17.5
|
a.
|
This is a weak acid problem. Setting up the standard equilibrium
table:
|
CH3COOH(aq)
|

|
H+(aq)
|
+ CH3COO-(aq)
|
Initial
(M):
|
0.40
|
|
0.00
|
0.00
|
Change
(M):
|
-x
|
|
+x
|
+x
|
Equilibrium
(M):
|
(0.40 - x)
|
|
x
|
x
|


x
= [H+] = 2.7 ´
10-3
M
pH
= 2.57
|
b.
|
In addition to the
acetate ion formed from the ionization of acetic acid, we also have
acetate ion formed from the sodium acetate dissolving.
CH3COONa(aq) ® CH3COO-(aq) + Na+(aq)
Dissolving 0.20 M sodium acetate initially
produces 0.20 M CH3COO-
and 0.20 M Na+. The sodium ions are not involved in any
further equilibrium (why?), but the acetate ions must be added to the
equilibrium in part (a).
|
CH3COOH(aq)
|

|
H+(aq)
|
+ CH3COO-(aq)
|
Initial
(M):
|
0.40
|
|
0.00
|
0.20
|
Change
(M):
|
-x
|
|
+x
|
+x
|
Equilibrium
(M):
|
(0.40 - x)
|
|
x
|
(0.20 + x)
|


x
= [H+] = 3.6 ´
10-5
M
pH =
4.44
Think About
It:
|
Could you have predicted whether the pH should have
increased or decreased after the addition of the sodium acetate to the
pure 0.40 M acetic acid in
part (a)?
An alternate way to work
part (b) of this problem is to use the Henderson-Hasselbalch equation.


|
|
|
|
17.6
|
a.
|
This is a weak base problem.
|

|

|

|
|

|
Initial (M):
|
0.20
|
|
|
0
|
0
|
Change (M):
|
-x
|
|
|
+x
|
+x
|
Equilibrium (M):
|
0.20 - x
|
|
|
+x
|
+x
|


x
= 1.9 ´
10-3
M =
[OH-]
pOH =
2.72
pH
= 11.28
|
b.
|
The initial concentration of NH is 0.30 M from the salt NH4Cl. We set up a table as in part (a).
|

|

|

|
|

|
Initial (M):
|
0.20
|
|
|
0.30
|
0
|
Change (M):
|
-x
|
|
|
+x
|
+x
|
Equilibrium (M):
|
0.20 - x
|
|
|
0.30 + x
|
+x
|


x
= 1.2 ´
10-5
M =
[OH-]
pOH =
4.92
pH
= 9.08
Alternatively, we could use the Henderson-Hasselbalch
equation to solve this problem. We
get the value of Ka
for the ammonium ion using the Kb
for ammonia (Table 16.7) and Equation 16.8. Substituting into the
Henderson-Hasselbalch equation gives:

pH =
9.25 -
0.18 = 9.07
Think About
It:
|
Is
there any difference in the Henderson-Hasselbalch equation in the cases
of a weak acid and its conjugate base and a weak base and its conjugate
acid?
|
|
|
|
17.9
|
Strategy:
|
What constitutes a buffer
system? Which of the solutions
described in the problem contains a weak acid and its salt (containing
the weak conjugate base)? Which
contains a weak base and its salt (containing the weak conjugate
acid)? Why is the conjugate base
of a strong acid not able to neutralize an added acid?
|
Solution:
|
The criteria for a buffer
system are that we must have a weak acid and its salt (containing the
weak conjugate base) or a weak base and its salt (containing the weak
conjugate acid).
a.
|
HCl (hydrochloric acid) is a strong acid. A buffer is a solution containing
both a weak acid and a weak base.
Therefore, this is not
a buffer system.
|
b.
|
H2SO4
(sulfuric acid) is a strong acid.
A buffer is a solution containing both a weak acid and a weak
base. Therefore, this is not a buffer system.
|
c.
|
This solution contains both a
weak acid, H2PO and its
conjugate base, HPO .
Therefore, this is a buffer system.
|
d.
|
HNO2 (nitrous
acid) is a weak acid, and its conjugate base, NO (nitrite ion, the anion of the salt
KNO2), is a weak base.
Therefore, this is a buffer system.
|
Only (c) and (d) are buffer systems.
|
|
|
17.10
|
a.
|
HCN is a weak acid, and its conjugate base, CN-,
is a weak base. Therefore, this is a buffer system.
|
b.
|
HSO is a weak
acid, and its conjugate base, SO is a weak
base. Therefore, this is a buffer system.
|
c.
|
NH3 (ammonia) is a
weak base, and its conjugate acid, NH is a weak
acid. Therefore, this is a buffer system.
|
d.
|
Because HI is a strong acid,
its conjugate base, I-, is an extremely weak base. This means that the I- ion will not combine
with a H+ ion in solution to form HI. Thus, this system cannot act as a
buffer system.
|
|
|
17.11
|
Strategy:
|
The pH of a buffer system
can be calculated using the Henderson-Hasselbalch equation (Equation
17.3). The Ka of a conjugate acid such as NH is
calculated using the Kb
of its weak base (in this case, NH3) and Equation 16.8.
|
Solution:
|
NH (aq) NH3(aq) + H+(aq)
Ka =
Kw /1.8 ´
10-5
= 5.56 ´
10-10
pKa = -log
Ka = 9.26
pH = 8.89
|
|
|
17.12
|
a.
|
We summarize the concentrations of the species at
equilibrium as follows:
|
CH3COOH(aq)
|

|
H+(aq)
|
+ CH3COO-(aq)
|
Initial
(M):
|
2.0
|
|
0
|
2.0
|
Change
(M):
|
-x
|
|
+x
|
+x
|
Equilibrium
(M):
|
2.0 - x
|
|
x
|
2.0 + x
|


Ka =
[H+]
Taking the -log
of both sides,
pKa =
pH
Thus, for a buffer system in
which the [weak acid] = [weak base],
pH =
pKa
pH =
-log(1.8
´
10-5) =
4.74
|
b.
|
Similar to part (a),
pH
= pKa = 4.74
Buffer (a) will be a more effective buffer
because the concentrations of acid and base components are ten times
higher than those in (b). Thus,
buffer (a) can neutralize 10 times more added acid or base compared to
buffer (b).
|
|
|
17.13
|
H2CO3(aq) HCO (aq) + H+(aq)







|
|
17.14
|
Step 1: Write the equilibrium that occurs between
H2PO and HPO . Set up a
table relating the initial concentrations, the change in concentration to
reach equilibrium, and the equilibrium concentrations.
|
|

|
|

|
Initial
(M):
|
0.15
|
|
0
|
0.10
|
Change
(M):
|
-x
|
|
+x
|
+x
|
Equilibrium
(M):
|
0.15
-
x
|
|
x
|
0.10
+ x
|
Step 2: Write the ionization constant expression
in terms of the equilibrium concentrations.
Knowing the value of the equilibrium constant (Ka), solve for x.

You can look up the Ka
value for dihydrogen phosphate (Ka2
for phosphoric acid) in Table 16.8 of your text.


x
= [H+] =
9.3 ´
10-8
M
Step 3: Having solved for the [H+], calculate the pH of the
solution.
pH
= -log[H+] = -log(9.3
´
10-8) = 7.03
|
|
17.15
|
Using the Henderson-Hasselbalch
equation:


Thus,

|
|
17.16
|
We can use
the Henderson-Hasselbalch equation to calculate the ratio [HCO3-]/[H2CO3]. The Henderson-Hasselbalch equation is:

For the
buffer system of interest, HCO3- is the
conjugate base of the acid, H2CO3. We can write:


The [conjugate base]/[acid] ratio
is:


The buffer should be more
effective against an added acid because ten times more base is present
compared to acid. Note that a pH of
7.40 is only a two significant figure number (Why?); the final result
should only have two significant figures.
|
|
17.17
|
For the first
part we use Ka for
ammonium ion. (Why?) The Henderson-Hasselbalch
equation is

For the second part, the acid-base
reaction is
NH3(g) + H+(aq)
® NH (aq)
We find the number of moles
of HCl added

The number of moles of NH3
and NH4+ originally present are

Using the acid-base reaction, we find the number of moles of NH3
and NH after
addition of the HCl.
|

|

|

|
|
Initial
(mol):
|
0.013
|
0.0010
|
|
0.013
|
Change
(mol):
|
-0.0010
|
-0.0010
|
|
+0.0010
|
Final
(mol):
|
0.012
|
0
|
|
0.014
|
We find the new pH:

|
|
17.18
|
As calculated in Problem 17.12, the pH of this buffer system is
equal to pKa.
pH
= pKa = -log(1.8 ´
10-5) = 4.74
a.
|
The added NaOH will react
completely with the acid component of the buffer, CH3COOH. NaOH ionizes completely; therefore,
0.080 mol of OH- are added to the
buffer.
Step 1: The neutralization reaction is:
|
CH3COOH(aq)
|
+ OH-(aq)
|
→
|
CH3COO-(aq)
|
+ H2O(l)
|
Initial
(mol):
|
1.00
|
0.080
|
|
1.00
|
|
Change
(mol):
|
-0.080
|
-0.080
|
|
+0.080
|
|
Final
(mol):
|
0.92
|
0
|
|
1.08
|
|
Step 2: Now, the acetic acid equilibrium is
reestablished. Since the volume of
the solution is 1.00 L, we can convert directly from moles to molar
concentration.
|
CH3COOH(aq)
|

|
H+(aq)
|
+ CH3COO-(aq)
|
Initial
(M):
|
0.92
|
|
0
|
1.08
|
Change
(M):
|
-x
|
|
+x
|
+x
|
Equilibrium
(M):
|
0.92 - x
|
|
x
|
1.08 + x
|
Write the Ka
expression, then solve for x.


x
= [H+] =
1.5 ´
10-5
M
Step 3: Having solved for the [H+], calculate the pH of the
solution.
pH =
-log[H+] =
-log(1.5
´
10-5) =
4.82
The pH of the buffer increased from 4.74 to 4.82 upon
addition of 0.080 mol of strong base.
|
b.
|
The added acid will react
completely with the base component of the buffer, CH3COO-. HCl ionizes completely; therefore, 0.12
mol of H+ ion are added to the buffer
Step 1: The neutralization reaction is:
|
CH3COO-(aq)
|
+ H+(aq)
|
→
|
CH3COOH(aq)
|
Initial
(mol):
|
1.00
|
0.12
|
|
1.00
|
Change
(mol):
|
-0.12
|
-0.12
|
|
+0.12
|
Final
(mol):
|
0.88
|
0
|
|
1.12
|
Step 2: Now, the acetic acid equilibrium is
reestablished. Since the volume of
the solution is 1.00 L, we can convert directly from moles to molar
concentration.
Step 2: Now, the acetic acid equilibrium is
reestablished. Since the volume of
the solution is 1.00 L, we can convert directly from moles to molar
concentration.
|
CH3COOH(aq)
|

|
H+(aq)
|
+ CH3COO-(aq)
|
Initial
(M):
|
1.12
|
|
0
|
0.88
|
Change
(M):
|
-x
|
|
+x
|
+x
|
Equilibrium
(M):
|
1.12 - x
|
|
x
|
0.88 + x
|
Write the Ka
expression, then solve for x.


x
= [H+] =
2.3 ´
10-5
M
Step 3: Having solved for the [H+], calculate the pH of
the solution.
pH =
-log[H+] =
-log(2.3
´
10-5) =
4.64
The pH of the buffer decreased from 4.74 to 4.64 upon
addition of 0.12 mol of strong acid.
|
|
|
17.19
|
In order for
the buffer solution to function effectively, the pKa of the acid component must be close to the
desired pH. We write


Therefore,
the proper buffer system is Na2A/NaHA.
|
|
17.20
|
For a buffer
to function effectively, the concentration of the acid component must be
roughly equal to the conjugate base component. According to Equation 17.3 of the text,
this occurs when the desired pH is close to the pKa of the acid, that is, when pH »
pKa,

or

To prepare a
solution of a desired pH, we should choose a weak acid with a pKa value close to the
desired pH. Calculating the pKa for each acid:
For
HA, pKa = -log(2.7 ´
10-3) =
2.57
For
HB, pKa = -log(4.4
´
10-6) = 5.36
For
HC, pKa = -log(2.6
´
10-9) =
8.59
The buffer solution with a pKa closest to the desired
pH is HC. Thus, HC is the
best choice to prepare a buffer solution with pH =
8.60.
|
|
17.21
|
1.
|
In order to function as a buffer, a solution must
contain species that will consume both acid and base. Solution (a)
contains HA-, which can consume
either acid or base:
HA- +
H+ ® H2A
HA- +
OH- ® A2- +
H2O
and A2-,
which can consume acid:
A2- +
H+ ® HA-
Solution (b) contains HA-,
which can consume either acid or base, and H2A, which can
consume base:
H2A +
OH- ® HA- +
H2O
Solution (c) contains only H2A and
consequently can consume base but not acid. Solution (c) cannot function as a
buffer.
Solution (d) contains HA-
and A2-.
Solutions (a), (b), and (c) can
function as buffers.
|
2.
|
Solution (a)
should be the most effective buffer because it has the highest
concentrations of acid- and base-consuming species.
|
|
|
17.22
|
1.
|
Solution (d) has
the lowest pH because it has the highest ratio of HA to A-
(6:4). Solution (a) has the highest pH because it
has the lowest ratio of HA to A-
(3:5).
|
2.
|
After addition of two H+ ions to solution (a), there
will be two species present:
HA and A-.
|
3.
|
After
addition of two OH- ions to solution
(b), there will be two species
present: HA and A-.
|
|
|
17.27
|
Since the
acid is monoprotic, the number of moles of KOH is equal to the number of
moles of acid.


|
|
17.28
|
We want to
calculate the molar mass of the diprotic acid. The mass of the acid is given in the
problem, so we need to find moles of acid in order to calculate its molar
mass.

The
neutralization reaction is:
2KOH(aq) + H2A(aq)
K2A(aq) + 2H2O(l)
From the
volume and molarity of the base needed to neutralize the acid, we can
calculate the number of moles of H2A reacted.

We know that
0.500 g of the diprotic acid were reacted (1/10 of the 250 mL was
tested). Divide the number of grams
by the number of moles to calculate the molar mass.

|
|
17.29
|
The neutralization
reaction is:
H2SO4(aq) + 2NaOH(aq) ® Na2SO4(aq) + 2H2O(l)
Since one mole of sulfuric
acid combines with two moles of sodium hydroxide, we write:


|
|
17.30
|
We want to calculate the
molarity of the Ba(OH)2 solution. The volume of the solution is given (19.3
mL), so we need to find the moles of Ba(OH)2 to calculate the
molarity.

The neutralization reaction is:
2HCOOH
+ Ba(OH)2 ® (HCOO)2Ba + 2H2O
From the volume and
molarity of HCOOH needed to neutralize Ba(OH)2, we can determine
the moles of Ba(OH)2 reacted.

The molarity of the Ba(OH)2 solution is:

|
|
17.31
|
a.
|
Since the acid is monoprotic, the moles of acid equals the moles
of base added.
HA(aq) + NaOH(aq) NaA(aq) + H2O(l)

We know
the mass of the unknown acid in grams and the number of moles of the
unknown acid.

|
b.
|
The number of moles of NaOH in 10.0 mL of solution is

The neutralization reaction is:
|
HA(aq)
|
+ NaOH(aq)
|
→
|
NaA(aq)
|
+ H2O(l)
|
Initial (mol):
|
0.00116
|
6.33 ´ 10-4
|
|
0
|
|
Change (mol):
|
-6.33 ´ 10-4
|
-6.33 ´ 10-4
|
|
+6.33 ´ 10-4
|
|
Final (mol):
|
5.3 ´ 10-4
|
0
|
|
6.33 ´ 10-4
|
|
Now, the weak acid equilibrium will be reestablished. The total volume of solution is 35.0 mL.


We can calculate the [H+] from the pH.
[H+] =
10-pH = 10-5.87
= 1.35 ´ 10-6 M
|
HA(aq)
|

|
H+(aq)
+
|
A-(aq)
|
Initial (M):
|
0.015
|
|
0
|
0.0181
|
Change (M):
|
-1.35 ´ 10-6
|
|
+1.35 ´ 10-6
|
+1.35 ´ 10-6
|
Equilibrium (M):
|
0.015
|
|
1.35 ´ 10-6
|
0.0181
|
Substitute the equilibrium
concentrations into the equilibrium constant expression to solve for Ka.

|
|
|
17.32
|
The resulting
solution is not a buffer system.
There is excess NaOH and the neutralization is well past the
equivalence point.


|
CH3COOH(aq)
|
+
NaOH(aq)
|
→
|
CH3COONa(aq)
|
+
H2O(l)
|
Initial (mol):
|
0.0500
|
0.0835
|
|
0
|
|
Change (mol):
|
-0.0500
|
-0.0500
|
|
+0.0500
|
|
Final (mol):
|
0
|
0.0335
|
|
0.0500
|
|
The volume of the resulting
solution is 1.00 L (500 mL + 500 mL = 1000 mL).




|
CH3COO-(aq)
|
+ H2O(l)
|

|
CH3COOH(aq)
|
+ OH-(aq)
|
Initial
(M):
|
0.0500
|
|
|
0
|
0.0335
|
Change
(M):
|
-x
|
|
|
+x
|
+x
|
Equilibrium
(M):
|
0.0500
-
x
|
|
|
x
|
0.0335
+ x
|


x = [CH3COOH] = 8.4 ´
10-10 M
|
|
17.33
|
HCl(aq) + CH3NH2(aq) CH3NH (aq) +
Cl-(aq)
Since the concentrations of acid and base are equal, equal
volumes of each solution will need to be added to reach the equivalence
point. Therefore, the solution
volume is doubled at the equivalence point, and the concentration of the
conjugate acid from the salt, CH3NH3+, is:

The conjugate acid
undergoes hydrolysis.
|

|

|

|
|

|
Initial (M):
|
0.10
|
|
|
0
|
0
|
Change (M):
|
-x
|
|
|
+x
|
+x
|
Equilibrium (M):
|
0.10
-
x
|
|
|
x
|
x
|
|
|
|
|
|
|
|


Assuming that, 0.10 -
x » 0.10
x
= [H3O+] =
1.5 ´
10-6
M
pH
= 5.82
|
|
17.34
|
Let's assume we react 1 L of HCOOH with 1 L of NaOH.
|
HCOOH(aq)
|
+ NaOH(aq)
|

|
HCOONa(aq)
|
+ H2O(l)
|
Initial
(mol):
|
0.10
|
0.10
|
|
0
|
|
Change
(mol):
|
-0.10
|
-0.10
|
|
+0.10
|
|
Final
(mol):
|
0
|
0
|
|
0.10
|
|
The solution volume has doubled
(1 L + 1 L = 2 L). The concentration
of HCOONa is:

HCOO-(aq) is a weak base. The hydrolysis is:
|
HCOO-(aq)
|
+ H2O(l)
|
→
|
HCOOH(aq)
|
+ OH-(aq)
|
Initial
(M):
|
0.050
|
|
|
0
|
0
|
Change
(M):
|
-x
|
|
|
+x
|
+x
|
Equilibrium
(M):
|
0.050
-
x
|
|
|
x
|
x
|


x =
1.7 ´
10-6
M
= [OH-]
pOH
= 5.77
pH
= 8.23
|
|
17.35
|
The reaction between CH3COOH and KOH is:
CH3COOH(aq) + KOH(aq) ® CH3COOK(aq) + H2O(l)
We see that 1 mole CH3COOH is stoichiometrically
equivalent to 1 mol KOH. Therefore,
at every stage of titration, we can calculate the number of moles of acid
reacting with base, and the pH of the solution is determined by the excess
acid or base left over. At the
equivalence point, however, the neutralization is complete, and the pH of
the solution will depend on the extent of the hydrolysis of the salt
formed, which is CH3COOK.
a.
|
No KOH has been added.
This is a weak acid calculation.
|
CH3COOH(aq)
|
+ H2O(l)
|

|
H3O+(aq)
|
+ CH3COO-(aq)
|
Initial
(M):
|
0.100
|
|
|
0
|
0
|
Change
(M):
|
-x
|
|
|
+x
|
+x
|
Equilibrium
(M):
|
0.100 - x
|
|
|
x
|
x
|


x
= 1.34 ´
10-3
M =
[H3O+]
pH
= 2.87
|
b.
|
The number of moles of CH3COOH originally present
in 25.0 mL of solution is:

The number of moles of KOH in 5.0 mL is:

We work with moles at this
point because when two solutions are mixed, the solution volume
increases. As the solution volume
increases, molarity will change, but the number of moles will remain the
same. The changes in number of
moles are summarized.
|
CH3COOH(aq)
|
+ KOH(aq)
|
→
|
CH3COOK(aq)
|
+ H2O(l)
|
Initial
(mol):
|
2.50 ´ 10-3
|
1.00 ´ 10-3
|
|
0
|
|
Change
(mol):
|
-1.00 ´ 10-3
|
-1.00 ´ 10-3
|
|
+1.00 ´ 10-3
|
|
Final
(mol):
|
1.50 ´ 10-3
|
0
|
|
1.00 ´ 10-3
|
|
At this stage, we have a
buffer system made up of CH3COOH and CH3COO-
(from the salt, CH3COOK).
We use the Henderson-Hasselbalch equation to calculate the pH.


pH = 4.56
|
c.
|
This part is solved similarly to part (b).
The number of moles of KOH in 10.0 mL is:

The changes in number of moles are summarized.
|
CH3COOH(aq)
|
+ KOH(aq)
|
→
|
CH3COOK(aq)
|
+ H2O(l)
|
Initial
(mol):
|
2.50 ´ 10-3
|
2.00 ´ 10-3
|
|
0
|
|
Change
(mol):
|
-2.00 ´ 10-3
|
-2.00 ´ 10-3
|
|
+2.00 ´ 10-3
|
|
Final
(mol):
|
0.50 ´ 10-3
|
0
|
|
2.00 ´ 10-3
|
|
At this stage, we have
a buffer system made up of CH3COOH and CH3COO-
(from the salt, CH3COOK).
We use the Henderson-Hasselbalch equation to calculate the pH.


pH = 5.34
|
d.
|
We have reached the equivalence
point of the titration. 2.50 ´
10-3
mole of CH3COOH reacts with
2.50 ´
10-3
mole KOH to produce 2.50 ´ 10-3
mole of CH3COOK. The
only major species present in solution at the equivalence point is the
salt, CH3COOK, which contains the conjugate base, CH3COO-. Let's calculate the molarity of CH3COO-. The volume of the solution is: (25.0 mL + 12.5 mL = 37.5 mL = 0.0375
L).

We set up the hydrolysis of CH3COO-,
which is a weak base.
|
CH3COO-(aq)
|
+ H2O(l)
|

|
CH3COOH(aq)
|
+ OH-(aq)
|
Initial
(M):
|
0.0667
|
|
|
0
|
0
|
Change
(M):
|
-x
|
|
|
+x
|
+x
|
Equilibrium
(M):
|
0.0667 - x
|
|
|
x
|
x
|


x
= 6.09 ´
10-6
M =
[OH-]
pOH =
5.22
pH
= 8.78
|
e.
|
We have passed the equivalence
point of the titration. The excess
strong base, KOH, will determine the pH at this point. The moles of KOH in 15.0 mL are:

The changes in number of moles are summarized.
|
CH3COOH(aq)
|
+ KOH(aq)
|
→
|
CH3COOK(aq)
|
+ H2O(l)
|
Initial
(mol):
|
2.50 ´ 10-3
|
3.00 ´ 10-3
|
|
0
|
|
Change
(mol):
|
-2.50 ´ 10-3
|
-2.50 ´ 10-3
|
|
+2.50 ´ 10-3
|
|
Final
(mol):
|
0
|
0.50 ´ 10-3
|
|
2.50 ´ 10-3
|
|
Let's calculate the molarity of the KOH in
solution. The volume of the
solution is now 40.0 mL =
0.0400 L.

KOH is a strong base. The pOH is:
pOH =
-log(0.0125) =
1.90
pH = 12.10
|
|
|
17.36
|
The reaction between NH3 and HCl is:
NH3(aq) +
HCl(aq) ® NH4Cl(aq)
We see that
1 mole NH3 is stoichiometrically equivalent to 1 mol HCl. Therefore, at every stage of titration,
we can calculate the number of moles of base reacting with acid, and the pH
of the solution is determined by the excess base or acid left over. At the equivalence point, however, the
neutralization is complete, and the pH of the solution will depend on the
extent of the hydrolysis of the salt formed, which is NH4Cl.
a.
|
No HCl has been added.
This is a weak base problem.


x
= 2.3 ´
10-3
M =
[OH-]
pOH =
2.64
pH
= 11.36
|
b.
|
The number of moles of NH3 originally present in
10.0 mL of solution is:

The number of moles of HCl in 10.0 mL is:

We work with moles at
this point because when two solutions are mixed, the solution volume
increases. As the solution volume
increases, molarity will change, but the number of moles will remain the
same. The changes in number of
moles are summarized.
|
NH3(aq) +
|
HCl(aq)
|
→
|
NH4Cl(aq)
|
Initial
(mol):
|
3.00 ´ 10-3
|
1.00 ´ 10-3
|
|
0
|
Change
(mol):
|
-1.00 ´ 10-3
|
-1.00 ´ 10-3
|
|
+1.00 ´ 10-3
|
Final
(mol):
|
2.00 ´ 10-3
|
0
|
|
1.00 ´ 10-3
|
NH3(aq) + HCl(aq) ® NH4Cl(aq)
At this stage, we have
a buffer system made up of NH3 and NH (from the
salt, NH4Cl). We use
the Henderson-Hasselbalch equation to calculate the pH.


pH = 9.55
|
c.
|
This part is solved similarly to part (b).
The
number of moles of HCl in 20.0 mL is:

The changes in number of moles are summarized.
|
NH3(aq) +
|
HCl(aq)
|
→
|
NH4Cl(aq)
|
Initial
(mol):
|
3.00 ´ 10-3
|
2.00 ´ 10-3
|
|
0
|
Change
(mol):
|
-2.00 ´ 10-3
|
-2.00 ´ 10-3
|
|
+2.00 ´ 10-3
|
Final
(mol):
|
1.00 ´ 10-3
|
0
|
|
2.00 ´ 10-3
|
At this stage, we have a
buffer system made up of NH3 and NH (from the
salt, NH4Cl). We use
the Henderson-Hasselbalch equation to calculate the pH.


pH = 8.95
|
d.
|
We have reached the equivalence
point of the titration. 3.00 ´
10-3
mole of NH3 reacts with
3.00 ´
10-3
mole HCl to produce 3.00 ´ 10-3
mole of NH4Cl. The only
major species present in solution at the equivalence point is the salt,
NH4Cl, which contains the conjugate acid, NH . Let's
calculate the molarity of NH . The
volume of the solution is: (10.0
mL + 30.0 mL = 40.0 mL = 0.0400 L).

We set up the hydrolysis of NH , which is a weak acid.
|

|

|

|

|

|
Initial (M):
|
0.0750
|
|
|
0
|
0
|
Change (M):
|
-x
|
|
|
+x
|
+x
|
Equilibrium (M):
|
0.0750 - x
|
|
|
x
|
x
|


x
= 6.5 ´
10-6
M =
[H3O+]
pH
= 5.19
|
e.
|
We have passed the equivalence
point of the titration. The excess
strong acid, HCl, will determine the pH at this point. The moles of HCl in 40.0 mL are:

The changes in number of moles are summarized.
|
NH3(aq) +
|
HCl(aq)
|
→
|
NH4Cl(aq)
|
Initial
(mol):
|
3.00 ´ 10-3
|
4.00 ´ 10-3
|
|
0
|
Change
(mol):
|
-3.00 ´ 10-3
|
-3.00 ´ 10-3
|
|
+3.00 ´ 10-3
|
Final
(mol):
|
0
|
1.00 ´ 10-3
|
|
3.00 ´ 10-3
|
Let's
calculate the molarity of the HCl in solution. The volume of the solution is now 50.0 mL = 0.0500
L.

HCl is a strong acid.
The pH is:
pH = -log(0.0200) =
1.70
|
|
|
17.37
|
a.
|
HCOOH is a weak acid and NaOH
is a strong base. Suitable
indicators are cresol red and
phenolphthalein.
|
b.
|
HCl is a strong acid and KOH is
a strong base. Most of the indicators in Table 17.3
are suitable for a strong acid-strong base titration. Exceptions are thymol blue and, to a
lesser extent, bromophenol blue and methyl orange.
|
c.
|
HNO3 is a strong
acid and CH3NH2 is a weak base. Suitable indicators are bromophenol blue, methyl orange,
methyl red, and chlorophenol blue.
|
|
|
17.38
|
CO2 in the air
dissolves in the solution:
CO2
+ H2O H2CO3
The carbonic acid
neutralizes the NaOH, lowering the pH.
When the pH is lowered, the phenolphthalein indicator returns to its
colorless form.
|
|
17.39
|
The weak acid equilibrium is
HIn(aq)
H+(aq)
+ In-(aq)
We can write a Ka
expression for this equilibrium.

Rearranging,

From the pH, we can calculate the H+
concentration.
[H+] =
10-pH =
10-4 =
1.0 ´
10-4
M

Since the
concentration of HIn is 100 times greater than the concentration of In-,
the color of the solution will be that of HIn, the nonionized formed. The color of the solution will be red.
|
|
17.40
|
When [HIn] » [In-]
the indicator color is a mixture of the colors of HIn and In-. In other words, the indicator color changes
at this point. When [HIn] »
[In-]
we can write:

[H+] = Ka =
2.0 ´
10-6
pH = 5.70
|
|
17.41
|
1.
|
Diagram (c)
|
2.
|
Diagram (b)
|
3.
|
Diagram (d)
|
4.
|
Diagram (a). The pH at the
equivalence point is below 7 (acidic) because the species in solution
at the equivalence point is the conjugate acid of a weak base.
|
|
|
17.42
|
1.
|
Diagram (b)
|
2.
|
Diagram (d)
|
3.
|
Diagram (a)
|
4.
|
Diagram (c). The pH at the
equivalence point is above 7 (basic) because the species in solution
at the equivalence point is the conjugate base of a weak acid.
|
|
|
17.49
|
a.
|
The solubility
equilibrium is given by the equation
AgI(s)
Ag+(aq) + I-(aq)
The expression for Ksp is given by
Ksp = [Ag+][I-]
The value of Ksp
can be found in Table 17.4 of the text.
If the equilibrium concentration of silver ion is the value given,
the concentration of iodide ion must be

|
b.
|
The value of Ksp for aluminum
hydroxide can be found in Table 17.4 of the text. The equilibrium expressions are:
Al(OH)3(s)
Al3+(aq) + 3OH-(aq)
Ksp =
[Al3+][OH-]3
Using the given value of the
hydroxide ion concentration, the equilibrium concentration of aluminum
ion is:

Think About
It:
|
What is
the pH of this solution? Will
the aluminum concentration change if the pH is altered?
|
|
|
|
17.50
|
In each part,
we can calculate the number of moles of compound dissolved in one liter of
solution
(the molar solubility). Then, from
the molar solubility, s, we can
determine Ksp.
a.
|

Consider the dissociation of SrF2 in
water. Let s be the molar solubility of SrF2.
|
SrF2(s)
|

|
Sr2+(aq)
|
+ 2F-(aq)
|
Initial (M):
|
|
|
0
|
0
|
Change (M):
|
-s
|
|
+s
|
+2s
|
Equilibrium (M):
|
|
|
s
|
2s
|
Ksp =
[Sr2+][F-]2 =
(s)(2s)2 =
4s3
The molar solubility (s) was calculated above. Substitute into the equilibrium
constant expression to solve for Ksp.
Ksp
= [Sr2+][F-]2 =
4s3 =
4(5.8 ´
10-4)3 = 7.8 ´
10-10
|
b.
|

(b) is solved in a
similar manner to (a)
The equilibrium equation
is:
|

|

|
|

|
Initial (M):
|
|
|
0
|
0
|
Change (M):
|
-s
|
|
+3s
|
+s
|
Equilibrium (M):
|
|
|
3s
|
s
|
Ksp =
[Ag+]3[PO43-] =
(3s)3(s)
= 27s4 = 27(1.6 ´
10-5)4 =
1.8 ´
10-18
|
|
|
17.51
|
For MnCO3
dissolving, we write
MnCO3(s)
Mn2+(aq) + CO (aq)
For every
mole of MnCO3 that dissolves, one mole of Mn2+ will
be produced and one mole of CO will be
produced. If the molar solubility of
MnCO3 is s mol/L, then
the concentrations of Mn2+ and CO are:
[Mn2+] =
[CO ] = s =
4.2 ´
10-6
M
Ksp =
[Mn2+][CO ] = s2 =
(4.2 ´
10-6)2 = 1.8 ´
10-11
|
|
17.52
|
First, we can
convert the solubility of MX in g/L to mol/L.

The equilibrium equation is:
|
MX(s)
|

|
Mn+(aq)
|
+ Xn-(aq)
|
Initial (M):
|
|
|
0
|
0
|
Change (M):
|
-s
|
|
+s
|
+s
|
Equilibrium (M):
|
|
|
s
|
s
|
Ksp =
[Mn+][Xn-] = s2 =
(1.34 ´
10-5)2 = 1.80 ´
10-10
|
|
17.53
|
The charges
of the M and X ions are +3 and -2, respectively (are other
values possible?). We first
calculate the number of moles of M2X3 that dissolve
in 1.0 L of water. We carry an
additional significant figure throughout this calculation to minimize
rounding errors.

The molar solubility, s,
of the compound is therefore 1.3 ´ 10-19
M. At equilibrium the concentration of M3+
must be 2s and that of X2-
must be 3s.
Ksp =
[M3+]2[X2-]3 =
[2s]2[3s]3 =
108s5
Since these are equilibrium concentrations, the value of Ksp can be found by
simple substitution
Ksp =
108s5 =
108(1.25 ´
10-19)5 = 3.3 ´
10-93
|
|
17.54
|
We can look up the Ksp
value of CaF2 in Table 17.4 of the text. Then, setting up the dissociation
equilibrium of CaF2 in water, we can solve for the molar
solubility, s.
Consider the dissociation of CaF2 in water.
|
CaF2(s)
|

|
Ca2+(aq)
|
+ 2F-(aq)
|
Initial (M):
|
|
|
0
|
0
|
Change (M):
|
-s
|
|
+s
|
+2s
|
Equilibrium (M):
|
|
|
s
|
2s
|
Recall, that the concentration of a pure solid does not
enter into an equilibrium constant expression. Therefore, the concentration of CaF2
is not important.
Substitute
the value of Ksp and
the concentrations of Ca2+ and F- in terms of s into the solubility product
expression to solve for s, the
molar solubility.
Ksp =
[Ca2+][F-]2
4.0
´
10-11 =
(s)(2s)2
4.0
´
10-11 =
4s3
s =
molar solubility = 2.2
´ 10-4 mol/L
The molar
solubility indicates that 2.2 ´ 10-4
mol of CaF2 will dissolve in 1 L of an aqueous solution.
|
|
17.55
|
Let s be the molar
solubility of Zn(OH)2.
The equilibrium concentrations of the ions are then
[Zn2+]
= s and [OH-]
= 2s
Ksp =
[Zn2+][OH-]2 =
(s)(2s)2 = 4s3 =
1.8 ´
10-14
s = 
[OH-]
= 2s = 3.30 ´
10-5
M and pOH = 4.48
pH
= 14.00 -
4.48 = 9.52
Think About It:
|
If the Ksp
of Zn(OH)2 were smaller by many more powers of ten, would 2s still be the hydroxide ion
concentration in the solution?
|
|
|
17.56
|
First we can
calculate the OH- concentration from the
pH.
pOH =
14.00 -
pH
pOH =
14.00 -
9.68 = 4.32
[OH-] =
10-pOH =
10-4.32 =
4.8 ´
10-5
M
The equilibrium equation is:
MOH(s)
M+(aq)
+ OH-(aq)
From the balanced equation we know that [M+]
= [OH-]
Ksp =
[M+][OH-] =
(4.8 ´
10-5)2 = 2.3 ´
10-9
|
|
17.57
|
According to the solubility rules, the only precipitate that
might form is BaCO3.
Ba2+(aq) + CO (aq) ® BaCO3(s)
The number of moles of Ba2+ present in the original
20.0 mL of Ba(NO3)2 solution is

The total volume after combining the two solutions is 70.0
mL. The concentration of Ba2+
in 70 mL is

The number of moles of CO present in
the original 50.0 mL Na2CO3 solution is

The concentration of CO in the 70.0
mL of combined solution is

Now we must compare Q
and Ksp. From Table
17.4 of the text, the Ksp
for BaCO3 is 8.1 ´ 10-9. As for Q,
Q
= [Ba2+]0[CO ]0
= (2.9 ´
10-2)(7.1
´
10-2) =
2.1 ´
10-3
Since (2.1 ´ 10-3)
> (8.1 ´
10-9),
then Q > Ksp.
Therefore, yes, BaCO3
will precipitate.
|
|
17.58
|
The net ionic
equation is:
Sr2+(aq) + 2F-(aq)
→ SrF2(s)
Let’s find
the limiting reagent in the precipitation reaction.


From the stoichiometry of the
balanced equation, twice as many moles of F- are required to
react with Sr2+. This
would require 0.0076 mol of F-, but we only
have 0.0045 mol. Thus, F-
is the limiting reagent.
Let’s assume
that the above reaction goes to completion.
Then, we will consider the equilibrium that is established when SrF2
partially dissociates into ions.
|
Sr2+(aq)
|
+ 2 F-(aq)
|
→
|
SrF2(s)
|
Initial
(mol):
|
0.0038
|
0.0045
|
|
0
|
Change
(mol):
|
-0.00225
|
-0.0045
|
|
+0.00225
|
Final
(mol):
|
0.00155
|
0
|
|
0.00225
|
Now, let’s
establish the equilibrium reaction.
The total volume of the solution is 100 mL = 0.100 L. Divide the above moles by 0.100 L to
convert to molar concentration.
|
SrF2(s)
|

|
Sr2+(aq)
|
+ 2F-(aq)
|
Initial (M):
|
0.0225
|
|
0.0155
|
0
|
Change (M):
|
-s
|
|
+s
|
+2s
|
Equilibrium (M):
|
0.0225
-
s
|
|
0.0155
+ s
|
2s
|
Write the
solubility product expression, then solve for s.
Ksp =
[Sr2+][F-]2
2.0
´
10-10 =
(0.0155 + s)(2s)2 » (0.0155)(2s)2
s
= 5.7 ´ 10-5
M
[F-] =
2s = 1.1 ´
10-4 M
[Sr2+] =
0.0155 + s = 0.016 M
Both sodium
ions and nitrate ions are spectator ions and therefore do not enter into
the precipitation reaction.


|
|
17.59
|
Strategy:
|
The addition of soluble salts to the solution changes
the concentrations of M2+ and A–. The salt which causes the largest
change in concentration (and therefore a change in the reaction quotient,
Q) will cause the precipitation
of the greatest quantity of MA2.
|
Setup:
|
The balanced equation is
MA2 M2+
+ 2A–
Therefore,
Ksp = [M2+][A–]2
|
Solution:
|
Since the anion, A–, in the Ksp equation is
squared, increasing its value will have the greater impact on the Q value than an equivalent change
in the cation, M2+.
AlA3 contains three anions, whereas NaA and
BaA2, contain only one or two.
M3(PO4)2 contains three cations,
while M(NO3)2 contains only one cation. Since the [A–] is squared in
the Ksp equation, changing its concentration will have a
greater impact on the precipitation than the three-fold change in [M2+]. Therefore, AlA3 will cause the greatest quantity of MA2 to precipitate.
|
|
|
17.60
|
Strategy:
|
If the reaction quotient, Q, is less than or equal to Ksp, then no precipitate will form.
|
Setup:
|
The equation for the dissociation is:
M2B
2M+
+ B–
Therefore,
Ksp = [M+]2[B–]
The saturated solution is at equilibrium. There are six cations and three
anions. This gives:
Ksp = [M+]2[B–]
= (6)2(3) = 108
Since the volumes are equal and additive when
combined, the volume doubles and the concentration of each ion is reduced
by one-half.
|
Solution:
|
a.
|
Diagram 2 has eight cations and diagram (a) has
fourteen anions. When added to
an equal volume of solution, the concentration is reduced by one-half,
yielding concentrations of four and seven, respectively.
This gives:
Q = [M+]2[B–]
= (4)2(7) = 112
Since 112 is greater than 108, a precipitate will
form.
|
b.
|
Diagram (b) has twelve anions. The number of cations was determined
in part (a). This gives:
Q = [M+]2[B–]
= (4)2(6) = 96
Since 96 is less than 108, no precipitate will form.
|
c.
|
Diagram (c) has four anions. This gives:
Q = [M+]2[B–]
= (4)2(2) = 32
Since 32 is less than 108, no precipitate will form.
|
d.
|
Diagram (d) has eight anions. This gives:
Q = [M+]2[B–]
= (4)2(4) = 64
Since 64 is less than 108, no precipitate will form.
|
e.
|
Diagram (e) has ten anions. This gives:
Q = [M+]2[B–]
= (4)2(5) = 80
Since 80 is less than 108, no precipitate will form.
|
Solutions (b),
(c), (d) and (e) can be added to the solution of MNO3
without causing a precipitate to form.
|
|
|
17.64
|
First let s be the molar solubility of CaCO3
in this solution.
|

|

|
|

|
Initial (M):
|
|
|
0.050
|
0
|
Change (M):
|
-s
|
|
+s
|
+s
|
Equilibrium (M):
|
|
|
(0.050
+ s)
|
s
|
CaCO3(s)
⇄ Ca2+(aq) + CO (aq)
CaCO3(s)
⇄ Ca2+(aq) + CO (aq)
Ksp =
[Ca2+][CO ] = (0.050 + s)s =
8.7 ´
10-9
We can assume 0.050 + s »
0.050, then

The mass of CaCO3 can then be found.

|
|
17.65
|
Strategy:
|
In parts (b) and (c), this is a common-ion
problem. In part (b), the common
ion is Br-,
which is supplied by both PbBr2 and KBr. Remember that the presence of a common
ion will affect only the solubility of PbBr2, but not the Ksp value because it is
an equilibrium constant. In part
(c), the common ion is Pb2+, which is supplied by both PbBr2
and Pb(NO3)2.
|
Solution:
|
a.
|
Set up a table to find
the equilibrium concentrations in pure water.
|
PbBr2(s)
|

|
Pb2+(aq)
|
+ 2Br-(aq)
|
Initial (M):
|
|
|
0
|
0
|
Change (M):
|
-s
|
|
+s
|
+2s
|
Equilibrium (M):
|
|
|
s
|
2s
|
Ksp =
[Pb2+][Br-]2
8.9 ´ 10-6 =
(s)(2s)2
s = molar solubility =
0.013 M or 1.3 ´ 10-2
M
|
b.
|
Set up a table to find the equilibrium concentrations
in 0.20 M KBr. KBr is a soluble salt that ionizes
completely giving an initial concentration of Br- = 0.20 M.
|
PbBr2(s)
|

|
Pb2+(aq)
|
+ 2Br-(aq)
|
Initial (M):
|
|
|
0
|
0.20
|
Change (M):
|
-s
|
|
+s
|
+2s
|
Equilibrium (M):
|
|
|
s
|
0.20 + 2s
|
Ksp =
[Pb2+][Br-]2
8.9 ´ 10-6 =
(s)(0.20 + 2s)2
8.9 ´ 10-6 » (s)(0.20)2
s = molar solubility =
2.2 ´
10-4 M
Thus, the molar solubility
of PbBr2 is reduced from 0.013 M to 2.2 ´ 10-4 M as a result of the common ion
(Br-) effect.
|
c.
|
Set up a table to find the equilibrium
concentrations in 0.20 M
Pb(NO3)2.
Pb(NO3)2 is a soluble salt that
dissociates completely giving an initial concentration of [Pb2+]
= 0.20 M.
|
PbBr2(s)
|

|
Pb2+(aq)
|
+ 2Br-(aq)
|
Initial (M):
|
0.20
|
|
0
|
|
Change (M):
|
-s
|
|
+s
|
+2s
|
Equilibrium (M):
|
|
|
0.20 + s
|
2s
|
Ksp =
[Pb2+][Br-]2
8.9 ´ 10-6 =
(0.20 + s)(2s)2
8.9 ´ 10-6 » (0.20)(2s)2
s = molar solubility =
3.3 ´
10-3 M
Thus,
the molar solubility of PbBr2 is reduced from 0.013 M to 3.3 ´
10-3 M as a result of the common ion
(Pb2+) effect.
|
|
Think About It:
|
You should also be able to predict the decrease in
solubility due to a common-ion using
Le Châtelier's principle. Adding
Br-
or Pb2+ ions shifts the system to the left, thus decreasing
the solubility of PbBr2.
|
|
|
17.66
|
We first calculate the concentration of chloride ion in the
solution.

|
AgCl(s)
|

|
Ag+(aq)
|
+ Cl-(aq)
|
Initial (M):
|
|
|
0.000
|
0.180
|
Change (M):
|
-s
|
|
+s
|
+s
|
Equilibrium (M):
|
|
|
s
|
(0.180
+ s)
|
If we assume that (0.180 + s) »
0.180, then
Ksp =
[Ag+][Cl-] =
1.6 ´
10-10

The molar solubility of AgCl is 8.9 ´
10-10 M.
|
|
17.67
|
a.
|
The equilibrium equation is:
|

|

|
|

|
Initial (M):
|
|
|
0
|
0
|
Change (M):
|
-s
|
|
+s
|
+s
|
Equilibrium (M):
|
|
|
s
|
s
|
Ksp =
[Ba2+][SO ]
1.1
´
10-10 =
s2
s =
1.0 ´
10-5
M
The molar
solubility of BaSO4 in pure water is 1.0 ´
10-5
mol/L.
|
b.
|
The initial concentration of SO is 1.0 M.
|

|

|
|

|
Initial (M):
|
|
|
0
|
1.0
|
Change (M):
|
-s
|
|
+s
|
+s
|
Equilibrium (M):
|
|
|
s
|
1.0 + s
|
Ksp =
[Ba2+][SO ]
1.1
´
10-10 =
(s)(1.0 + s)
» (s)(1.0)
s =
1.1 ´
10-10
M
Due to the
common ion effect, the molar solubility of BaSO4
decreases to 1.1 ´ 10-10
mol/L in
1.0 M SO (aq)
compared to 1.0 ´ 10-5
mol/L in pure water.
|
|
|
17.68
|
When the
anion of a salt is a base, the salt will be more soluble in acidic solution
because the hydrogen ion decreases the concentration of the anion (Le
Chatelier's principle):
B-(aq) + H+(aq)
HB(aq)
a.
|
BaSO4
will be slightly more soluble because SO is a base
(although a weak one).
|
b.
|
The solubility of PbCl2
in acid is unchanged over the solubility in pure water because HCl is a
strong acid, and therefore Cl-
is a negligibly weak base.
|
c.
|
Fe(OH)3 will be more soluble in acid because OH-
is a base.
|
d.
|
CaCO3 will be more soluble in acidic solution
because the CO ions
react with H+ ions to form CO2 and H2O. The CO2 escapes from the
solution, shifting the equilibrium.
|
|
|
17.69
|
a.
|
I-
is the conjugate base of the strong acid HI
|
b.
|
SO (aq) is a weak base
|
c.
|
OH-(aq) is a strong base
|
d.
|
C2O (aq)
is a weak base
|
e.
|
PO (aq)
is a weak base
|
The solubilities of the above (b – e) will increase in acidic
solution. Only (a), which contains
an extremely weak base (I- is the conjugate base
of the strong acid HI) is unaffected by the acid solution.
|
|
17.70
|
In water:
Mg(OH)2 Mg2+
+ 2OH-
s 2s
Ksp =
4s3 =
1.2 ´
10-11
s = 1.4 ´
10-4 M
In a buffer at pH = 9.0
[H+] =
1.0 ´
10-9
[OH-] =
1.0 ´
10-5
1.2
´
10-11 =
(s)(1.0 ´
10-5)2
s = 0.12 M
|
|
17.71
|
From Table 17.4, the value of Ksp for iron(II) hydroxide is 1.6 ´
10-14.
a.
|
At pH = 8.00, pOH = 14.00 -
8.00 = 6.00, and [OH-] = 1.0 ´
10-6
M

The molar solubility of iron(II) hydroxide
at pH = 8.00 is 0.016 M or 1.6 ´
10-2
M
|
b.
|
At pH = 10.00, pOH = 14.00 -
10.00 = 4.00, and [OH-] = 1.0 ´
10-4
M

The molar solubility of iron(II) hydroxide
at pH = 10.00 is 1.6 ´
10-6
M.
|
|
|
17.72
|
The solubility product expression
for magnesium hydroxide is
Ksp =
[Mg2+][OH-]2 =
1.2 ´
10-11
We find the hydroxide ion
concentration when [Mg2+] is 1.0 ´ 10-10
M

Therefore the concentration
of OH-
must be slightly greater than 0.35 M.
|
|
17.73
|
We first determine the effect of the added ammonia. Let's calculate the concentration of NH3. This is a dilution problem.
MiVi
= MfVf
(0.60
M)(2.00 mL) = Mf(1002 mL)
Mf =
0.0012 M NH3
Ammonia is a weak base (Kb =
1.8 ´
10-5).
|

|

|

|

|

|
Initial (M):
|
0.0012
|
|
|
0
|
0
|
Change (M):
|
-x
|
|
|
+x
|
+x
|
Equilibrium (M):
|
0.0012
-
x
|
|
|
x
|
x
|


Solving the resulting quadratic equation gives x = 0.00014, or [OH-] = 0.00014 M
This is a solution of iron(II) sulfate, which contains Fe2+
ions. These Fe2+ ions
could combine with OH- to precipitate Fe(OH)2. Therefore, we must use Ksp for iron(II)
hydroxide. We compute the value of Qc for this solution.
Fe(OH)2(s)
Fe2+(aq) + 2OH-(aq)
Q =
[Fe2+]0[OH-] = (1.0 ´ 10-3)(0.00014)2 =
2.0 ´
10-11
Note that when adding 2.00 mL of NH3 to 1.0 L of FeSO4,
the concentration of FeSO4 will decrease slightly. However, rounding off to 2 significant
figures, the concentration of 1.0 × 10-3 M does not change. Q
is larger than Ksp
[Fe(OH)2] = 1.6 ´ 10-14. The concentrations of the ions in
solution are greater than the equilibrium concentrations; the solution is
saturated. The system will shift
left to reestablish equilibrium; therefore, a precipitate of Fe(OH)2
will form.
|
|
17.74
|
First find the molarity of the copper(II) ion


Because of the magnitude of
the Kf for formation
of Cu(NH3) , the position of equilibrium will be far to the right. We assume essentially all the copper ion
is complexed with NH3.
The NH3 consumed is 4 ´ 0.0174 M = 0.0696 M. The uncombined NH3
remaining is (0.30 - 0.0696) M, or 0.23 M. The equilibrium concentrations of Cu(NH3) and NH3
are therefore 0.0174 M and 0.23 M,
respectively. We find [Cu2+]
from the formation constant expression.

[Cu2+] = 1.2 ´
10-13 M
|
|
17.75
|
Strategy:
|
The addition of Cd(NO3)2 to the
NaCN solution results in complex ion formation. In solution, Cd2+ ions will
complex with CN- ions. The concentration of Cd2+
will be determined by the following equilibrium
Cd2+(aq) + 4CN-(aq) Cd(CN)
From Table 17.5 of the text, we see that the formation
constant (Kf) for
this reaction is very large (Kf =
7.1 ´
1016). Because Kf is so large, the
reaction lies mostly to the right.
At equilibrium, the concentration of Cd2+ will be very
small. As a good approximation, we
can assume that essentially all the dissolved Cd2+ ions end up
as Cd(CN) ions.
What is the initial concentration of Cd2+ ions? A very small amount of Cd2+
will be present at equilibrium.
Set up the Kf
expression for the above equilibrium to solve for [Cd2+].
|
Solution:
|
Calculate the initial concentration of Cd2+
ions.

If we assume that the above equilibrium goes to
completion, we can write
|

|

|

|

|
Initial (M):
|
4.2 ´ 10-3
|
0.50
|
|
0
|
Change (M):
|
-4.2 ´ 10-3
|
-4(4.2 ´ 10-3)
|
|
+4.2 ´ 10-3
|
Equilibrium (M):
|
0
|
0.48
|
|
4.2 ´ 10-3
|
To find the concentration of free Cd2+ at
equilibrium, use the formation constant expression.

Rearranging,

Substitute the equilibrium concentrations calculated
above into the formation constant expression to calculate the equilibrium
concentration of Cd2+.

[Cd(CN) ]
= (4.2 ´
10-3
M) -
(1.1 ´
10-18) =
4.2 ´
10-3
M
[CN-] =
0.48 M + 4(1.1 ´
10-18
M) =
0.48 M
|
Think About It:
|
Substitute the equilibrium
concentrations calculated into the formation constant expression to
calculate Kf. Also, the small value of [Cd2+]
at equilibrium, compared to its initial concentration of 4.2 ´
10-3
M, certainly justifies our
approximation that almost all the Cd2+ ions react.
|
|
|
17.76
|
The reaction
Al(OH)3(s) + OH-(aq)
Al(OH) (aq)
is the sum of the two known
reactions
Al(OH)3(s)
Al3+(aq) + 3OH-(aq) Ksp =
1.8 ´
10-33
Al3+(aq) + 4OH-(aq)
Al(OH) (aq) Kf =
2.0 ´
1033
The equilibrium constant is

When pH = 14.00, [OH-]
= 1.0 M, therefore
[Al(OH) ] = K[OH-] =
(3.6)(1 M) =
3.6 M
This
represents the maximum possible concentration of the complex ion at pH
14.00. Since this is much larger
than the initial 0.010 M, the complex ion will be the predominant
species.
|
|
17.77
|
Silver
iodide is only slightly soluble. It
dissociates to form a small amount of Ag+ and I-
ions. The Ag+ ions then
complex with NH3 in solution to form the complex ion Ag(NH3) . The
balanced equations are:
AgI(s)
Ag+(aq) + I-(aq) Ksp = [Ag+][I-] =
8.3 ´
10-17
Overall: AgI(s)
+ 2NH3(aq) Ag(NH3) (aq) + I-(aq) K
= Ksp ´ Kf = 1.2 ´ 10-9
If s is the molar solubility of AgI then,
|

|

|

|

|

|
Initial (M):
|
|
1.0
|
|
0.0
|
0.0
|
Change (M):
|
-s
|
-2s
|
|
+s
|
+s
|
Equilibrium (M):
|
|
(1.0
-
2s)
|
|
s
|
s
|
Because Kf is large, we can assume all of the silver ions
exist as Ag(NH3) . Thus,
[Ag(NH3) ] = [I-] = s
We can write the equilibrium
constant expression for the above reaction, then solve for s.

s
= 3.5 ´ 10-5 M
At equilibrium, 3.5 ´
10-5
moles of AgI dissolves in 1 L of 1.0 M
NH3 solution.
|
|
17.78
|
Strategy:
|
The formula of hydroxyapatite is given in Section
17.4: Ca5(PO4)3OH. Write the dissociation equation for Ca5(PO4)3OH
and solve for molar solubility using the equilibrium expression.
|
Setup:
|
The equation for the dissociation of Ca5(PO4)3OH
is
Ca5(PO4)3OH(s)
5Ca2+(aq) + 3PO (aq) +
OH–(aq)
and the equilibrium expressions is Ksp = [Ca2+]5[PO ]3[OH–].
|
Solution:
|
|

|

|

|

|

|
Initial (M):
|
|
|
0
|
0
|
0
|
Change (M):
|
|
|
+5s
|
+3s
|
+s
|
Equilibrium (M):
|
|
|
5s
|
3s
|
s
|
Therefore,
2 × 10–59
= (5s)5(3s)3(s)
2 × 10–59 = (3125s5)(27s3)(s)
2.4 × 10–64 = s9
s = 9 × 10–8 M
|
|
|
17.79
|
Strategy:
|
Substitute the molar solubility into the equilibrium
expression to determine Ksp.
|
Setup:
|
s = 7 × 10–8 M
The equation and the equilibrium expression for the
dissociation of Ca5(PO4)3F are:
Ca5(PO4)3F(s)
5Ca2+(aq) + 3PO (aq) +
F–(aq)
and
Ksp = [Ca2+]5[PO ]3[F–]
|
Solution:
|
Substituting the molar solubility into the equilibrium
expression gives
Ksp = [Ca2+]5[PO ]3[F–] = [5(7 × 10–8 )]5[(3(7 × 10–8 )]3[7 × 10–8]
Ksp = (5.3 × 10–33)(9.3
× 10–21)(7 × 10–8)
Ksp = 3 × 10–60
|
|
|
17.82
|
a.
|
The solubility
product expressions for both substances have exactly the same
mathematical form and are therefore directly comparable. The substance having the smaller Ksp (AgI) will precipitate first.
(Why?)
|
| | |