1 atm.

Direction of the Spontaneous Reaction to Equilibrium

¾ To determine the direction that a given reaction will proceed to


equilibrium, we need to compare the instantaneous reaction quotient,
Q, to the equilibrium constant K:

Q < K , the reaction will proceed from L→ R to reach equilibrium

Q > K , the reaction will proceed from R→ L to reach equilibrium

Q = K , the system is at equilibrium already, therefore no net rxn will


occur
Free Energy Criteria for Reaction Spontaneity

∆Gnon-std < 0, the reaction is spontaneous from L → R

∆Gnon-std > 0, the reaction is spontaneous from R → L

∆Gnon-std = 0, the mixture is at equilibrium, no net rxn will occur

∆Go = ∆Gnon-std = 0, equilibrium is standard state

Note: Non-standard free energy change is ∆Gnon-std not ∆Go !

The sign of ∆Go gives the direction of spontaneity starting at standard


conditions.

The driving force and the direction of the reaction depend upon the
concentrations of reactants/products (i.e. upon Q) and also on the
magnitude/sign of the ∆Gnon-std for the reaction.
¾ As the reaction proceeds, the instantaneous value of
∆Gnon-std varies as Q changes.

The non-standard free energy change is given by

ΔG non-std = ΔG o + RTlnQ
J
R = 8.3145 K• mol ; K is the temperature in Kelvins

lnQ is the natural logarithm of the reaction quotient

¾ As the reaction proceeds from reactants to products, the increasing


product concentrations result in an increasing value for Q, ∆Gnon-std will
become less negative and thus the driving force for the forward reaction
will decline in strength.
Relationship Between ∆Go and the Equilibrium Constant

¾ At equilibrium, ∆Gnon-std = 0 and Q = K. (∆Go = 0 only when


equilibrium is standard state)

¾ Therefore, non-standard free energy change equation at


equilibrium has the form

∆Gnon-std = 0 = ∆Go + RTlnK

¾ Rearrangement to isolate ∆Go on one side gives the


equation that relates the equilibrium constant to the standard
free energy change:
ΔG o = − RT ln K
¾ For reactions involving only gases K = KP

¾ For reactions involving solutes in solution K = Kc

¾ For reactions involving both solutes in solution and


gases K = Keq

Why is ∆Gnon-std = 0 at equilibrium?


Answer: The minimum energy state (and thus the target
state) for the system occurs where
Free energies for the reactants = Free energies for the
products
or when ∆Gnon-std = 0 kJ/mol
Practice Problem: Non-Standard Free Energy Changes

For the reaction…

2NOCl(g)
2NO(g) + Cl2(g)

KP = 2.7 x 10-2 atm at 773.2 K

1. Calculate ∆Go at 773.2 K


2 .Calculate ∆Gnon-std at 773.2 K when each gas
initially has a partial pressure 0.0050 atm
Temperature Dependence of the Equilibrium Constant

Replacing the ∆Go term in ∆Go = -RTlnK with ∆Ho - T∆So and solving
for K gives

ΔH o ΔSo
ln K = − +
RT R
(Van’t Hoff Equation)

Application of this equation to a given reaction at two different


temperatures, T1 and T2, gives two different equations with two
equilibrium
; constants, K1 and K2. The subtraction of one ln equation
from the other affords the Van’t Hoff Two-Point Equation, which relates
the ratio of the rate constants to the two temperatures and the enthalpy
change for the reaction:

⎛ K 2 ⎞ ΔH o ( T2 − T1 )
ln ⎜ ⎟= × Where T2, T1 are temperatures in Kelvin; K2 and K1 are

⎝ K1 ⎠ R ( T2 × T1 ) equilibrium constants for the two temperatures


T2 and T1 ;
kJ
R = 8.3145 × 10−3
K• mol
(Van’t Hoff Two-Point Equation)
∆Ho = Std. Molar Enthalpy change for the reaction
¾ If K2 > K1 when T2 > T1 , then the reaction is endothermic
(+ ve ∆H)

¾ If K2 < K1 when T2 > T1 , then the reaction is exothermic


(-ve ∆H)

Finally, consider a reaction with ∆H = + ve and ∆S = -ve.


Even though ∆Go becomes increasingly more positive with
increasing T, the equilibrium constant increases as well
because the T increases faster than ∆Go and this makes K
larger at higher temperatures.
−ΔG o
Recall K = e ΔG o
RT ; gets smaller with increasing T
RT
Sample Problem That Involves the Use of the Van’t Hoff
Two-Point Equation

For the reaction…

N2O4(g)
2NO2(g)

KP = 0.13 atm at 298 K. Calculate KP at 350 K.


Electrochemistry

Definition: Study of the interchange of chemical energy and


electrical energy (chemistry and electricity both involve electrons).

¾ Electrochemistry is based primarily upon processes that involve


oxidation-reduction reactions.

¾ The two types of processes that we will study in this course are:

1. Spontaneous chemical reactions that generate an


electric current
2. Non-spontaneous reactions that require the use of
an external current to produce chemical change.

¾ Let us begin by describing redox reactions, which form the


foundation of electrochemistry.
Oxidation-Reduction Reactions

¾ Oxidation-reduction reactions involve the complete transfer


of electrons from one chemical species to another.

¾ Most of the redox reactions that we will encounter take


place in aqueous solution.

¾ Recall from CHEM 101/CHEM 103 that


oxidation involves the loss of electrons and concomitant
increase in the oxidation state for the oxidized atom.

¾ Reduction, on the other hand, involves the gain of electrons


and a concomitant decrease in the oxidation state for the
reduced atom.
Balancing Redox Equations by the Half-Reaction
Method

I) Reactions in Acidic Solution

¾ First remove all spectator ions from the original equation

¾ Separate the net ionic equation into a reduction half-reaction and


oxidation half-reaction

¾ Balance each half reaction by using the following priority order:


1. Start with all atoms except H and O
2. Balance for O by adding H2O to the side deficient in O
3. Balance for H by adding H+ to the side deficient in H
4. Balance the charge with electrons
5. Find the lowest common denominator for the two half
reactions
6. Multiply each half reaction with the appropriate integer
7. Add the half reactions and cancel out the electrons
Sample Problem: Balancing Redox Reaction in Aqueous
Solution

Balance the following reaction by the half-reaction method:

Cr2O72-(aq) + C2O42-(aq)
Cr2+(aq) + CO2(g)
II) Reactions in Basic Solution
Same basic approach as with balancing in aqueous
solution, but with two additional steps:

1. Neutralize any H+ ions in the equation by adding an


equal number of OH- ions to both sides of the equation,
then cancel out any duplications of H2O(l) formed by the
reaction H+(aq) + OH-(aq) H2O(l)

Sample Problem: Balancing Redox Reactions in Basic


Solution

Balance the redox reaction below in basic solution:

MnO4-(aq) + N2H4(aq)
MnO2(s) + N2(g)
Galvanic Cells

¾ A galvanic cell is device, built by humans, in which a spontaneous


redox reaction is harnessed to provide electron flow through an
external circuit.

¾ In essence a galvanic cell, or battery, converts chemical energy into


electrical energy that can be used to do work.

¾ The individual oxidation-reduction half-reactions take place in


separate compartments called half-cells

¾ The circuit for the cell is closed via a conducting wire and a salt
bridge or semi-permeable membrane.

¾ One of the earliest galvanic cells assembled was the Zn-Cu battery.
This particular cell is based upon the following redox reaction:

Cu2+(aq) (1.00 M) + Zn(s) Zn2+(aq)(1.00 M) + Cu(s)


At the instant of hook up, with a sodium sulfate bridge in place, this
standard cell has a voltage of 1.103 V. A diagram of this cell is shown
below:
¾ The left half-cell in the diagram is the anodic compartment. Oxidation
occurs in the anodic compartment. For all galvanic cells, the anode is
the negative electrode, and it is usually connected to the voltmeter by a
black wire. So, in this case we have

Anode : Oxidation half-reaction taking place in the compartment:

Zn(s) Zn2+(aq) + 2e-

¾ The right half-cell in the diagram is the cathodic compartment.


Reduction occurs in the cathodic compartment. For all galvanic cells,
the cathode is the positive electrode, and it is usually connected to the
voltmeter by a red wire. So, in this case we have

Cathode : Reduction half-reaction taking place in the compartment:

Cu2+(aq) + 2e- Cu(s)

Overall Cell Rxn: Cu2+(aq) + Zn(s) ↔ Zn2+(aq) + Cu(s)


¾ To recapitulate, in this particular cell, in the anode, Zn metal is oxidized
to Zn2+(aq) ions, providing two electrons that are transported through
the connecting wire to the Cu metal strip in the cathodic compartment.
Cu2+(aq) ions are then reduced to Cu(s) on the surface of the cathode
by the two electrons that were supplied by the Zn anode. The net effect
is that energized electrons are pumped through the battery circuit via
the driving force behind the spontaneous redox reaction upon which
the cell is based. These energized electrons can then be employed to
do work, such as power a flashlight.

¾ Remember: “Anions flow into the anode, while cations flow into the
cathode”.

¾ These ions can pass between the two half cells either through a salt
bridge or a porous membrane.

¾ Without a salt bridge or membrane, the cathode would become highly


negatively charged (excess of anions), while the anode would become
highly positively charged (excess of cations) and this would shut down
the cell. A salt-bridge or membrane allows the ions to migrate from one
compartment to the other, thereby keeping the net charge in each cell
equal to zero.
Standard Cell Notation

¾ Galvanic cells are usually reported in standard cell notation (vide infra):

Anode/Anode Electrolyte//Cathode Electrolyte/Cathode

¾ This notation allows one to avoid providing a detailed diagram each


time reference is made to the cell.

¾ A single slash represents a phase boundary while a double slash


indicates a salt bridge or membrane.

¾ Only reacting or current carrying materials are included.

¾ Pt metal is the electrode material whenever the half-cell involves only


ions in solution or a gas ( e.g. for 2H+(aq) + 2e- H2(g) ; Pt is the
electron conduit)

Concentrations of the ions involved in the reaction are usually provided


e.g.
Zn/Zn 2+ (aq) (1.00 M)//Cu 2+ (aq) (1.00 M)/Cu
Electrical Work and Free Energy

¾ The difference in the electrical potential between the cathode and the
anode of a voltaic cell creates an electromotive force that drives
electrons through the external circuit.

electrical work (joules)


volts = ∴ one volt = one joule per coulomb
electrical charge (coulombs)
J
1V = 1
C

¾ The potential (voltage) can be measured with a voltmeter and expressed


in volts (named after Alessandro Volta). Thus, voltage is a measure of
the average amount of energy present in each electron that is passed
through the circuit.
¾ Thus, voltage is a measure of the average amount of energy present in
each electron that is passed through the circuit.

¾ The maximum electrical work that can be done during the transfer of Q
coulombs of charge by an electromotive force of E volts is given by

Maximum electrical work = ΔG = − Q × E


(i.e. the number of electrons passed through the circuit
multiplied by the force with wihich each electron is
driven through the circuit)

Sample Problem: Electrical Work:


Calculate the maximum electrical work done by a 12.0-volt battery that
supplies a current of 10.0 amps for 1.00 minute.
∆G and Cell Voltage

¾ The maximum useful work that can be generated by a spontaneous


reaction is ∆G

¾ Therefore,
Maximum work = ∆G = − Q × E

Also, Q = n×F (n = # moles of electrons,

F = Faraday constant = charge of 1 mol of e- = 96,485 coulombs);


Thus

ΔG = − n × F × E
(E = max . possible potential for the cell at std. conditions )
Spontaneous Redox Reactions

¾ For spontaneous reactions in voltaic cells, ∆G < 0 and E > 0 (i.e. a


positive cell potential indicates a spontaneous reaction)

Standard States and Standard Cell Voltages

¾ A standard cell is one in which all substances are in their standard


states.(ions at 1M; gases at 1 atm pressure and 25oC)

¾ The standard cell voltage (Eo) is a quantitative


measure of the tendency of the reactants in their standard states to
form products in their std. states

Thus, at std. conditions , ΔG o = − n F E o


Sample Problem: Voltage and Free Energy

The standard voltage of the cell


Zn/Zn2+ (1.00 M)//Cu2+( 1.00 M)/Cu
at the instant of hook-up is 1.10 V. Calculate ∆G for one
mole of zinc metal being dissolved in the cell reaction
(Note: once the current begins to flow, the voltage for the
cell starts to drop).
Standard Electrode (Half-Cell) Potentials

¾ It is not possible to measure the absolute potential of a


single half-cell (electrode).

¾ However, we can assign a voltage to each half-reaction on


the basis of the sum of the overall cell voltage.

¾ Standard half-cell potentials have been arbitrarily


referenced to the standard hydrogen electrode (SHE),
which has been assigned a value of zero volts.

i.e. Pt/H2(g) (1 atm)/H+(aq) ( 1 M )


with the half-cell reaction
2H+(aq) + 2e- H2(g) (Eo = 0 V by definition)
¾ By convention, because the reduction of H+(aq) was set at 0 V, all half-
cell potentials are given as standard reduction processes in the
standard table (vide infra):
e.g. Zn2+(aq) + 2e- Zn(s) ; Eo = -0.76 V ,

thus this half-reaction is non-spontaneous relative to the SHE

Cu2+(aq) + 2e- Cu(s) ; Eo = + 0.34 V ,

thus this half-reaction is spontaneous relative to the SHE


¾ Combining two half reaction to give a balanced full-cell
equation requires two manipulations:

1. One of the reduction half-reactions must be reversed.


Therefore, the sign of the potential must also be
reversed. The half-reaction with the more negative
reduction potential is the one that must be reversed.

2. Since the number of electrons lost must equal the


number gained, the half reactions must be multiplied by
the appropriate integers as necessary to balance the
equation.

VERY IMPORTANT: The value for Eo must never be


changed ! Eo must not be multiplied by the integer used to
balance the equation! Voltage is an intensive property,
meaning it does not depend upon the amount of
substance. It is a measure of the difficulty with which
electrons are added to or removed from a particular atomic
centre. (electronegativity is another intensive property)
Calculating Standard Cell Voltage from Standard Electrode Potentials

¾ Spontaneous reactions in voltaic cells proceed in the direction that


gives a positive cell potential and hence a negative ∆Go.

¾ Therefore, when two half-cells are added to make a full cell,


1. the half-cell with the more positive reduction potential Eo
proceeds as a reduction, making it the cathode
2. while the half-cell with the more negative reduction potential Eo
proceeds as an oxidation, making it the anode.
The standard cell potential for a voltaic cell is given by:

E o cell = E o (cathode)red − E o (anode) red


(1/2 rxn with the (1/2 rxn with the
more "+" red. pot.) more "−"red. pot.)

(Just plug in the reduction potentials directly


from the table)
Practice Problem: Calculating Cell Potentials

The voltaic cell


Pt/H2(g)(1.00 atm)/H+(aq) (1.00 M)//Os2+ (aq)(1.00 M)/Os
has a voltage of 0.44 V. By using the std. reduction
potential for the H2/H+ couple, calculate the std. reduction
potential for the Os2+/Os couple.
Relative Strengths of Oxidizing Agents and Reducing
Agents

¾ The standard reduction potentials table has the following


format:

Oxidizing agent + ne- Reducing agent

¾ More positive Eo means a more negative ∆Go, therefore,

¾ The more positive the standard reduction potential of a


couple, the stronger is the oxidizing agent and the weaker is
the reducing agent in the couple.

¾ Conversely, the more negative the standard reduction


potential of a couple, the weaker is the oxidizing agent and
the stronger is the reducing agent in the couple.
All spontaneous redox reactions follow the format:

First Conjuage Redox Pair

Stronger Oxidizing Agent + Stronger Reducing Agent Weaker Reducing Agent + Weaker Oxidizing Agent

Second Conjugate Redox Pair

Note: The cell reaction with the highest positive potential is the one that will
occur preferentially, unless it is kinetically-controlled.

Sample Problem: Spontaneous Std. Cell Reactions


Give two species that are capable of oxidizing Sn metal to Sn2+, but not
all the way to Sn4+. Include the appropriate cell reactions and their
associated voltages in your answer.
Disproportionation Reactions

¾ Disproportionation reactions are chemical


transformation in which a single species is both
oxidized and reduced.

¾ The reacting atomic centre in a disproportionation


reaction is in an intermediate oxidation state
between higher and lower oxidation states that are
known for the element.

e.g. 2Cu+(aq) Cu2+(aq) + Cu(s)


(I) (II) (0)
Addition of Two Half-Reactions to Obtain a Third Half
Reaction (all of the same type)

¾ When two half reactions are obtained to give a third half-reaction, all of
the same type, the potential for the third half-reaction is not the sum of
the potentials for the first two, but rather the weighted average of the first
two potentials. This is the case because the half-reactions are all of the
same type (no electrons are cancelled out) voltages are blended
proportionately rather than simply added together.

¾ To find Eo for the third half-reaction, we need to use free energy


changes, which are additive.

i.e. ΔG o 3 = ΔG o1 + ΔG o 2
e.g. Determine the reduction potential for
Cu2+(aq) + 2e- Cu(s) given…

Cu2+(aq) + e- Cu+(aq) Eored = 0.153 V and


Cu+(aq) + e- Cu(s) Eored = 0.521 V

Solution:

a) Cu2+(aq) + e- Cu+(aq) Eored = 0.153 V


+
b) Cu+(aq) + e- Cu(s) Eored = 0.521 V
___________________________________________
c) Cu2+(aq) + 2e- Cu (s) Eored = ? V (weighted average)
Δ G o c = Δ G o a + ΔG o b

∆Go = -nFEo ; So,

−nFE o c = − nFE o b + − nFE o a

−2FE o c = − 1F(0.153 V) + − 1F(0.521 V)

Dividing both sides by -2F gives

Eoc =
( −1 × 0.153 V + − 1 × 0.521 V )
= 0.337 V (weighted average)
−2
Practice Sample Problem: Calculating Half –Cell Potentials/Latimer
Diagrams

A Latimer diagram is a reduction series for a particular element that


shows the potential differences between the various stable oxidation
states. Given below is a Latimer diagram for a series of nitrogen
species:
?V

- 0 . 94 V 0 . 98 V
NO3 HNO2 NO
(V) (III) (II)

a) Write the equation for the NO3-/NO couple in acidic solution and
calculate its reduction potential.
b) Write a balanced equation for the disproportionation of HNO2. Is the
disproportionation of HNO2 spontaneous at std. conditions/25oC?
Cell Potential and Concentration: The Nernst Equation

¾ Up until now, we have been using std. reduction potentials


and cell voltages at std. conditions and 25oC and thus we
have not needed to consider other concentrations.

¾ Reaction potentials do, however, depend upon reactant and


product concentrations.

¾ In fact, measuring cell voltage is a method of determining


ion concentration or gas partial pressure.

e.g. pH meter measures [H+] indirectly by measuring the


voltage of the test solution.
Qualitative Effect of Concentration on Cell Voltage

¾ How concentration affects voltage can be understood by applying Le


Chatelier’s principle to the cell reaction.

e.g. consider the electrochemical cell based on

H2(g) + 2AgCl(s) 2Ag(s) + 2H+(aq) + 2Cl-(aq)

Applying Le Chatelier’s principle we would predict that

1. An increase in H2 pressure will drive the reaction to the


right, thereby making the cell voltage more positive.

2. An increase in [H+] or [Cl-] will drive the reaction back to the left,
thereby making the cell voltage less positive.

3. Changing the quantity of AgCl(s) does not affect the ion


concentrations and thus changing the mass of AgCl(s) in the cell has
no effect on the cell potential.
¾ Similarly, for the redox couple

Ox + ne- Red

The reduction potential becomes more positive with increasing [Ox] and
decreasing [Red].

Nernst Equation for a Redox Reaction

¾ Recall that ∆G non-standard is given by

ΔG = ΔG o + RTlnQ , combining this with


and ΔG o = − nFE o and ΔG = − nFE
we get
− nFE o + RTlnQ = − nFE
¾ Rearranging to get E by itself on one side we end up with the
Nernst Equation (second term on the RHS accounts for any deviation
from std. state):

RT
E = E − o
ln Q where
nF

E = non-standard voltage (V)


Eo = std. voltage for all species in their std. states
F = Faraday constant (96, 485 coulombs of charge per mole of electrons)
T = Temperature in K
Q = Reaction quotient (initial ion concentrations and initial gas pressure)
n = # of moles of electrons transferred in the balanced equation

¾ This powerful equation gives the voltage for non-standard cells and
half-cells (i.e. cells in which at least one concentration or pressure is
different from 1M or 1 atm).
¾ By combining all of the constants, fixing the temperature
at T = 298 K and converting ln into log base 10 we get the
more commonly used version of the Nernst equation,
shown below:

0.0592
E = E − o
logQ (for 298 K only!)
n

Note: The Nernst equation work for both half-cells and


full cells. Also remember, if Q = 1 , E = Eo.
Sample Problems Using The Nernst Equation

1. Calculate the electrode potential versus the SHE at


25oC for the following electrode reaction:
Pt/H+(aq) (pH = 4.00)/ H2 (g) (1.00 atm)

2. Calculate the cell voltage at 25oC for the following cell:


Al(s)/Al3+(aq) (0.020 M)//Sn2+(aq)(0.10 M)/Sn(s)
Concentration Cells

¾ A concentration cell is an electrochemical cell in which the


anode/cathode have the same oxidant/reductant pair, but the ion
concentration in the two half-cells is different, and this difference
creates a positive potential. See, for instance, the Cu/Cu2+
concentration cell below:
¾ The Nernst equation predicts that a voltage should exist for
such a cell due solely to unequal concentrations.

¾ Voltage develops from the natural tendency of two


solutions of different concentrations to achieve the same
concentration (i.e. the concentration gradient provides the
voltage)

¾ Concentration cells can be used to measure ion


concentrations and to determine equilibrium constants.

Measuring Ion Concentrations with Electrochemical Cells

¾ If we know the standard potential of a reaction and


measure a non-standard voltage, we can calculate Q and
solve for the unknown concentration of a reactant or
product in solution, which can help us to find a Kc, Kp, Ka ,
Kb, Ksp or Kf value.
Concentration Cell Question
Consider the following cell:
Ag(s)/Ag+(aq) (0.10 M)//Ag+(aq) (1.0 M)/Ag(s)
By using the Nernst equation, calculate the cell potential caused by the
concentration gradient.

Solution:

As usual, we first must write down the cell reaction:

Cathode: Ag+(aq) (1.0 M ) + e- Ag(s)


+
Anode : Ag(s) Ag+(aq) (0.10 M) + e-

Ag+(aq) (1.0 M) (Cathode) Ag+(aq) (0.10 M) (Anode)


Eo = 0 V (True for all concentration cells) :
Ag+(aq) (1 M) (Cathode) Ag+(aq) (1 M) (Anode)

(Cell reaction is from higher conc. ½ cell to lower conc. ½ cell)

¾ Electrons flow from the anode into the cathode; this will cause an
increase in [Ag+] in the anode and a decrease [Ag+] in the cathode.

⎡⎣ Ag + (aq )(anode) ⎤⎦
n = 1e- ; Q =
⎡⎣ Ag + (aq )(cathode) ⎤⎦

0.0592 ⎛ [0.10 M ] ⎞
E = 0.000V − log ⎜ ⎟V
1 ⎝ [1.0 M ] ⎠
E = 0.0592 V
Sample Problem: Concentration Cells
Given that the initial voltage of the cell below is 1.41 V at
25oC, calculate the Ksp for Tl3PO4.
Tl(s)/Tl+(aq) ( x M)//Tl+(aq) (1.00 M)/Tl(s) (Assume that the
anodic solution is saturated Tl3PO4(aq) produced by
dissolving solid Tl3PO4 in pure water)
Calculating Keq Values from Standard Potentials

¾ At equilibrium, Q = K, E = O V, which corresponds to a dead cell (i.e.


the redox reaction that forms the basis of the electrochemical cell has
gone to completion).
Therefore,
RT
0V = E − o
ln K eq
nF
By combining constants and by converting natural log to log base 10
we get at 25oC:

n
log K = Eo
0.0592
(relates std. cell voltage to equilibrium constant at 25oC)
Practice Problem: Std. Voltage and the Equilibrium
Constant

The cell below had a voltage of 0.118 V at the instant of hook-up.


Calculate Kc for the reaction at 25oC.

3Pu(s) + 2Ir3+(aq)(0.221 M) 2Ir(s) + 3Pu2+(aq)(0.331 M)


Electrolytic Cells

¾ Oxford dictionary definition: The splitting of chemical


entities with electricity.

¾ Chemical dictionary definition: The use of direct current to


accomplish a non-spontaneous redox reaction.

¾ The process of electrolysis converts electrical energy into


chemical energy.
¾ Electrolysis is performed in an electrolytic cell. A typical electrolytic cell
is given below:

¾ The voltage for this cell at the instant of hook-up is – 1.10 V. (The
voltage for an electrolytic cell is always negative)

¾ The Zn/Zn2+ electrode is the cathode, the Cu/Cu2+ electrode the anode
¾ The negative terminal of the power source sends electrons
to the cathode while electrons from the anode are pulled to
the positive terminal of the power source.

IMPORTANT: The electrode reactions are the reverse of


those in voltaic cells and thus the electrode polarities are
reversed. The cathode is negative since it is connected to
the negative terminal while the anode is positive because
it is attached to the positive terminal.

Cell Cathode Anode

Voltaic + -

Electrolytic - +
¾ Also, keep in mind that in both voltaic and electrolytic cells,
reduction occurs at the cathode and oxidation occurs at the anode.

¾ To force electrons in the direction opposite to that of the spontaneous


voltaic cell requires the application of an external D/C voltage greater
than the negative potential of the electrolytic cell (more positive than
1.10 V for the cell above).
i.e. for current to flow, Eapplied > - Eelectrolytic cell

¾ In practice, much more than the minimum voltage is needed for


electrolysis to begin because of the natural resistance of the cell (see
overvoltage below)

¾ Finally, it must be mentioned that electrolysis requires that there are


sufficient numbers of ions present in solution to allow for the current to
flow through the cell.

e.g. If Pt electrodes are placed in pure water and hooked up to a D/C


power source, no electrolysis will be observed because pure water
contains insufficient ions to allow for the flow of current at moderately
high voltages. To start the electrolysis, a small amount of an ionic
compound must be added.
Overvoltage

¾ Additional voltage over and above the EMF for the voltaic cell is required
to initiate electrolysis.

¾ The major reason for the need to use higher voltage is kinetic
overpotential, which is the extra voltage needed to overcome the
activation energy for each half-reaction and drive the electrons across
the electrode surface at a reasonable rate.

Electrolysis of Molten Salts

¾ Ion movement necessary for current flow can be achieved by melting


ionic solids and using the melt as the electrolyte.

¾ Compounds of highly reactive metals typically exist as oxide salts here


on the Earth.

¾ The most practical way to obtain pure samples of these metals is to


electrolyze their metal oxides at high temperatures.
e.g. Hall-Heroult electrolytic process for producing aluminum
metal:
molten cryolite

2Al2O3(s) + 3C(s) 4Al(s) + 3CO2(g)


External current

¾ Bauxite or Al2O3 is dissolved in molten Na3AlF6 (cryolite)


(ca. 1400oC), which serves as the solvent.

¾ Al3+ ions are reduced at the cathode, while molecular


oxygen reacts with graphite at the anode, forming CO2(g)

¾ Thus, the anode is consumed in the reaction (i.e. converted


into CO2(g)) and must therefore, constantly be replaced with
fresh graphite.
Electrolysis of Aqueous Ion Mixtures: Predicting Electrode
Reactions

¾ Consider for example, that a solution containing Cu2+(aq), Ag+(aq),


Zn2+(aq) is to be electrolyzed by gradually increasing the voltage.

One question that arises is: “In what order would the metals plate out
into the cathode?”

To answer this, we have to look at standard reduction potentials for


these three ions:

1. Ag+(aq) + e- Ag(s) Eo red = 0.80 V

2. Cu2+(aq) + 2e- Cu(s) Eo red = 0.34 V



3. Zn2+(aq) + 2e- Zn(s) Eo red = - 0.76 V

¾ The reaction with the more positive Eo has the greatest tendency to
proceed.
¾ This is the correct order but only because of the high
overpotential associated with the reduction of H2O(l).

¾ Electrolysis in aqueous solution is complicated by the fact


that water can be reduced at the cathode or oxidized at the
anode.

Reduction of H2O(l)

2H2O(l) + 2e- H2(g) + 2OH-(aq)


Ered = - 0.414 V (pH = 7)

Oxidation of H2O(l)

2H2O(l) O2(g) + 4H+(aq) + 4e-


- Ered = - 0.815 V (pH = 7)
Upshot?

¾ Theoretically, it is not possible to reduce a cation in pH = 7 solution that


has an Eored more negative than -0.414 V.

¾ However, due to the high overvoltage for the reduction of liquid water,
the actual reduction of pure liquid H2O is - 0.814 V.

¾ Thus, one can actually reduce metal ions with Eored as low as – 0.80 V
in liquid water.

¾ Similarly, one cannot, in principle, oxidize a species in aqueous solution


with an oxidation potential(-Eored) more negative than – 0.815 V, but
because of overvoltage, the actual empirical oxidation of pure liquid
water is –1.415 V.

¾ Thus, because of overpotential, it is possible to oxidize dissolved


species in liquid water with oxidation potentials as low as ca. –1.4 V.
Sample Problem: Predicting the Products of an Aqueous
Electrolysis Reaction

A 1.0 M solution of CdF2(aq) is electrolyzed between two


inert Pt electrodes. Predict which reactions will occur at the
cathode and anode and write equations for each half-
reaction, the full cell reaction and calculate the cell voltage.
Quantitative Aspects of Electrolysis

¾ The extent to which a reaction takes place in an


electrochemical cell is directly proportional to the quantity
of electrical charge passed through the cell.

e.g. To completely reduce 1 mol of Al3+ ions to Al,


3 moles of electrons must be passed through the cell.

Quantity of charge (q) present in 1 mol of electrons is


called the Faraday C
1 Faraday = 96,485 mol e- ; Remember as well that
⎛C ⎞
q = I⎜ ⎟ × t (s )
⎝s ⎠
Sample Problems: Quantitative Aspects of Electrolysis

1. Elemental rhenium (Re) can be produced by the


electrolysis of molten ReCl3.

a) What mass of Re would be produced at the


cathode if a current of 6.5 x 103 amps
is applied for 48 hours?

b) Calculate the total amount of electrical work


needed to accomplish the electrolytic process
described above if the applied voltage is 5.00 V.
2. Determine the oxidation number for the chlorine atom in
a sodium salt of an oxychloroanion of unknown formula
if electrolysis of a molten sample of the salt for
1.50 h with a current of 10.0 amps produces 3.97 g of
molecular chlorine gas at the cathode.

Vous aimerez peut-être aussi