Sie sind auf Seite 1von 39

PHR 123: Physical Pharmacy-I, chapter-2

IONIC EQUILIBRIA, ACIDS & BASES, BUFFERS &


BUFFERING AGENTS

Ionic Equilibria
OSTWALD’S DILUTION LAW
According to the Arrhenius Theory of dissociation, an
electrolyte dissociates into ions in water solutions.
These ions are in a state of equilibrium with the
undissociated molecules. This equilibrium is called the
Ionic equilibrium. Ostwald noted that the Law of
Mass Action can be applied to the ionic equilibrium
as in the case of chemical equilibria.
Let us consider a binary electrolyte AB which
dissociates in solution to form the ions A+ and B–.

Let C moles per litre be the concentration of the


electrolyte and α (alpha) its degree of dissociation. The
concentration terms at equilibrium may be written as :

Applying the Law of Mass Action:


At equilibrium :

The equilibrium constant Kc is called the Dissociation


constant or Ionization constant. It has a constant
value at a constant temperature.
If one mole of an electrolyte be dissolved in V litre of
the solution, then

V is known as the Dilution or the solution. Thus the


expression (1) becomes

This expression which correlates the variation of the


degree of dissociation of an electrolyte with dilution,
is known as Ostwald’s Dilution Law.
Limitation of Ostwald’s Law
Ostwald’s Dilution law holds good only for weak
electrolytes and fails completely when applied to
strong electrolytes. For strong electrolytes, which are
highly ionised in solution, the value of the dissociation
constant K, far from remaining constant, rapidly falls
with dilution.
Factors that explain the failure of Ostwald’s law in
case of strong electrolytes
(1) The law is based on Arrhenius theory which
assumes that only a fraction of the electrolyte is
dissociated at ordinary dilutions and complete
dissociation occurs only at infinite dilution. However,
this is true for weak electrolytes. Strong electrolytes are
almost completely ionised at all dilutions and V/
does not give the accurate value of .
(2) The Ostwald’s law is derived on the assumption that
the Law of Mass Action holds for the ionic equilibria as
well. But when the concentration of ions is very high,
the presence of charges affects the equilibrium. Thus
the Law of Mass Action in its simple form cannot be
applied.
(3) The ions obtained by dissociation may get hydrated
and may affect the concentration terms. Better results
are obtained by using activities instead of
concentrations.

THEORY OF STRONG ELECTROLYTES


A number of theories have been put forward by
different workers in order to explain the high
conductance of strong electrolytes. Southerland (1906)
held the view that ions in solution were surrounded by a
large number of ions of opposite charge. Due to the
weakening of interionic forces, the ionic velocities were
accelerated. This resulted in the increase of conductance
of the electrolyte solution. However, not much notice
was taken of Southerland’s view.

Ghosh’s Formula
In 1918 J.C. Ghosh revived the above theory. He
assumed that though the electrolyte is completely
ionised, all the ions are not free to move owing to the
influence of electric charges and it is only the mobile
ions which contribute to the conductance of the
solution. The value α represents the “active” proportion
of the electrolyte and can be determined by purely
electrical data, the Law of Mass Action playing no part
whatsoever. His formula

was applicable to univalent strong electrolytes.


Debye-Huckel Theory
In 1923 Debye and Huckel and in 1926 Onsagar put
forward the modern theory of strong electrolytes in
which account is taken of the electrostatic forces
between the ions. Without going into its mathematical
details, a brief outline of the main ideas of the theory is
given below :

(1) The strong electrolyte is completely ionised at all


dilutions. The present position as it has emerged from
the study of Raman spectra, X-ray analysis of crystals,
Distribution coefficients and vapour pressures is that
there is a very small amount of unionised substance also
present and therefore instead of saying ‘completely
ionised’ we should say ‘almost completely ionised’.

(2) Since oppositely charged ions attract each other,


it suggests that anions and cations are not
uniformly distributed in the solution of an
electrolyte but that the cations tend to be found in
the vicinity of anions and vice-versa (Fig. 2.1).
Though the solution is on the whole neutral, there is in
the vicinity of any given ion a predominance of ions of
opposite charge which we call as counter ions.
The ions are all the time on the move in all directions
but on the average, more counter ions than like ions
pass by any given ion. This spherical haze of opposite
charge is called ionic atmosphere.
Fig. 2.1: The ions of a particular charge are surrounded by more ions of
the opposite charge and solvent molecules. Because of the large number of
ions in concentrated solution, the ion activity is reduced due to hindered
movement of the ions.

(3) Decrease in equivalent conductance with increase


in concentration is due to fall in mobilities
of the ions due to greater inter-ionic effect and vice-
versa.
(4) The ratio V/ does not correctly give the degree of
dissociation α for strong electrolytes but only the
conductance or conductance coefficient fc.
(5) In spite of almost complete ionisation, V is much
less than.
The observed deviations are due to the following
reasons:
(1) Asymmetry or Relaxation Effect. Imagine a
central negative ion. This is surrounded by a number of
positively charged ions which form its ‘ionic
atmosphere.’ This atmosphere is symmetrically situated
in the absence of any electrical field and the force of
attraction exerted by the atmosphere on the central ion
is uniform in all directions. When an electric field is
applied, the negative ion moves towards the anode and
the positive ionic atmosphere towards the cathode. This
leaves a large number of positive ions behind it than
there are in front of the negative ion with the result that
the symmetry of the atmosphere about the central
negative ion is destroyed (Fig. 2.2) and it becomes
distorted.

Whereas initially the force of attraction exerted by the


atmosphere on the central ion was uniform it becomes
greater now behind the ion than in front. As a result of
this the negative ion experiences a force which tends to
drag it backwards and this slows down its movement in
the forward direction. This behavior is known as
Asymmetry Effect. It may be said that the negative ion
which leaves the ionic atmosphere of positive ions
behind to die away would build a new ionic atmosphere
and the asymmetry would be corrected. If this process
of building up and dying away were instantaneous,
there would be no cause for asymmetry and the
atmosphere would always be symmetrically placed
about the ion. But, as it is, the formation of the new
ionic atmosphere does not take place at the same rate at
which the old one decays and the latter lags behind or
takes more time, known as ‘relaxation time’. During
this interval, there is a preponderance of positive ions to
the left of the central negative ion which is under move
and these tend to drag it back. For this reason
asymmetry effect is also known as Relaxation Effect.
Moreover, the central ion moves into a place where
there is an excess of ions of its own sign and this has
also a retarding effect.

Fig. 2.2: (a ) Symmetrical ionic atmosphere at rest; (b ) Asymmetrical


ionic atmosphere under the influence of applied field.

(2) Electrophoretic Effect. Another factor which acts


as a drag and tends to retard the motion of an ion in
solution is the tendency of the applied field to move the
ionic atmosphere (to which solvent molecules are also
attached) in a direction opposite to that in which the
central ion associated with solvent molecules is moving.
Thus the central negative ion moving towards the anode
has to make its way through the ionic atmosphere with
its associated solvent molecules which is moving in the
opposite direction i.e., towards the cathode. This causes
a retarding influence on the movement of the ion the
effect of which is equal to the increase in the viscous
resistance of the solvent. By analogy to the resistance
acting on the movement of a colloidal particle under an
electrical field, this effect is called Electrophoretic
Effect.

Both the above causes reduce the velocity of the ion and
operate in solutions of strong electrolytes with the result
that a value of equivalent conductance (V) lower than
the value at infinite dilution () is obtained. At infinite
dilution since the electrical effects are practically
absent, the two values tend to approach each other.
Debye-Huckel-Onsagar Conductance Equation takes
these causes into account and for a univalent electrolyte
supposed to be completely dissociated is written in the
form

where A and B are constants and c is the concentration


in gm-equivalents per litre.

DEGREE OF DISSOCIATION
When a certain amount of electrolyte (A+ B–) is
dissolved in water, a small fraction of it dissociates
to form ions (A+ and B–). When the equilibrium has
been reached between the undissociated and the
free ions, we have

The fraction of the amount of the electrolyte in


solution present as free ions is called the Degree
of dissociation.
If the degree of dissociation is represented by x, we can
write

The value of x can be calculated by applying the Law of


Mass Action to the ionic equilibrium
stated above :

If the value of the equilibrium constant, K, is given, the


value of x can be calculated.
THE COMMON–ION EFFECT
When a soluble salt (say A+C–) is added to a solution of
another salt (A+B–) containing a common ion (A+), the
dissociation of AB is suppressed.

By the addition of the salt (AC), the concentration of A +


increases. Therefore, according to Le Chatelier’s
principle, the equilibrium will shift to the left, thereby
decreasing the concentration of A+ ions. Or that, the
degree of dissociation of AB will be reduced.

The reduction of the degree of dissociation of a salt


by the addition of a common-ion is called the
Common-ion effect.
The common-ion effect can be applied to the ionic
equilibrium of a weak acid as HF.

NaF is added to the equilibrium mixture. The


concentration of F– (common ion) is increased. Thus the
equilibrium shifts to the left. In other words, the degree
of dissociation of HF decreases. It was found by
experiment that the degree of dissociation of HF in 1M
solution is 2.7, while the value reduces to 7.2 × 10 – 4
after the addition of 1M NaF.

FACTORS WHICH INFLUENCE THE DEGREE


OF DISSOCIATION
The degree of dissociation of an electrolyte in solution
depends upon the following factors:
(1) Nature of Solute
The nature of solute is the chief factor which
determines its degree of dissociation in solution. Strong
acids and strong bases, and the salts obtained by their
interaction are almost completely dissociated in
solution. On the other hand, weak acids and weak bases
and their salts are feebly dissociated.
(2) Nature of the solvent
The nature of the solvent affects dissociation to a
marked degree. It weakens the electrostatic forces of
attraction between the two ions and separates them.
This effect of the solvent is measured by its ‘dielectric
constant’. The dielectric constant of a solvent may be
defined as its capacity to weaken the force of attraction
between the electrical charges immersed in that solvent.
The dielectric constant of any solvent is evaluated
considering that of vaccum as unity. It is 4.1 in case of
ether, 25 in case of ethyl alcohol and 80 in case of
water. The higher the value of the dielectric
constant the greater is the dissociation of the electrolyte
dissolved in it because the electrostatic
forces vary inversely as the dielectric constant of the
medium. Water, which has a high value of dielectric
constant is, therefore, a strong dissociating solvent. The
electrostatic forces of attraction between the ions are
considerably weakened when electrolytes are dissolved
in it and as a result, the ions begin to move freely and
there is an increase in the conductance of the solution.
(3) Concentration
The extent of dissociation of an electrolyte is inversely
proportional to the concentration of its solution. The
less concentrated the solution, the greater will be the
dissociation of the electrolyte. This is obviously due to
the fact that in a dilute solution the ratio of solvent
molecules to the solute
molecules is large and the greater number of solvent
molecules will separate more molecules of the solute
into ions.

(4) Temperature
The dissociation of an electrolyte in solution also
depends on temperature. The higher the temperature
greater is the dissociation. At high temperature the
increased molecular velocities overcome the forces of
attraction between the ions and consequently the
dissociation is great.

Acids and Bases


There are three concepts of acids and bases.
(a) Arrhenius concept
(b) Bronsted-Lowry concept
(c) Lewis concept
ARRHENIUS CONCEPT
Savante Arrhenius (1884) proposed his concept of acids
and bases. According to this concept, an acid is a
compound that releases H+ ions in water; and a base
is a compound that releases OH– ions in water.
For example, HCl is an Arrhenius acid and NaOH is an
Arrhenius base.

Limitations of Arrhenius Concept


Arrhenius concept of acids and bases proved to be very
useful in the study of chemical reactions. However, it
has the following limitations :
(1) Free H+ and OH– ions do not exist in water. The
H+ and OH– ions produced by acids and bases
respectively do not exist in water in the free state. They
are associated with water molecules to form complex
ions through hydrogen bonding. Thus the H+ ion forms
a hydronium ion :

Similarly, OH– ion forms the complex H3O2- . Although


the hydrogen and hydroxyl ions are associated with
water molecule, for simplicity we shall generally write
them H+ and OH–.
(2) Limited to water only. Arrhenius defined acids and
bases as compounds producing H+ and OH– ions in
water only. But a truly general concept of acids and
bases should be appropriate to other solvents as well.
(3) Some bases do not contain OH–. Arrhenius base is
one that produces OH– ions in water. Yet there are
compounds like ammonia (NH3) and calcium oxide
(CaO) that are bases but contain no OH– ions in their
original formulation.
Arrhenius models of acids and bases, no doubt, proved
very helpful in interpreting their action.
However on account of its limitations the Arrhenius
concept needed to be modified.

BRONSTED–LOWRY CONCEPT
In 1923 J.N. Bronsted and J.M. Lowry independently
proposed a broader concept of acids and bases.
According to this theory, an acid is any molecule or
ion that can donate a proton (H +) a base is any
molecule or ion that can accept a proton
For brevity we can say that an acid is a proton donor
while a base is a proton acceptor.
An acid qualifying Bronsted-Lowry concept is termed a
Bronsted-Lowry acid or simply Bronsted
acid.
A base qualifying Bronsted-Lowry concept is termed a
Bronsted-Lowry base or simply Bronsted
base.
Examples of Bronsted acids and bases
(1) HCl gas and H2O. When dry HCl gas dissolves in
water, each HCl molecule donates a proton to a water
molecule to produce hydronium ion.

Thus HCl gas is a Bronsted acid and water that accepts


a proton is a Bronsted base.
Bronsted-Lowry concept is superior to Arrhenius
concept
(1) Much wider scope. Arrhenius concept of acids and
bases is restricted to the study of substances which can
release H+ or OH– ions in water. Bronsted-Lowry
concept embraces all molecules and ions that can
donate a proton (acids) and those which can accept a
proton (bases).
(2) Not limited to aqueous solutions. The Bronsted-
Lowry model is not limited to aqueous solutions as is
the case with Arrhenius model. It can be extended even
to the gas phase. For example, gaseous ammonia (a
Bronsted base) can react with hydrogen chloride gas (a
Bronsted acid) to give ammonium chloride.

Here a proton is donated by HCl to NH 3 as shown


above. Note that this is not considered as an acid-base
reaction according to Arrhenius concept.
(3) Release of OH– not necessary to qualify as a base.
Arrhenius base is a substance that releases OH – ions in
water. On the other hand, Bronsted base is a substance
that accepts a proton. Thus liquid ammonia (NH 3) does
not produce OH– ions in water but it is a recognised
base. But according to Bronsted-Lowry model, it
qualifies as a base since it can accept a proton to form
NH4+ (an acid).

Conjugate Acid-Base pairs


In an acid-base reaction the acid (HA) gives up its
proton (H+) and produces a new base (A–). The
new base that is related to the original acid is called a
conjugate (meaning related) base. Similarly the
original base (B–) after accepting a proton (H+) gives a
new acid (HB) which is called a conjugate acid.
A hypothetical reaction between the acid HA and the
base B– will illustrate the above definitions.

Fig. Illustration of conjugate pairs.


The acid (HA) and the conjugate base (A–) that are
related to each other by donating and accepting
a single proton, are said to constitute a conjugate
Acid-Base pair.
It may be noted that in any acid-base reaction, there are
two conjugate acid-base pairs, Thus, in
the above equation, the two conjugate pairs are : HA
and A–; and HB and B–.
Thus we can conclude that :
(a) a weak base has strong conjugate acid
(b) a weak acid has a strong conjugate base
LEWIS CONCEPT OF ACIDS AND BASES
In the early 1930s, G.N. Lewis proposed even a more
general model of acids and bases. According
to Lewis theory,
an acid is an electron-pair acceptor
a base is an electron-pair donor
Lewis pictured an acid and base as sharing the electron
pair provided by the base. This creates a
covalent bond (or coordinate bond) between the Lewis
acid and the Lewis base. The resulting
combination is called a Complex. If the Lewis acid be
denoted by A and the Lewis base by B, then the
fundamental equation of the Lewis theory can be
written as :

It may be noted that : (1) all cations or molecules short


of an electron-pair act as Lewis acids; and
(2) all anions or molecules having a lone electron-pair
act as Lewis bases.
Examples of Lewis reactions
(1) Between H+ and NH3. Proton (H+) is a Lewis acid
as it can accept an electron-pair. Ammonia
molecule (: NH3) has an electron-pair which it can
donate and is a Lewis base. Thus the Lewis reaction
between H+ and NH3 can be written as :

(2) Between H+ and OH–. A proton (H+) is an


electron-pair acceptor and, therefore, a Lewis acid.
The OH– is an electron-pair donor and hence a Lewis
base. Thus Lewis reaction between H+ and OH–
can be written as :

(3) Between BF3 and NH3. BF3 has six valence


electrons with B atom which can accept an
electron-pair and is a Lewis acid. The N atom of : NH3
has a lone electron-pair and is a Lewis base.
Lewis reaction between BF3 and NH3 may be written
as :
Superiority of Lewis model of acids and bases
The useful but limited model of Arrhenius was replaced
by a more general model of Bronsted and
Lowry. Even a more general model was proposed by
Lewis. However, the Bronsted-Lowry model is
now used in common practice.

The advantages of the Lewis acid-base model are:


(1) All the Bronsted-Lowry acid base reactions are
covered by the Lewis model. It is so because
the transfer or gain of a proton is accompanied by the
loss or donation of an electron-pair in both
types of reactions.
(2) Many reactions which do not involve transfer of a
proton e.g.,
BF3 + NH3 ⎯⎯→ BF3 — NH3
pH is one of the primary influences on the solubility of
most drugs that contain ionisable groups (Figure 2.2):
_ Acidic drugs, such as the non-steroidal anti-
inflammatory agents, are less soluble in acidic solutions
than in alkaline solutions because the predominant
undissociated species cannot interact with water
molecules to the same extent as the ionised form which
is readily hydrated. The equation relating the solubility,
S, of an acidic drug to the pH of the solution is:

where So is the solubility of the undissociated form of


the drug.
_ Basic drugs such as ranitidine are more soluble in
acidic solutions where the ionised form of the drug is
predominant. The equation relating the solubility, S, of
a basic drug to the pH of the solution is:

_ Amphoteric drugs such as the sulfonamides and


tetracyclines display both basic and acidic
characteristics. The zwitterion has the lowest solubility,
So, and the variation of solubility with pH is given by:
at pH values below the isoelectric point and

at pH values above the isoelectric point.

Figure 2.2 Solubility of acidic, basic and amphoteric drugs as a function of pH.

Ionisation of drugs in solution


Ionisation of weakly acidic drugs and their salts
_ If the weak acid is represented by HA, its ionisation in
water may be represented by the equilibrium:

_ The equilibrium constant, Ka, is referred to as the


ionization constant, dissociation constant or acidity
constant and is given by:
_ The negative logarithm of Ka is referred to as pKa, i.e.
pKa = –logKa.
_ When the pH of an aqueous solution of the weakly
acidic drug approaches to within 2 pH units of the pKa
there is a very pronounced change in the ionisation of
that drug (Figure 2.3):

Figure 2.3 Percentage ionisation of weakly acidic and weakly basic drugs as a
function of pH.

_ The percentage ionisation at a given pH can be


calculated from:

_ Weakly acidic drugs are virtually completely


unionised at pHs up to 2 units below their pKa and
virtually completely ionized at pHs greater than 2 units
above their pKa. They are exactly 50% ionised at their
pKa values.
_ Salts of weak acids are essentially completely ionised
in solution, for example when sodium salicylate (salt of
the weak acid, salicylic acid, and the strong base
NaOH) is dissolved in water, it ionises almost entirely
into the conjugate base of salicylic acid, HOC 6H5COO–,
and Na+ ions. The conjugate acids formed in this way
are subject to acid–base equilibria described by the
general equations above.

Ionisation of weakly basic drugs and their salts


_ If the weak acid is represented by B, its ionisation in
water may be represented by the equilibrium:

_ The equilibrium constant, Kb, is referred to as the


ionization constant, dissociation constant or basicity
constant and is given by:

_ The negative logarithm of Kb is referred to as pKb, i.e.


pKb = –logKb.
_ The percentage ionisation at a given pH can be
calculated from:
_ Weakly basic drugs are virtually completely ionised at
pHs up to 2 units below their pKa and virtually
completely unionized at pHs greater than 2 units above
their pKa. They are exactly 50% ionised at pHs equal to
their pKa values.
_ Salts of weak bases are essentially completely ionised
in solution; for example, ephedrine hydrochloride (salt
of the weak base, ephedrine, and the strong acid HCl)
exists in aqueous solution in the form of the conjugate
acid of the weak base, C6H5CH(OH)CH(CH3)N+H2CH3,
together with its Cl– counterions. The conjugate bases
formed in this way are subject to acid–base equilibria
described by the general equations above.

Ionisation of amphoteric drugs


_ These can function as either weak acids or weak bases
in aqueous solution depending on the pH and have pKa
values corresponding to the ionisation of each group.
_ If pKa of the acidic group, pKa acidic, is higher than
that of the basic group, pKa basic, they are referred to as
ordinary ampholytes and exist in solution as a cation, an
unionised form, and an anion depending on the pH of
the solution. For example, the ionisation of m-
aminophenol (pKaacidic = 9.8 and pKabasic = 4.4) changes
as the pH increases as follows:

_ If pKa acidic < pKa basic they are referred to as


zwitterionic ampholytes and exist in solution as a
cation, a zwitterion (having both positive and negative
charges), and an anion depending on the pH of the
solution. Examples of this type of compound include
the amino acids, peptides and proteins. Glycine (pKaacidic
= 2.3 and pKabasic = 9.6) ionises as follows:

_ The ionisation pattern of both types is more complex,


however, with drugs in which the difference in pKa of
the two groups is much smaller (< 2 pH units) because
of overlapping of the two equilibria.

Ionisation of polyprotic drugs


_ Several acids, for example citric, phosphoric and
tartaric acid, are capable of donating more than one
proton and these compounds are referred to as
polyprotic or polybasic acids.
Similarly, polyprotic bases are capable of accepting two
or more protons. Examples of polyprotic drugs include
the polybasic acids amoxicillin and fluorouracil, and the
polyacidic bases pilocarpine, doxorubicin and aciclovir.
_ Each stage of the dissociation of the drug may be
represented by an equilibrium expression and hence
each stage has a distinct pKa or pKb value. For example,
the ionisation of phosphoric acid occurs in three stages:

_ If the pKa values of each stage of dissociation are far


apart it is usually possible to assign them to the
ionisation of specific groups, but if they are within
about 2 pH units of each other this is not possible. For a
more complete picture of the dissociation it is necessary
to take into account all possible ways in which the
molecule may be ionised and all the possible species
present in solution. In this case the constants are called
microdissociation constants.

pH of drug solutions
The pH of a strong acid such as HCl is given by pH = –
log[H+]. This is because strong acids are completely
ionised in solution. However, as seen above, weak acids
and bases are only slightly ionised in solution and the
extent of their ionisation changes with pH and so
therefore does their pH. The pH at any particular
concentration, c, can be calculated from the pKa value:
_ Weakly acidic drugs: pH = ½ pKa – ½ log c.
_ Weakly basic drugs: pH = ½ pKw + ½ pKa + ½ log c.
_ Drug salts:
– Salts of a weak acid and a strong base:
pH = ½ pKw + ½ pKa + ½ log c
– Salts of a weak base and a strong acid:
pH = ½ pKa – ½ log c
– Salts of a weak acid and a weak base:
pH = ½ pKw + ½ pKa – ½ pKb
(note that there is no concentration term in this
equation, meaning that the pH does not vary with
concentration).

Buffers
_ Buffers are usually mixtures of a weak acid and its
salt (that is, a conjugate base), or a weak base and its
conjugate acid.
_ A mixture of a weak acid HA and its ionised salt (for
example, NaA) acts as a buffer because the A – ions
from the salt combine with the added H + ions, removing
them from solution as undissociated weak acid:

Added OH– ions are removed by combination with the


weak acid to form undissociated water molecules:

_ A mixture of a weak base and its salt acts as a buffer


because added H+ ions are removed by the base B to
form the salt and OH– ions are removed by the salt to
form undissociated water:

_ The concentration of buffer components required to


maintain a solution at the required pH may be
calculated from the Henderson–Hasselbalch equations:

_ The effectiveness of a buffer in minimising pH


change is expressed as the buffer capacity, β, calculated
from:

where c0 is the total initial buffer concentration.


A plot of β against pH (Figure 2.4) shows that:
– The buffer capacity is maximum when pH = pKa.
– Maximum buffer capacity, βmax, = 0.576co.
– If, instead of using a single weak monobasic acid,
which has a maximum buffer capacity at pH = pKa, you
use a suitable mixture of polybasic and monobasic
acids, it is possible to produce a buffer which is
effective over a wide pH range because each stage of
the ionisation of the polybasic acid has its own βmax
value. Such solutions are referred to as universal
buffers. A typical example is a mixture of citric acid
(pKa1 = 3.06, pKa2 = 4.78 and pKa3 = 5.40), Na2HPO4
(pKa of conjugate acid, H2PO4− = 7.2), diethylbarbituric
acid (pKa1 = 7.43) and boric acid (pKa1 = 9.24). This
buffer is effective over a pH range 2.4 to 12.

Figure 2.4 Buffer capacity of a weak acid/salt buffer as a function


of pH, showing maximum buffer capacity at pKa.

NUMERICAL PROBLEMS BASED ON pH

Das könnte Ihnen auch gefallen