Fundamentals
Fundamentals
pH
13
12 pK = 15
11 pK = 11
10
pK = 9
9
pK = 7
7
pK = 5
5
4
pK = 3
3
pK = 1
V [mL]
1 2 3 4 5 6 7 8 9 10
Fundamentals of Titration
Contents
Page
1 Introduction 5
4 Indication methods 27
4.1 Electrochemical indication 28
4.1.1 Galvanic cells 28
4.1.2 Reference electrodes 32
4.1.3 Metal electrodes 34
4.1.4 Glass electrodes 34
4.1.5 Ion selective electrodes 39
4.1.6 Measurement technique 41
4.2 Photometric indication 44
4.2.1 The METTLER phototrode 46
4.3 Special indication methods 47
4.3.1 Conductometric indication 47
Fundamentals of Titration 1
Fundamentals of Titration
Page
5 Types of titration 50
5.1 The direct titration 50
5.2 The back titration 51
5.3 The inverse titration 52
5.4 The substitution titration 53
5.5 The collective titration 54
5.6 The selective titration 55
5.7 The sequential titration 55
6 Titration curves 59
6.1 Measurement signal as a function of the titrant volume: E = f(V) 59
6.2 Measurement signal as a function of time: E = f(t) 63
6.3 Titrant volume as a function of time: V = f(t) 65
2 Fundamentals of Titration
Fundamentals of Titration
Page
Appendices 130
Appendix A: Tables of buffer values 131
Appendix B: Statistical tables 135
Appendix C: Tables of primary standards for the most important titrants 137
Index 141
Fundamentals of Titration 3
Fundamentals of Titration
4 Fundamentals of Titration
1 Introduction
1 Introduction
A balance, a burette and a suitable chemical reaction suffice to solve many quantitative
analytical problems. The analytical technique employed is called titration or titrimetric analysis
(titrimetry). The expression “volumetric analysis” is not recommended [1]. In a titration, part of
the sample containing the substance to be analyzed (the analyte) is dissolved in a suitable
solvent. A second chemical compound, the titrant, is added as a solution of known concentra-
tion in a controlled manner until the analyte has reacted quantitatively. From the consumption
and concentration of the titrant as well as the weight of sample used in the analysis, the content
of the analyte can be calculated.
From the above definition it follows that the following requirements must be fulfilled before a
titration can be performed:
– The basic chemical reaction – the titration reaction – must be rapid, straightforward and
quantitative.
– It must be possible to either prepare a titrant of exactly known content or determine the
reacting strength (titer) of the solution accurately.
– The course of the titration must be observable. The method used to follow the titration
progress is called indication.
– Determination of the equivalence point – the point at which the number of entities
(equivalents) of the titrant added is the same as the number of entities of sample analyte
present – must be unambiguous.
The titration reaction, the indication, the control and evaluation of the titration as well as the
assessment of the results (statistics) form the focal points of this supplement to the operating
instructions for the titrator DL70.
Fundamentals of Titration 5
2 Base units
The base units and calculation parameters of titrimetric analysis are associated with the base
quantity amount of substance and its base unit mole of the international system of units (SI)
[1]. The concepts and definitions regarding amount of substance and the quantities derived
from it are defined in [2].
Mole
The SI base unit for amount of substance is the mole (symbol of unit: mol). The mole is the
amount of substance of a system that contains just as many elementary entities as there are
atoms in 12 g of the carbon 12C isotope. One mole of a substance contains 6.022 • 1023
elementary entities. These can be atoms, molecules, ions, groups of atoms or electrons.
Molar mass M
The molar mass (symbol M) is a quantity related to the amount of substance. The molar mass
of a substance X is defined as its mass m divided by amount of substance n(X).
m
M(X) =
n(X)
6 Fundamentals of Titration
2 Base units
n(X)
c(X) =
V
Titer t
The titer (symbol t) of a titrimetric solution is the quotient of the actual concentration (ACTUAL
value) and the expected concentration (NOMINAL value).
c( X , ACTUAL )
t =
c( X , NOMINAL )
Fundamentals of Titration 7
2 Base units
The equivalent entity, abbreviated to equivalent, is the fraction 1/z* of such an entity. The
number of equivalents z* of each entity X is called the equivalent number.
a. HCl: z* = 1
H+ + Cl- + Na+ + OH- —> Na+ + Cl- + H2O
b. H2SO4: z* = 2
2H+ + SO42- + 2 Na+ + 2 OH- —> 2 Na+ + SO42- + 2 H2O
2. Redox equivalent: In a redox reaction, the reaction partners change their oxidation
number.
VII II II III
K + MnO4 + 5Fe + 8H —> K + Mn + 5Fe3+ + 4H2O
+ - 2+ + + 2+
VII II lV
+ - 2+ - +
2 K + 2 MnO4 + 3 Mn + 4 OH —>2 K + 5 MnO2 + 2 H2O
8 Fundamentals of Titration
2 Base units
– amount of substance
– molar mass
– concentration.
1
n( X ) = z* • n(X)
z*
1 M(X)
M( X) =
z* z*
Fundamentals of Titration 9
2 Base units
1
The following relation holds: c( X ) = z* • c(X)
z*
1 1•6
c( K 2 Cr 2 O 7 ) = = 0.408 mol/L
6 294.185 • 0.05
10 Fundamentals of Titration
2 Base units
Concentration of a titrant
The concentration of a titrant should be specified as equivalent concentration.
c( 1/z* X ) • M(X) • V
m =
z*
[1] Bureau International des Poids et Mesures, Le Système International d’Unités (SI), 5th French and English
Edition, BIPM, Sèvres 1985
[2] IUPAC Compendium of Analytical Nomenclature, Pergamon Press, 1978, page 175 ff. See also DIN 32625
Fundamentals of Titration 11
3 Titration reaction
The basis of each titrimetric method of analysis is the chemical reaction of the analyte with the
titrant. In order to understand the demands on this titration reaction, a brief introduction into
the fundamentals of the thermodynamics of chemical reactions is called for.
v1
aA + bB + cC <====> xX + yY + zZ
v2
x y z
X • Y • Z
= K
a b c
A • B • C
The concentrations of entities X, Y and Z are designated here by [X], [Y] and [Z] ([X] = c(X)).
This notation is usual in analytical chemistry.
Strictly speaking, the law of mass action can not be applied directly to the analytical
concentrations of the reaction partners. Real chemical systems are distinguished by the
mutual interaction of all molecules present. In solution, interactions occur between the
molecules of the dissolved substance and the solvent molecules. Here, it is only the quasi-free
or “effective” concentrations of the substances participating in reaction, the so-called activities,
that are decisive for the chemical reaction and the law of mass action. But for a formal
understanding and the calculations, only analytical concentrations are considered in the
present discussion.
12 Fundamentals of Titration
3 Titration reaction
The demand that the titration reaction proceed quantitatively and to completion is fulfilled when
the equilibrium constant K is so large that the equilibrium concentration of the analyte is
infinitely small in comparison with its concentration before titration.
The equilibrium constant K provides no direct information regarding the rate of the titration
reaction. Decisive for this is the rate of the forward reaction v1, the reaction of the analyte with
the titrant.
The following sections treat the thermodynamic constants that appear most frequently in
potentiometry, the solubility product of sparingly soluble salts, the ionic product of water and
the acidity constant of weak acids.
Many salts are only slightly soluble. If solutions of the corresponding ions are mixed,
precipitates are formed. The processes at the surface of a salt in contact with a saturated
solution lead to the establishment of a heterogenous equilibrium. Ions from the salt constantly
pass into solution, and ions from the solution are incorporated in the salt lattice.
AB(solid) <===> A+ + B-
+ -
A • B
= K
AB
As long as solid salt AB is present as precipitate, the concentration of AB remains constant and
is thus included in the equilibrium constant. This gives rise to the solubility product Ksp:
+ -
A B = K sp
A sparingly soluble salt is always precipitated when the solubility product of the participating
ions is exceeded. The lower the solubility product, the more insoluble the salt.
The solubility product of salts having the general formula AB2 has the following form:
2
2+ -
K sp = A B
Fundamentals of Titration 13
3 Titration reaction
AgCl 10-10
AgBr 5 10-13
•
AgI 10-16
PbSO4 10-8
When the conductivity of water is examined using very sensitive instruments, it is apparent that
even ultrapure, repeatedly distilled water has a very low conductivity. It is due to the following
reaction:
The forward reaction describes the proton transfer from one water molecule to another. This
equilibrium is present not only in pure water but also in all aqueous solutions. The correspond-
ing law of mass action is:
+ -
H 3O • OH
= K
2
H 2O
The concentrations of the H3O+ and OH--ions in solution can be changed drastically by addition
of an acid or base. But the concentration of the H2O molecules (55.5 mol/L) remains constant
in dilute solutions. The law of mass action can thus be simplified:
+ -
H 3O OH = Kw
The equilibrium constant Kw is known as the ionic product of water. It depends on the tempe-
rature and is 10-14 mol2/L2 at 23°C.
In dilute aqueous solutions the product of [H3O+] and [OH-] is thus constant. If one of the two
concentrations is known, the other can be calculated from a knowledge of Kw. In a neutral
solution the concentrations [H3O+] and [OH-] are equal:
+ - -7
H 3O = OH = K w = 10 mol/L
14 Fundamentals of Titration
3 Titration reaction
If, for example, the H3O+ concentration is increased to 10-2 mol/L by addition of acid, the OH-
concentration decreases to 10-12 mol/L. Specification of one of these concentrations allows
unequivocal identification of the nature of an aqueous solution. This led to the introduction of
the pH concept as
+
pH = –log H 3O
In acidic solutions([H3O+] > 10-7) the pH is less than 7, whereas in alkaline solutions it is greater
than 7. The pH of a neutral solution is 7.
The reaction of weak acids with water is described by the following equilibrium:
Acid HA reacts with the base H2O to form the conjugate base A- of HA and the conjugate acid
of H2O, namely H3O+.
+ -
H 3O • A
= K
HA • H 2O
In dilute solutions ([H2O] = constant) the following formula applies:
+ -
H 3O • A
= Ka
HA
The equilibrium constant Ka is known as the acidity constant of acid HA and characterizes
the strength of an acid. Strong acids have a large acidity constant, weak acids a very low one.
The negative logarithm of Ka is frequently employed in calculations:
pKa = –logKa
Fundamentals of Titration 15
3 Titration reaction
The reaction of the base A- with water can be described in an analogous manner:
The corresponding basicity constant Kb follows from the law of mass action:
-
HA • OH
= Kb
-
A
+ -
K a • K b = H 3O OH = Kw
Polyprotic acids or polyequivalent bases that donate (accept) protons in steps have a separate
acidity (basicity) constant for each ionization step.
16 Fundamentals of Titration
3 Titration reaction
This section contains a summary of the titration reactions important in titration practice.
In the titration of an acid HA with a strong base (e.g. NaOH) the following two chemical
equilibria occur:
+ -
+ - H 3O • A
HA + H2O <===> H3O + A = Ka
HA
+ -
2 H2O <===> H3O+ + OH- H 3O OH = Kw
Acid-base reactions are very fast, and the chemical equilibrium is established extremely
rapidly. Acid-base reactions in aqueous solutions are thus ideal for titrations. If the solutions
used are not too dilute, the shape of the titration curves depends only on the acidity constant
Ka as the following figure shows.
Notes: – Very weak acids are difficult to titrate in aqueous solution. In the figure below
it can be seen that for pKa values greater than 10, the corresponding titration
curve no longer exhibits any jump in the region of the equivalence point.
– Bases can be titrated with a strong acid in an analogous manner. The same
titration curves result if Ka is substituted by Kb and pH by pOH (pH + pOH = pKw
= 14).
– Polyprotic acids (e.g. the first two ionization steps of phosphoric acid) and
mixtures of acids can easily be titrated separately if the acidity constants differ
by at least two pK units.
Fundamentals of Titration 17
3 Titration reaction
pH
13
12 pK = 15
11 pK = 11
10
pK = 9
9
pK = 7
7
pK = 5
5
4
pK = 3
3
pK = 1
V [mL]
1 2 3 4 5 6 7 8 9 10
18 Fundamentals of Titration
3 Titration reaction
Titrations can also be performed in nonaqueous solvents ([1], [2], [3], [4]). The use of
nonaqueous solvents is advantageous under the following conditions:
Like water, each suitable solvent HS for acid-base titrations acts both as an acid and a base:
+ -
+ - HS H • S
Acid: HS <===> H + S Ka =
HS
+
HS SH 2
Base: HS + H+ <===> SH2+ Kb =
+
HS • H
The sum of the above two equilibria gives the autoprotolysis constant KsHS of the medium:
HS + - HS HS
2 HS <===> SH2+ + S- KS = SH 2 S = Ka • Kb
The solvent is thus characterized by the acidity constant KaHS, the basicity constant KbHS and
the autoprotolysis constant KsHS.
– If the acid HA to be determined is very weak, the acidic behavior of the solvent must be less
pronounced than that of water (small KaHS).
– If the base B to be determined is very weak, the basic behavior of the solvent must be less
pronounced than that of water (small KbHS).
– The smaller the autoprotolysis constant KsHS, the greater the potential jump at the
equivalence point.
Fundamentals of Titration 19
3 Titration reaction
– The coefficient of expansion of organic solvents is considerably larger than that of water.
The temperature dependence of the titer can thus be very large (up to 0.2% for a
temperature change of 1°C).
– Many nonaqueous solvents are more volatile than water and are sensitive to CO2. It is thus
essential to check the titer frequently.
+ -
Ag+ + X- <===> AgX(solid) K sp = Ag Cl
X- = Cl-, Br-, I-
– At the start of the titration, the solution may become supersaturated before a precipitate is
formed.
– With solutions that are too concentrated, inclusions of sample and/or titrant may occur in
the precipitating solid, thereby falsifying the result. An effective countermeasure is rapid
stirring.
20 Fundamentals of Titration
3 Titration reaction
3.2.4 Complexometry
Complexometric methods allow the titration of a large number of metal ions. Typical of such
methods is the formation of chelates between the metal ion M and the complexing agent Y.
MY
M + Y <===> MY K MY =
M•Y
The most important complexing agent is the disodium salt of ethylenediaminetetraacetic acid
(EDTA: abbreviation Na2H2Y 2H2O, molar mass: 327.24 g/mol):
•
NaOOC – H2 C CH 2 – COOH
N – CH 2 – CH2 – N • 2 H2 O
HOOC – H2 C CH 2 – COONa
*The illustration is taken from the special edition of MERCK SPECTRUM “Titration and Electrochemistry”.
Fundamentals of Titration 21
3 Titration reaction
Among the applications of the complexometric titration, the determination of water hardness
(Ca, Mg) has achieved the greatest importance.
If two reaction partners can be interconverted by the gain or loss of electrons, a redox system
is present. The process underlying this chemical reaction is called a redox reaction (= electron
shift). The two partners are known as a conjugate redox couple.
One of these entities gives up electrons. This process is called oxidation. The other entity gains
these freed electrons. This process is known as reduction.
Substances that can oxidize other substances are called oxidizing agents. Substances that
can reduce other substances are known as reducing agents.
But since electrons never occur in the free state in perceptible concentration, oxidation and
reduction reactions can only occur together. One reaction releases exactly the same number
of electrons as the other reaction requires. There must thus always be two reaction couples
participating in a redox reaction.
22 Fundamentals of Titration
3 Titration reaction
This redox series shows a representative selection of redox couples that includes not only the
most well-known titrants but also several metals and the halogens. It can be seen immediately
from this table that, for example, metallic iron in copper(II) solutions will be oxidized.
Fundamentals of Titration 23
3 Titration reaction
Manganometry
Manganometry is based on the powerful oxidizing effect of potassium permanganate. The
overwhelming number of redox titrations with KMnO4 are performed in sulphuric acid solutions
according to the following scheme:
Iodometry
One of the most important redox couples is iodide/iodine. The fundamental process
In this manner reducing agents can be determined directly with iodine solution as titrant.
The liberated iodine is titrated with a suitable reducing agent. Sodium thiosulfate is used almost
exclusively today.
24 Fundamentals of Titration
3 Titration reaction
Dichromate method
Chromium with oxidation number +6 is reduced by a large number of reducing agents in acidic
solution. Use is made of this property in the cleaning of glass vessels with chromosulfuric acid.
The dichromate ion Cr2O72- is stable in acidic solution but can be reduced to the chromium(III)
ion Cr3+ in the presence of hydrogen ions by gain of six electrons (three per chromium(VI))
according to the equation
Dichromate as a titrant has gained practical importance in the determination of the chemical
oxygen demand (COD) in wastewater analysis [5]. The COD determination is based on the
oxidation of organic compounds with chromosulfuric acid using silver sulfate catalyst.
Cerimetry
Cerium(VI) sulfate is a powerful oxidizing agent. The oxidation number of cerium changes only
by one:
The cerium(IV) sulfate solution (prepared in H2SO4) has a stable titer and is insensitive to both
light and heat. In contrast to permanganate solutions it can also be used for titrations in highly
concentrated hydrochloric acid solution. It is thus extremely versatile.
NO
Fundamentals of Titration 25
3 Titration reaction
[2] J. Kucharsky and L. Safarik, Titrations in nonaqueous solvents, Elsevier Publishing Company, Amsterdam,
1965
26 Fundamentals of Titration
4 Indications
4 Indication methods
The progress of the titration, the chemical reaction and the determination of the end point must
be observable. Traditionally, the titration was followed visually, usually by addition of color
indicators to the solution as only a few reactions are self-indicating (e.g. reactions with iodine
and permanganate).
Over the years, a great many disadvantages, some of them serious, for example
– only the end point and not the complete titration profile is indicated,
– recognition of the color by human eye is not objective,
– many titrations cannot be indicated visually,
– with color indicators an arbitrary end point of the titration is defined that does not coincide
with the equivalence point,
– the color indicator is also titrated and this distorts the result and
– the cost of the chemicals and sample pretreatment is usually greater than in indication using
an electrochemical sensor
have led to the replacement of visual indication by electrochemical and photometric indication
capable of being automated.
With electrochemical sensors, namely electrodes, charge transfers and charge separations
that arise at phase boundary surfaces can be determined (potentiometry) or generated and
altered by means of an imposed current (voltametry, amperometry).
With photometric sensors the decrease in intensity of a light beam passing through the sample
can be measured at a specified wavelength.
Fundamentals of Titration 27
4 Indications
The processes that occur at an electrode in a galvanic cell form the basis of electrochemical
indication methods.
A galvanic cell comprises two electrodes and one solution or two solutions separated from
each other but connected by an electrically conducting salt bridge (= half cells). Such an
arrangement generates electrical energy through electrochemical processes. Galvanic cells
are also popularly referred to as batteries.
An oxidation takes place at one of the electrodes and a reduction at the other. The electrons
released in the oxidation process in the first half cell are transported across the external bridge
to the other electrode where reduction occurs. There is thus a potential difference between the
two electrodes.
A simple galvanic cell can be demonstrated using the example of the reaction of metallic zinc
and copper ions.
Zn Salt bridge Cu
Zn 2+ Cu
2+
SO4 2- SO4
2-
28 Fundamentals of Titration
4 Indications
The voltage (potential difference) measurable at the voltmeter, the slow dissolution of the zinc
rod and the deposition of copper on the copper rod are proof of the occurrence of the
electrochemical process.
Considerably more important for use in titration are inert electrodes (e.g. platinum) that are not
changed by the redox reaction:
Pt Salt bridge Cu
3+
Fe 3+
2+
Fe 2- Cu2+
2-
SO4 SO4
In the right half cell the copper rod slowly dissolves. At the platinum electrode in the left half
cell Fe3+ ions take up electrons and are reduced to Fe2+ ions. The platinum rod itself remains
unchanged and is thus referred to as an inert electrode.
Fundamentals of Titration 29
4 Indications
It is impossible to measure the potential of a single electrode directly; only the difference in
potential of two electrodes is accessible. The resulting potential of such an electrode assembly
Etot (so-called electromotive force) is given by the difference between the potentials E1 and E2
of the two electrodes:
E tot = E 1 - E 2
The potential of a single electrode depends on the ionic concentration of the solution used to
complete the half cell. This dependency is described by the Nernst equation:
R • T • In 10 Ox
E = E0 + • log
n • F Red
where
E0: is the standard potential ([Ox]/[Red] = 1) of the electrode
R: the molar gas constant
T: the temperature (in K)
n: the number of electrons transferred in the electrode reaction and
F: the Faraday constant.
[Ox] and [Red] are the concentrations of the oxidized and reduced ionic species participating
in the reaction.
59.16 Ox
E = E0 + • log mV
n Red
A change in the concentration ratio by a factor of ten causes a change in the electrode potential
by 59.16/n mV.
The half cells with zinc and copper rods mentioned in the above examples are so-called
electrodes of the 1st kind. Each metal that is immersed in a solution of one of its salts and
can develop a reversible potential is called an electrode of the 1st kind.
30 Fundamentals of Titration
4 Indications
Ag <—> Ag+ + e-
R • T • In 10 +
E = E0 + • log Ag
F
The potential of this half cell depends only on the silver ion concentration [Ag+] in the solution.
A metal that is coated with a layer of one of its sparingly soluble salts and immersed in a solution
that contains the anion of the coating is called an electrode of the 2nd kind.
An example is a silver rod coated with silver chloride immersed in a chloride solution.
Ag Ag / AgCl
3+ 3+
+ -
Ag Cl
Electrodes of the 2nd kind are extremely important as reference systems of reference
electrodes.
Fundamentals of Titration 31
4 Indications
A typical experimental setup in titration comprises one sensing electrode and one reference
electrode. The task of the sensing electrode is to record all changes in the composition of the
solution. The reference electrode must be capable of supplying a stable reference potential
that is independent of these changes.
The nature of the reference system, the frit and the reference electrolyte determines the
properties of the reference electrode. The reference electrode chiefly used today (see
illustration) is an electrode of the 2nd kind described above, with the reference system Ag/
AgCl.
Cable
Head
Filling aperture
Plug-in contact
Reference element
Reference electrolyte
Frit
32 Fundamentals of Titration
4 Indications
The reference system is in the form of a cartridge and contains an ample supply of silver and
silver chloride. The cartridge is connected to the reference electrolyte (e.g. KCl: c(KCl) = 3 mol/
L) via an internal frit.
The external frit ensures electrical contact between the reference electrode and the analysis
solution. It must fulfill the following requirements:
– chemically inert
– low outflow rate of reference electrolyte at low electrical resistance
– no ion exchanger properties.
In addition to fine-pored ceramic frits, sleeve frits made of glass or plastic are used.
A calomel electrode (reference system Hg/Hg2Cl2) is often used as reference electrode and
is constructed on similar lines to the silver chloride reference electrode.
To avoid a reaction of the reference electrolyte with constituents of the sample or titrant (e.g.
Cl- with Ag+), double junction reference electrodes comprising two electrodes of the 2nd
kind are often used.
In routine analysis, combined electrodes have gained wide acceptance. Here, the sensing and
reference electrodes are integrated in the same shaft (glass or plastic) (see Section 4.1.4).
Fundamentals of Titration 33
4 Indications
Metal electrodes, usually manufactured as electrodes of the 1st kind, have a wide range of
uses in titration.
Electrodes of the noble metals platinum and gold are used as redox electrodes. They are
eminently suitable for the indication of redox titrations.
In addition to measurement of the silver ion concentration (silver ion activity), metal electrodes
made of silver can be used for the indication of precipitation titrations (determination of
halogens).
The glass electrode is the most important and most widely used sensor in analysis.
that are in contact on both sides with a solution containing H+ ions develop an electrical
potential that depends on the difference in pH value of the boundary solutions. This
phenomenon is based on the following physicochemical processes:
Each glass membrane of a pH electrode reacts with water to form a hydrated, gel-like layer
(see figure). This hydrated layer is not visible since it has a thickness of only 5 - 500 nm, but
it is of fundamental importance to the operating principle of the glass electrode.
34 Fundamentals of Titration
4 Indications
The basis of the glass membrane is a three dimensional network of silicon and oxygen atoms
with each silicon atom being surrounded by four oxygen atoms and each oxygen atom by two
silicon atoms. The interstitial spaces in this irregular network are occupied by cations to ensure
electroneutrality of the glass membrane.
The alkali ions diffuse into the aqueous solution, leaving behind a virtually completely
protonated Si-O skeleton. The internal glass membrane remains anhydrous.
The number of hydrogen ions in the hydrated layer depends on the silicic acid skeleton and
is constant and independent of the nature of the analysis solution. Owing to the movement of
cations through the glass membrane, the electrical potential at the hydrated layer is transferred
to the inner surface of the membrane where a hydrated layer with a phase boundary potential
is also present. The total membrane potential results from the difference between the two
phase boundary potentials.
Fundamentals of Titration 35
4 Indications
In routine work a glass electrode with integrated reference electrode (so-called combination
or combined electrode) is usually employed.
Filling aperture
Reference electrode
Reference electrolyte (3 M KCI saturated
with AgCI)
Ag/AgCl reference element
Ceramic frit
Sensing electrode
Internal buffer (electrolyte)
Ag/AgCl (internal) lead-off
Glass membrane
When the pH values in the two hydrated layers are identical (ideal case) and the pH value of
the internal electrolyte (see figure) of the glass electrode remains constant, the membrane
potential is given by
2.303 • RT +
E membrane = Constant + • log H
F
36 Fundamentals of Titration
4 Indications
In practice, this equation is never satisfied exactly. Three factors contribute to the nonideal
behavior:
1. Asymmetry potential
The membrane potential should be zero when the glass membrane is in contact with
identical solutions on both sides. Generally, however, a potential of a few mV is observed
owing to the different history of the two sides of the membrane. A small asymmetry
potential is unimportant since it is compensated in the calibration.
2. Alkaline error
In alkaline solutions the H+ ions of the hydrated layer are partly replaced by alkali ions,
especially sodium ions. The smaller H+ concentration in the hydrated layer thus leads to
lower pH values.
The difference between the theoretical and experimental pH value is called the alkaline
error. In the types of lithium-based glass used today, the pH deviations do not start until
above pH 13. The alkaline error increases with increasing pH value, increasing alkali
concentration and increasing temperature.
3. Acid error
In highly acidic solutions (pH < 1) the glass electrode exhibits deviations from the ideal pH
function. Through uptake of acid molecules, the hydrogen ion activity of the hydrated layer
is increased, thereby leading to positive pH shifts. The acid error is less disturbing than the
alkaline error.
E [mV]
Acid error
200
-200
Allkaline error
pH
0 7 14
Fundamentals of Titration 37
4 Indications
A further characteristic of the glass membrane is its high electrical resistance, which lies
between 10 and 10 000 MΩ, depending on the composition of the glass, the temperature and
the size of the membrane. This high resistance thus places increased demands on the
measuring system.
The internal electrolyte ensures a constant phase boundary potential at the inner surface of
the glass membrane and a constant potential at the internal lead-off. Usually a silver wire
coated with Ag/AgCl is used as internal lead-off, whose potential is determined by the chloride
ion activity of the internal electrolyte.
38 Fundamentals of Titration
4 Indications
Ion selective electrodes are electrochemical half cells in which a potential difference arises at
the phase boundary electrode/solution that depends on the concentration (more correctly
activity) of a certain ion in the solution.
Glass electrodes are also ion selective electrodes. The setup of an ion selective electrode
assembly is similar to that of the pH electrode and comprises an ion selective membrane and
a reference electrode of constant potential.
Instead of a pH scale, an ion scale is defined, for example a pNa or a pCl scale.
Many cations and anions, neutral gases such as NH3, CO2 and SO2 and even organic
substances such as amino acids can be measured quantitatively with ion selective electrodes
directly. Even ions or neutral substances that are not measurable directly can be determined
indirectly if a chemical auxiliary reaction is run in which a substance that can be detected by
a sensing electrode is released or bound.
In the ideal case the electrode assembly potential of an ion selective electrode is described
by an expanded Nernst equation, the so-called Nicolsky equation:
ni/nj
E = E0 ± S • log a i + ∑K ij • aj
j
where
E: is the electrode assembly potential
E0 : the electrode assembly potential at the reference point (ai = 1, aj = 0)
S: the slope (S = 2.301 R T/ni F). The sign is + for cations and – for anions.
• • •
The selectivity coefficients of the interfering ions are a measure of the selectivity of the
electrode. They should be as small as possible so that the interfering ions in question make
no appreciable contribution to the ion selective potential change at the measuring cell. With
a value of Kij = 1 the contribution of the interfering ions to the potential change is exactly the
same as that of the analyte ion (assuming same charge numbers). A completely selective
electrode, i.e. one that responds only to one type of ion under all conditions does not exist.
Fundamentals of Titration 39
4 Indications
The glass electrode has a selectivity coefficient towards sodium ions of 10-12 – 10-13, in other
words good pH electrodes are disturbed by sodium ions only at Na+ concentrations greater
than 0.1 mol/L and pH values above 12.
The astonishing selectivity of glass electrodes is by no means attained by other ion selective
electrodes. Selectivity coefficients of 10-5 – 10-6 are typical.
Cable
Electrode head From the schematic setup of such an
electrode it is apparent that the membra-
Shaft ne is the electrode sensor at which the
potential is developed.
Reference electrolyte
The shaft of the electrode is closed at its
bottom end by the membrane and usually
comprises an aqueous reference solu-
tion and an electrode of the 2nd kind.
Various types of material are used for the
membrane.
Reference system
Membrane
The glass membranes of the glass pH electrode are mostly made by glass blowing and fused
onto the electrode shaft.
40 Fundamentals of Titration
4 Indications
All electrochemical measurements have one thing in common: they are performed using an
electrode assembly consisting of a sensing and a reference electrode.
Potentiometry
The direct measurement of the galvanic potential developed by an electrode assembly is
called potentiometry, while the performance of a titration by use of this method is referred to
as a potentiometric titration.
The potential U that develops should be measured, if at all possible, at zero current with a high
impedance signal amplifier for the following reasons:
– The basis of potentiometry is the Nernst equation, derived for electrodes in chemical and
electrical equilibrium. An excessive current flow across the phase boundary surfaces
concerned would disturb this equilibrium.
– A further reason for use of a high impedance measuring input results from the special
construction of pH and ion selective electrodes. The measuring circuit includes the ion
selective membrane, whose electrical resistance can easily be 100-1000 MΩ. If the
experimental error due to the voltage divider effect is to be kept below 0.1%, the input
impedance of the measuring instrument should be at least 1000 times greater. This can be
seen from the following equation:
Relectrode assembly
Error in % = • 100
Relectrode assembly + Rinput
For very high resistance electrodes, signal amplifiers with an input impedance of 1012 Ω are
thus necessary.
Fundamentals of Titration 41
4 Indications
Voltametry
This indication technique involves the currentless measurement of the potential difference
between two polarizable metal electrodes polarized by a small current applied externally
(direct or alternating current).
R Ipol
U
The stabilized power supply source provides the current. The resistance R connected in the
circuit must be selected such that a current Ipol can be generated in the range 0.1 - 20 µA. The
potential U that develops between the electrodes is measured exactly as in potentiometry.
One of the main applications of voltametric indication is the determination of water by the Karl
Fischer method.
42 Fundamentals of Titration
4 Indications
Amperometry
Amperometric indication like voltametric indication makes use of polarized electrodes but uses
a constant potential instead of a constant current. The measured variable is here the current
flowing through the electrodes and the titration solution. The amperometric titration curve is
thus a current-volume curve. The following experimental setup is needed:
Upol
R
I
A constant potential Upol is applied between the two electrodes with a voltage divider. The
resulting current I is measured with a microammeter.
Fundamentals of Titration 43
4 Indications
I I
T = ( or % T = • 100 )
I0 I0
where
T: is the transmission
I0: the intensity of the incident light and
I: the intensity ot the transmitted light.
If all light is absorbed, then I = 0 and hence T = 0. If no light is absorbed, I = I0 and T = 1 (or
%T = 100%).
In photometry, work is frequently performed using absorption as the measured variable. The
relation between transmission and absorption is described by the Bouguer-Beer-Lambert
Law:
A = –log T = ε • c • d
where
A: is the absorption
ε: the absorptivity
c: the concentration ot the absorbing substance and
d: the path length of the light through the solution.
From the above relation it can be seen that there is a linear relation between absorption A and
concentration c. This is the basis of the direct photometric measurement.
44 Fundamentals of Titration
4 Indications
– they are easier to use (no refilling of electrolyte solutions, no clogging of the frit)
– longer lifetime (they are virtually unbreakable)
– they can be used to perform all classical titrations to a color change (no change in traditional
procedures and standards).
Fundamentals of Titration 45
4 Indications
The METTLER DP550 and DP660 Phototrodes are probes for photometric titration in the
visible region.
Compared with traditional titration instruments with photometers, the phototrodes are distin-
guished primarily by integration of the light source and signal processing in the probe. The
measurement principle is shown schematically in the following figure:
6 Connection
7 Control knob
1 Photodiode
5 Detector
2 Light guide
3 Sample liquid
4 Concave mirror
The photodiode (1) built into the probe emits modulated light that passes through the sample
liquid (3) via light guide (2). The light reflected by the concave mirror (4) is converted by the
detector (5) into an electrical signal that is amplified and led to the titrator via connection (6).
The signal amplification can be adjusted by means of a control knob (7). Sunlight and artificial
lighting have no influence on the measurement, as interference due to external light sources
is virtually completely eliminated thanks to the high frequency light modulation.
46 Fundamentals of Titration
4 Indications
Conductometric indication [1], [2] [3] makes use of the ability of aqueous solutions to conduct
an electric current. This conductivity is based on the dissociation of acids, bases and salts into
electrically charged species (ions) in aqueous solution. In an electric field the anions migrate
to the positively charged anode and the cations to the negatively charged cathode. Faraday’s
law states that per mole equivalent entity the same quantity of electricity, namely 96'485
coulombs, will be transported to the electrodes.
I
R = ρ•
q
The proportionality factor ρ is called the resistivity. The following relation holds between the
conductivity and the resistivity:
ρ •χ = 1
The conductivity is thus obtained from the measured resistance R and the dimensions of the
conductivity cell:
1•I
χ = = G•Z
R q
The factor 1/R is also known as the conductance G. The conductance has dimensions µS or
mS (S = Siemens). The quantity 1/q is referred to as the cell constant Z. The cell constant has
dimensions of cm-1. Typical values of cell constants are between 0.1 and 10 cm-1. It is always
specified by manufacturers of conductivity cells and should be selected to match the
concentration of the solution being titrated.
Fundamentals of Titration 47
4 Indications
The conductivity should be specified in units of µS/cm or mS/cm, depending on its magnitude.
For the determination of the conductivity, alternating current must be used for the resistance
measurement. If a direct current flows between the electrodes, electrolysis takes place and the
contribution of the ohmic resistance, which is the sole variable of interest, becomes so small
that its measurement is impossible.
+
H
OH—
+
Na
Cl —
V
VEQ
48 Fundamentals of Titration
4 Indications
The measured conductivity at every point on the titration curve is composed of the sum of the
conductivity of the individual ions. The titration diagram overleaf shows the contributions of the
individual ions to the total conductivity (dilution not taken into account).
The titration curves are actually straight lines as long as each ionic species present reacts
quantitatively or not at all. The typical curve character is due to the fact that one ionic species
of the solution disappears (in our case H+) only to be replaced by a new one from the titrant
(here OH-). When the equivalence point is exceeded, an increase in the conductivity is always
observed in the absence of any further reaction.
Fundamentals of Titration 49
5 Types of titration
5 Types of titration
Titrations can be classified in various ways. The classification by means of indication method
and analytical reaction has been discussed in earlier sections. This section describes the
classification of titrations according to the manner in which they are performed.
In the direct titration the titrant reacts directly with the analyte. The performance of a direct
titration can be represented as follows:
Amount of analyte
Q
Result: indicated equivalence point
through equivalent amount of
titrant
Under the experimental conditions usual in the practical procedure, not every reaction fulfills
the requirements described in section 1 for the titration reaction. Further, under certain circum-
stances indication of the equivalence point may also be poor. In such cases an indirect method
is often employed to obtain the result.
50 Fundamentals of Titration
5 Types of titration
In a back titration an excess of titrant is added to the sample. After a sufficiently long waiting
time, this excess is then backtitrated with a second titrant. The difference between the added
amount of the first and second titrant then gives the equivalent amount of the analyte. The back
titration is used mainly in cases where the titration reaction of the direct titration is too slow or
direct indication of the equivalence point is unsatisfactory.
Amount of analyte
Q1
Amount of titrant 1 (excess)
Q2
indicated equivalence point
through amount of titrant 2
Q1 - Q2
Result: calculated equivalent
amount that has reacted
with the analyte
Fundamentals of Titration 51
5 Types of titration
By initial addition of a metered volume of titrant followed by titration with the sample solution
(= reverse of titration), the titration reaction may, under certain circumstances, be faster than
in the direct titration. The classic example of an inverse titration is the determination of sugar
using Fehling’s solution.
Q
Result: indicated equivalence point
through equivalent amount of
the analyte
52 Fundamentals of Titration
5 Types of titration
The action of the substitution titration is based on the addition of a reagent to the sample
solution that reacts with the analyte. Here, a component of the added reagent is released in
a stoichiometric amount and is then determined by direct titration.
Amount of analyte
stirring/warming
Q
Result: indicated equivalence point
through equivalent amount
of titrant
Fundamentals of Titration 53
5 Types of titration
A B C
Amount of components A + B + C
Q
Result: indicated equivalence point for the
sum of the components A,B and
through equivalent amount of
titrant
54 Fundamentals of Titration
5 Types of titration
A B C
Amount of components A + B + C
Masking of C
Q
Result: indicated equivalence point for the
sum of the components A and B
through equivalent amount of
titrant
Acid-base titrations
Selective acid-base sequential titrations are possible when the pKs of the different acidic
or basic components differ by at least two units. The choice of a nonaqueous solvent often
allows an improved differentiation.
Fundamentals of Titration 55
5 Types of titration
Complexometric titrations
In complexometric sequential titrations the following criteria must be met:
– the effective stability constants must show a difference of at least five logarithmic units
– the minimum value of the relevant stability constant (in logarithmic units) must be at least
seven.
Redox titrations
Selective redox titrations are also possible. The potential difference between the equivalen-
ce points in question must be at least 300 mV.
A B C
Amount of analytes A + B + C
Q1 Q2 Q3
Result A, B, C:
indicated equivalence points 1,2 and 3
through equivalent amount of titrant in
each case
56 Fundamentals of Titration
5 Types of titration
A B
Amount of analytes
A+B
Increase in the pH
Q2
Result B: indicated equivalence point 2 through
equivalent amount of titrant
A B
Amount of analytes
A+B
Masking of B
Q1
Result A: indicated equivalence point 1 through
equivalent amount of titrant
Demasking of B
Fundamentals of Titration 57
5 Types of titration
A B
Amount of analytes
A+B
Q2
Amount of titrant 1 (exess)
stirring/warming/pH change
Q2 - Q3
Result B: calculated equivalent amount of
component B
58 Fundamentals of Titration
6 Titration curves
6 Titration curves
Titration curves illustrate the qualitative progress of a titration. They allow a rapid assessment
of the titration method. A distinction is made between logarithmic and linear titration curves.
Logarithmic curves appear when the measured signal has a logarithmic dependence on the
equilibrium concentration. Such curves are obtained with all indication methods that follow the
Nernst equation. This includes all potentiometric titrations.
If there is a linear relation between the measurement signal and the concentration, the titration
curves are said to be linear. The most important application is photometric titration. Further
examples are titrations with conductometric, amperometric and thermometric indication.
Linear curves can be evaluated graphically and by calculation.
This representation can be used for the graphical determination of the equivalence point (see
section 8). The various potentiometric indication methods produce in part very different
titration curves in the range -1600 mV to +1600 mV.
Fundamentals of Titration 59
6 Titration curves
60 Fundamentals of Titration
6 Titration curves
In titrations of mixtures the extreme values of the first derivative of the curve are in some cases
not readily apparent when the titration curve comprises flat jumps in addition to a very steep
jump. For such cases the following logarithmic representation of the first derivative has proved
useful:
sign(x) = 1, if x > = 0
sign(x) = –1, if x < 0
This representation gives greater prominence to small maxima over relatively large ones.
Fundamentals of Titration 61
6 Titration curves
∆E/∆V
log ∆E/∆V
62 Fundamentals of Titration
6 Titration curves
Representation of the signal as a function of time (plotted on an analog recorder or a dot matrix
printer) aids the development of new methods and the optimization of the equilibrium condition
for the measured value acquisition. The time dependance of the measurement signal allows
assessment of the response behavior of the electrode and the rate of the titration reaction.
The following figures show a few representative examples. The shape of the curve is
influenced by the following parameters:
This example shows the case of a slow reaction observed using an electrode with a fast
response. The abrupt change in the signal at the start shows the immediate response of the
electrode to addition of the titrant. The stabilization of the equilibrium signal is the result of the
subsequent titration reaction.
Fundamentals of Titration 63
6 Titration curves
This example is typical of a rapid reaction with an electrode having a rapid response. The
electrode can follow the chemical reaction instantaneously, hence the exponential profile of
the time signal.
A rapid reaction with the use of a slowly responding electrode is demonstrated by this example.
It will appear often when work is performed with a dirty or poorly maintained electrode. The
electrode does not respond until after a certain incubation time, but meanwhile the chemical
reaction is already well advanced.
64 Fundamentals of Titration
6 Titration curves
The concepts of “rapid” and “slow” are naturally relative in this context. They describe the rate
of the titration reaction relative to the response behavior of the electrode.
This form of the titration curve – especially the first derivative dV/dt = f(t) – is an important
representation of pH-stating reactions and Karl Fischer titrations.
– The amount of titrant added as a function of time describes the progress of the chemical
reaction under investigation. The rate of titrant addition dV/dt is directly proportional to the
reaction rate.
– In Karl Fischer titrations the representation of the rate of titrant addition – expressed in µg
H2O/min – as a function of time allows a simple assessment of the drift before and above
all after the titration.
Example of a V/t curve: Determination of the activity of pancreatic lipase 250 (pH-stating).
V [mL]
t [min]
2 4 6
Fundamentals of Titration 65
7 Control
A titration curve is represented by the measured signal E (unit: mV or a quantity derived from
it such as pH) and the volume V of the added titrant (unit: mL). The signal describes the
dependence of the progress of the titration reaction on the titrant addition.
E [mV]
V [mL]
In modern titrators titrant addition and measured value acquisition are intimately linked by a
control system (with the aim of providing an accurate and reproducible titration result within
the shortest possible time).
66 Fundamentals of Titration
7 Control
The titrant can be added in two ways: continuously at a defined rate of dispensing or
incremental with individual volume steps.
The continuous addition of titrant is the classic way to perform a titration with motorized piston
burette, pH meter and analog recorder. In automatic titrators the continuous titrant addition is
used in the so-called recording titration and in end point titrations. In the recording titration
the signal is plotted on an analog recorder or a dot matrix printer as a function of the added
volume.
The dispensing rate must be matched to the rate of the titration reaction and the response time
of the electrode. Particularly in cases where a large potential change occurs in the vicinity of
the equivalence point, diffusion phenomena appear at the frit and lead to a delay in stabilization
of the potential. If the dispensing rate is too high, with a slow reaction the result is too low and
with a slow response time of the electrode too high.
E [mV]
slow electrode:
slow reaction:
results too high
results too low
rapid reaction,
rapid electrode:
EP correct results
V [mL]
Fundamentals of Titration 67
7 Control
Modern titrators solve the problem with a variable dispensing rate. The titrant addition is
controlled as a function of the measured signal change such that a distortion of the titration
curve due to lag of the potential adjustment is avoided even in the transition interval.
The titrant is added at a high rate up to a defined control band. Within the control range the
rate decreases exponentially. In the vicinity of the end point single pulses are dispensed. A
pulse is the smallest increment and equals 1/5000 of the burette volume of METTLER burettes.
E [mV – pH]
5
+ 100
8
-100 End point
9
10
-200
V [mL]
A large control range leads to an accurate but slow titration. A narrower range results in a rapid
titration, but there is an inherent danger of overtitrating. With an S-shaped, steadily progres-
sing titration curve only about the last 10% of the volume to be added should lie within the
control range. To avoid overtitrating, it is advisable to position the burette tip so that the stirrer
transports the titrant to the electrode by the shortest route.
68 Fundamentals of Titration
7 Control
The time from the attainment of the end point up to the definitive termination of the titration is
called the delay. If the signal of the sample solution deviates from the original end point signal
during this time, additional increments are added until the end point is again reached. A large
value of the delay (typical value: 10 s) should be selected with:
The continuous titrant addition in end point titrations is suitable only for steep titration curves.
With flat curves (see figure), wrong selection of the end point (EP’ instead of EP) or a drifting
electrode leads to a false equivalence volume (VEQ’ instead of VEQ). For traditional reasons
(old standards, procedures), however, even when the curves are flat, titration must sometimes
be performed to the preset end point by means of continuous titrant addition. Before such
determinations the appropriate electrode must always be calibrated to allow detection of the
exact end point.
E E
EP
EP‘
V V
VEQ VEQ › VEQ‘
Fundamentals of Titration 69
7 Control
The accuracy of the result of a titration depends principally on the quality of the experimental
data available for the evaluation. This has led to the development of a new type of reagent
addition, the incremental titration.
The titrant is added in single volume steps. After dispensing, the change in the signal is
recorded accurately. Thus, for each dispensing step, the measurement method furnishes a
measured value, a data point, of the titration curve.
If the addition is effected with constant, relatively large volume steps, with steep titration curves
there are very few data points in the vicinity of the equivalence point (see figure).
E [mV]
∆E14
∆E13
∆E12
V [mL]
∆V12 ∆V14
∆V = konstant
70 Fundamentals of Titration
7 Control
With smaller volume increments, more data points are obtained, but the titration takes longer
and in the steep part of the curve, important for the calculation, the increase in the number of
data points is relatively small.
E [mV]
∆E27
∆E26
∆E
25
V [mL]
∆V23 ∆V27
∆V = konstant
Fundamentals of Titration 71
7 Control
An obvious move is thus to size the titrant addition so that the signal change always remains
about the same. This type of titrant addition is known as dynamic titrant addition and the
corresponding titration control as dynamic titration. In the flat, uninteresting part of the titration
curve, large volume increments are dispensed and in the steep part of the curve, important for
the calculation, smaller increments.
E [mV]
∆E ~ ∆E(set)
∆E5
∆E4
∆E3
∆V3 ∆V4
V [mL]
72 Fundamentals of Titration
7 Control
The last three data points of the current titration and the signal change ∆E(set) aimed for are
used to calculate the next increment ∆V by extrapolation.
E [mV]
∆E ~ ∆E(set)
3rd data point
∆E2 ∆V
2rd data point
∆E1 ∆V2
1st data point
∆V1
V [mL]
The formula proposed by Ebel [1] has well proved its worth in practice:
∆E 1
m1 =
∆E ( set ) ∆V 1
∆V =
( 2 • m1 – m2 ) ∆E 2
m2 =
∆V 2
If the calculated ∆V is negative, the following formula is used:
∆E ( set )
∆V =
( 2 • m2 – m1 )
In addition, a lower limit ∆V(min) and an upper limit ∆V(max) are defined for the calculated ∆V.
The choice of ∆E(set), ∆V(min) and ∆V(max) depends on the shape of the titration curve.
The value of ∆E(set) that should be chosen is defined by the height of the jump. For a jump
height of e.g. 250 mV, a typical value of ∆E(set) would be 10 mV.
With very steep titration curves the calculation of ∆V could in some cases result in very small
volume values (fractions of a single burette pulse). This can be prevented by judicious
selection of ∆V(min), e.g. 0.01 mL.
Fundamentals of Titration 73
7 Control
On very flat parts of the titration curve the calculation of ∆V can lead to nonsensically large
volume increments. Limitation of ∆V(max), e.g. 0.5 mL, can prevent this.
For a better understanding of these parameters, see following table of measured values
obtained for a titration of an H3PO4 solution with NaOH (0.1 mol/L).
Titration curve
EQP 1
EQP 2
74 Fundamentals of Titration
7 Control
2 4
1. The added volume increment never exceeds the value of ∆V(max) in the flat part of the
curve.
2. The added volume increment is never less than the value of ∆V(min) in the steep part of
the curve.
3. The signal change is always around 10 mV in the middle part of the curve thanks to the
dynamic control.
4. The time between two increments varies from 3 s (t(min)) to 30 s (t(max)), depending on
the parameters of the equilibrium-controlled measured value acquisition.
Fundamentals of Titration 75
7 Control
If the titration curve exhibits a sharp kink before the equivalence point, a very small value of
∆V(max) must be selected to ensure that the jump is not missed. The value of ∆V(max) then
approaches the value of ∆V(min). In such cases it is appropriate to perform the titration with
small, constant volume increments.
E [mV]
V [mL]
76 Fundamentals of Titration
7 Control
7.1.3 Predispensing
Predispensing can shorten the time needed for titration considerably. There is no sense to
acquire accurate data points during the initial stage of the titration if such data are not needed
for the evaluation. The following predispensing modes are possible:
The first two methods need no signal measurement. The two last methods take into account
a variable amount and a variable content of a sample, but necessitate control of the titrant
addition with simultaneous measurement of the signal. The titrant addition can be continuous
or incremental here. Exact signal measurement is not necessary; the important factor is a rapid
predispensing.
The second method, predispensing to nominal content, also takes a variable sample quantity
into consideration, but requires no signal recording. If the expected value of the sample content
(= nominal content) remains within certain limits, the volume V to be dispensed, that corres-
ponds to a partial content of the sample, can be calculated taking into account the amount
(weighing or initial volume) of sample:
where
metered amount: amount to be dispensed in % of the nominal consumption
nominal content: nominal content of the sample (any unit)
m (or U): weight (or volume) of the sample
C: calculation constant for the conversion from mmol to the unit of the nominal
content, e.g: conversion to unit %: C = M/10 z •
If a dynamic titration follows the fixed predispensing, it is advantageous to split the predispens-
ing into several steps (e.g. three) with simultaneous measurement of the signal after each step.
This procedure allows optimum calculation of the first increment after the predispensing.
Several predispensing modes can be used at the same time and performed one after the other.
Fundamentals of Titration 77
7 Control
After every addition of a volume increment a measured value must be acquired. This can occur
in two ways:
– timed-increment or
– equilibrium-controlled.
E [mV]
Measured value
t [s]
∆t1 = 3s ∆t2 = 3s ∆t3 = 3s
78 Fundamentals of Titration
7 Control
In the equilibrium-controlled measured value acquisition [2] the measured value after
every increment addition is first acquired when an equilibrium has been established in the
solution and the signal is stable.
The measured value is then acquired when the potential no longer changes by a specified
amount ∆E within a specified time ∆t, in other words the measured drift of the electrode
potential must be less than the defined quotient ∆E/∆t (time window) during the period ∆t. This
equilibrium condition can be established at the earliest at the defined time t(min) and should
be established at the latest at the defined time t(max). At t(max) the measured value is acquired
at all events, even if the equilibrium condition has not yet been met.
The measured values are acquired in this mode at different times. If the signal change after
increment addition is small or the equilibrium is established rapidly, the measured value is
acquired immediately after t(min) is exceeded. If the signal change is large or the equilibrium
is established slowly, the waiting time is correspondingly longer. This ensures optimum
matching of the measured value acquisition to the chemical reaction and to the response
behavior of the sensor during the entire titration.
Default values for the parameters of the equilibrium-controlled measured value acquisition
are:
∆E = 0.5 mV
∆t = 1 s
t(min) = 3 s
t(max) = 30 s
For definition of the optimum equilibrium condition, observation of the signal as a function of
time is necessary.
Note: If the measured signal is very unstable (e.g. owing to noise, large fluctuations in
the response behavior of the sensor, etc.) or if it drifts, preference should be given
to the timed-increment measured value acquisition.
Fundamentals of Titration 79
7 Control
E [mV]
164
163
162
∆t = 2 s
161
a ∆E = 1 mV
160
159
∆t = 2 s
158
a ∆E = 1 mV
157 t(min)
156
155 b
154
t(min)
153
b
152
Increment addition
t [s]
0 1 t(min) 5 0 1 t(min) 5 7
b: The equilibrium condition has been met for the first time after 5.4 or 6.9 s.
80 Fundamentals of Titration
7 Control
With electrodes having a slow response or if the analytical reaction is slow, a sufficiently large
value of time t must be selected. With a large minimum time t(min), premature acquisition of
the measured value in the case of an oscillating signal profile can be avoided (see section 6.2).
E [mV]
120
119
∆t =1 s
118
a ∆E = 0.5 mV
117
116
t(min)
115
114
b
113 Increment addition
112
t [s]
0 1 3 5 t(min) 8 0
b: The equilibrium condition has been met for the first time after 8.5 s.
Fundamentals of Titration 81
7 Control
Termination of the titration at the proper time can be triggered by selection of various
parameters:
[1] S. Ebel and B. Beyer, Fres. Z. Anal. Chem., 312, 346 (1982).
82 Fundamentals of Titration
8 Equivalence point determination
The end of a titration is reached when an amount of titrant equivalent to the substance being
analyzed (the analyte) has been added. From the volume of the titrimetric solution required to
reach this point – the equivalence point – and its known concentration, the amount of analyte
can be calculated if the stoichiometry of the reaction is known. The correctness (precision and
accuracy, see Section 10) of the result depends to a large degree on the method chosen to
detect the equivalence point. The methods for recognition and exact calculation of the
equivalence point are treated in this section.
In the immediate vicinity of the equivalence point, those titration curves that can be evaluated
in practice exhibit a change in either the slope (so-called linear or segmented titration curves,
e.g. titrations with amperometric, conductometric and photometric indication) or the direction
of curvature (so-called logarithmic or S-shaped titration curves, e.g. titrations with potentiome-
tric and voltametric indication). These break or inflection points are influenced by the
equilibrium constants of the titration reaction, the initial concentrations, the change in volume
due to the amount of added titrant and other factors. Each form of a titration curve in theory
requires a specific type of equivalence point determination if optimum accuracy of the analysis
result is to be achieved.
The end point of a potentiometric titration is often taken as the inflection point of the titration
curve. The example of the titration of a strong acid with a strong base will be used here to
demonstrate that determination of the inflection point provides a serviceable approximation of
the equivalence point.
To represent the titration curve, the dependence of the H3O+ concentration or the pH value
on the amount of added titrant is needed. The following table shows the H3O+ concentration
or pH value and the extent of titration as a function of the amount of sodium hydroxide
of concentration 0.1 mol/L added to a solution of 50 mL hydrochloric acid of concentration
0.01 mol/L.
Fundamentals of Titration 83
8 Equivalence point determination
From the pH values and their differences shown in the table, it follows that even when the
dilution due to addition of the titrant is taken into account, the titration curve is practically
symmetrical in the region of the equivalence point (extent of titration ~1) and exhibits an
inflection point that coincides with the equivalence point within experimental error.
In titrations of strong acids with strong bases, the difference in volume between the
equivalence point and inflection point depends only on the dilution and is negligibly small. If
the titration curve is highly asymmetric, the error in the determination of the inflection point can
be so large that another type of calculation of the equivalence point becomes necessary (e.g.
with heterovalent redox or precipitation titrations).
84 Fundamentals of Titration
8 Equivalence point determination
Before any attempt is made to calculate equivalence points exactly, they must first be located
by means of the measured data points.
With S-shaped titration curves the inflection point of the titration curve can be used as a
criterion of the equivalence point recognition. The first derivative of the titration curve is used
here.
∆E/∆V
If the absolute value of the slope between the individual data points increases at least twice
before a maximum and then decreases at least twice after a maximum, it is virtually certain that
the titration curve shows an inflection point (situation a).
Situation b is unclear. Either the maximum represents a true, but ill-defined inflection point or
it has been caused by an error in the recording of a data point (e.g. disturbance in the
measurement process).
The recognition of inflection points in this manner is not always unequivocal and is influenced
among other things by the data point density (size of the volume increments), the stability of
the measured signal and the magnitude of the signal changes between the individual data
points.
Fundamentals of Titration 85
8 Equivalence point determination
The degree of uncertainty in the recognition of the equivalence point can be lessened
appreciably by:
The tendency defines the titration direction. This can thus be used to filter out all equivalence
points whose titration direction is not in accordance with the tendency. The following illustration
shows a schematic representation of a typical titration curve of a photometrically indicated
surfactant determination. With the aid of the tendency, only one of the two equivalence points
possible in principle will be evaluated.
E Equivalence point
No equivalence point
(wrong tendency)
Tendency
86 Fundamentals of Titration
8 Equivalence point determination
The threshold value (unit: mV/mL) allows suppression of flat jumps with S-shaped curves.
All maxima in the first derivative that lie below the threshold value are ignored (see illustration).
∆E/∆V
Equivalence point
No equivalence point
(below threshold)
Threshold
Fundamentals of Titration 87
8 Equivalence point determination
With the equivalence point range a measured value range is defined, within which the
equivalence points must occur. All equivalence points outside this range are not evaluated
(see illustration).
Equivalence point
quivalenzpunktbereich
Equivalence point range
No equivalence point
(outside the equiva-
lence point range)
The combination of these criteria leads to a high degree of reliability in the recognition of
equivalence points.
88 Fundamentals of Titration
8 Equivalence point determination
Equivalence points of segmented curves are recognized in a similar manner. If the first
derivative of a segmented curve is calculated, an S-shaped curve analogous to logarithmic
titration curves is obtained. In this case it is not the first, but rather the second derivative that
is used in location of the maximum and in the application of the threshold value criterion.
E ∆E/∆V ∆2E/∆V2
V V V
Fundamentals of Titration 89
8 Equivalence point determination
Many procedures have been described to determine the equivalence point of a titration. They
can be classified into three different groups:
– approximation procedures
– interpolation procedures
– mathematical procedures.
The aim of all these methods is to determine the equivalence point from the values of the
potential by calculation.
The approximation procedures take into account only a few data points in the region of the
equivalence point. Mathematical knowledge of the profile of the titration curve is not required.
These procedures are thus independent of the nature of the titration reaction and the
indication. Consequently, they do not determine the true equivalence point.
Mathematical procedures are methods to determine the equivalence point that require a
mathematical description (physical model) of the titration progress. These procedures
determine the true equivalence point. The model to be used depends on the nature of the
titration reaction and the indication. The mathematics are usually very time-consuming, and
the evaluation requires extensive computation.
The approximation procedures discussed here presuppose constant volume steps. The
equivalence point is determined from the three largest potential changes.
The traditional procedure of Kolthoff and Furmann [1] starts with the assumption that
equivalence point and inflection point coincide and the titration curve is symmetrical about this
point. The inflection point is given mathematically by the zero crossing of the second derivative.
It is also assumed that the calculated quotients of the differences ∆E/∆V (3 values) or ∆2E/∆V2
(a positive and a negative value) describe the first and second derivative of the titration curve
with sufficient accuracy. From the positive and negative value of the second derivative thus
determined, the zero point is calculated by means of linear interpolation.
90 Fundamentals of Titration
8 Equivalence point determination
The underlying mathematical concept is based on the similarity of the titration profile and the
inverse function of the hyperbolic-trigonometric functions. This relation is treated in detail in
the following section on mathematical procedures to determine the equivalence point.
Wolf [3] and Keller–Richter [4] have described numerical approximations of these nomo-
grams.
With the introduction of an auxiliary variable r (see below), the equivalence point can be
calculated for all procedures using the following equation:
VEQ = Vmax + a ∆V + b r ∆V
• • •
E [mV]
∆E n ∆E 2
∆E0
∆E ∆E1
v
∆V ∆V ∆V
V [mL]
V max
∆E0, ∆E1 and ∆E2 designate the three potential steps in descending order of magnitude.
Fundamentals of Titration 91
8 Equivalence point determination
∆E 0 – ∆E 1 1 – R1
Kolthoff–Furmann [1] r = = ––––
2∆E 0 – ∆E 1 – ∆E 2 2 – R1 ( 1 + R2 )
The procedure of Keller–Richter can be employed only under the specified condition. If this is
not fulfilled, two potential values can be combined, i.e. ∆E0 + ∆E1 becomes a new ∆E0.
In addition to the condition of constant volume steps, these procedures have the further
disadvantage that only four data points in the vicinity of the equivalence point are enlisted for
the evaluation. Thus, the only data points used in the calculation are the very points usually
associated with errors caused by time-dependent phenomena (response time of the electrode,
kinetics of the analysis reaction, etc.).
It must be pointed out that strictly speaking all approximation methods mentioned here apply
only to the titration of a strong acid with a strong base and to isovalent precipitation titrations.
Nevertheless, good results have also been obtained with other titrations.
A systematic study by Ebel [5] of the accuracy of these approximation procedures showed that
the method of Keller-Richter provided the most reliable results.
92 Fundamentals of Titration
8 Equivalence point determination
Interpolation procedures include in the evaluation either all data points or those that are
relevant for the interpolation. They are thus less strongly influenced by errors in the
measurement of the potential in the vicinity of the equivalence point.
A well-known procedure is the interpolation of the titration curve using model functions,
particularly the so-called spline functions. The inflection point is determined from the
maximum of the first derivative or the zero crossing of the second derivative of the approxima-
ted titration curve.
The method fails with highly asymmetric curves. The difference between the inflection point
and the true equivalence point is then too large and leads to a systematic evaluation error.
Tubbs procedure
Tubbs [6] has described a graphical procedure for the evaluation of asymmetric, analog
recorded titration curves. It has well proved its worth in routine analyses, as titration curves
often do not exhibit the theoretical profile predicted by a mathematical model (e.g. precipitation
and redox titrations).
M2
Equivalence point
M1
Fundamentals of Titration 93
8 Equivalence point determination
A circle of curvature of minimum radius can be drawn in both branches of the titration curve.
The ratio of the two radii is determined by the asymmetry of the curve. The intersection point
of the straight line drawn between the midpoints M1 and M2 of the circles with the titration curve
represents the equivalence point. Theoretical calculations show that the true equivalence
point with asymmetric titration curves always lies between the inflection point and that branch
of the titration curve with the greater curvature (the smaller circle of curvature). The result of
the Tubbs evaluation approximates this true equivalence point very closely when the titration
curve profile is regular and allows calculation of the circles of curvature of the two branches.
A mathematical variation of the Tubbs procedure for digitally recorded titration curve has been
described by Ebel [7].
This involves approximation of those parts of the titration curve that lie in the region of the
greatest curvature by a hyperbola. For each approximated hyperbola the vertex is determined.
This point on the hyperbola lies at the position of greatest curvature. The midpoints of the
assigned smallest circles of curvature are the foci of the two hyperbolae. As in the graphical
version, the intersection point of the straight line joining the two foci with the titration curve gives
the equivalence point.
The evaluation requires at least six measured points in the region of greatest curvature, both
before and after the inflection point of the titration curve (see illustration).
94 Fundamentals of Titration
8 Equivalence point determination
Y = f(X; P1,P2,P3,...)
with the aid of N experimental data pairs (xi, yi) so that the sum of the squares of the errors
N
∑(y – f(x ; P
2
S = i i 1 ,P 2 , P 3 ,…))
i=1
The time required for the mathematics and computation is enormous. Optimization of functions
with more than four or five parameters exceeds the limits of possibilities of the titrators on the
market today.
E = f(V)
such that there is a linear relationship between the new variables X and Y:
E = f(V) : E —> Y
V —> X : —> Y = A X + B
•
Depending on the model, the equivalence point can then be calculated from one of the
following quantities:
– Slope A
– Intercept B (Y axis)
– Intercept -B/A (X axis)
This procedure was first used by Gran [10] to evaluate acid-base titrations.
Fundamentals of Titration 95
8 Equivalence point determination
Example: Titration of a strong acid (e.g. HCl) with a strong base (e.g. NaOH)
In aqueous solution HCl and NaOH are completely dissociated throughout the titration:
+ + - -
H 3O + Na = CI + OH
During the titration, the mass balances of the ions that do not participate in the titration reaction
are:
where
CONC: the concentration of the titrant NaOH [mol/L]
V0: the initial volume of the titration [mL]
V: the volume of titrant added [mL]
VEQ: the titrant consumption up to the equivalence point [mL]
Insertion of the mass balances in the charge balance equation and solving for [H+] ([H+] =
[H3O+]) gives a quadratic equation
2
+ + CONC • ( VEQ – V )
H – H • – Kw = 0
V0 + V
96 Fundamentals of Titration
8 Equivalence point determination
+
E = E 0 + S • pH = E 0 – S • log H
2
CONC • ( VEQ – V ) CONC • ( VEQ – V )
+ + 4K w
V0 + V V0 + V
E = E 0 – S • log
2
This equation has five parameters: the electrode parameters E0 and S, the consumption of the
titrant VEQ up to the equivalence point , the initial volume V0 to take into account the dilution
and the ionic product of water Kw.
This equation for the titration curve profile can also be described with the aid of the inverse
function of the hyperbolic-trigonometric function:
Division of
- + + - CONC • ( V – VEQ )
OH – H = NA – CI =
V0 + V
+
Kw H CONC • ( V – VEQ )
– =
+ Kw Kw • ( V0 + V )
H
+
Kw H Kw Kw pH – pK w / 2
– = exp In – exp – In = 2 sinh
H
+ Kw H
+
H
+ log e
Fundamentals of Titration 97
8 Equivalence point determination
pK w -1 CONC • ( V – VEQ )
pH = + ( log e ) sinh
2 2 K w• ( V 0 + V )
-1 CONC • ( V – VEQ )
E = α + β • sinh
2 Kw • ( V0 + V )
where α = E0 + pKw/2
and β = S • (log e).
If only a few data points are considered in the region of the equivalence point (V ~ VEQ), the
denominator in the sinh-1 function can be combined in a single parameter g:
-1 CONC • ( V – VEQ )
E = α + β • sinh
γ
where γ = 2 K w • (V 0 + VEQ)
The equation for the titration curve is thus reduced to four parameters α, β, γ and VEQ.
The equation with the four unknowns can be solved exactly with four data pairs (Ei, Vi; i=1,...4).
From the four measured values, the parameter α can be eliminated through formation of three
differences ∆Eij and the parameter β through formation of two differential quotients ∆Eij/∆Ekl.
The resulting system of equations with the two unknowns γ and VEQ cannot be solved for VEQ
[2]. The value of VEQ must therefore be determined iteratively using the procedure of Newton.
The restriction to constant volume steps found in the procedure of Keller–Richter does not exist
in this method. It can thus also be used for titrations with dynamic titrant addition.
A wrong measurement point due to a disturbing influence on the measured value acquisition
has a very great effect on the result of this procedure. It is even possible that the iteration can
fail and no result is found.
98 Fundamentals of Titration
8 Equivalence point determination
The effect of this error source on the evaluation is less if the procedure of nonlinear regression
described above is used. Owing to the smoothing properties of the regression calculation, a
single erroneous measurement point is of less consequence than if an attempt is made to solve
the system of equations exactly.
The DL70 uses this method as a standard procedure for the evaluation of S-shaped titration
curves. The equivalence point is determined using the Marquardt procedure [9] and the model
of the titration of a strong acid with a strong base just described.
The principle of the mathematical linearization of the titration curve can also be demonstrated
using this example.
- + + - CONC • ( V – VEQ )
OH – H = Na – CI =
V0 + V
V0 + V + Kw
G = VEQ – V = H –
CONC H
+
V0 + V +
G 1 = VEQ – V = • H
CONC
holds.
V0 + V Kw
G 2 = V – VEQ = •
CONC H
+
is valid.
Fundamentals of Titration 99
8 Equivalence point determination
When the transformed data are represented graphically, two straight lines are obtained with
slopes of –1 and +1 that intersect the V axis at V = VEQ.
V0 + V -pH
G1 = ¥ 10
CONC -pKw
V0 + V 10
G2 = ¥ -pH
CONC 10
VEQ
This graph, often called a Gran’s plot, demonstrates the potential advantages of this method:
There is no need to titrate to the equivalence point; it can be determined using a few
experimental points either graphically or by calculation using linear regression. However, a
requirement for this is that the electrode parameters (zero point and slope) and the start volume
V0 are known exactly. Otherwise, curves exhibiting partial curvature result and the determina-
tion of VEQ is inaccurate.
It should also be noted that for each titration reaction this method requires both an individual
transformation function G and a knowledge of additional model parameters such as the
stability constants.
The break point of a linear titration curve can be obtained through extrapolation of the
bordering straight parts of the curve and calculation of their intersection points.
The main problem here lies in finding parts of the curve that can be regarded as representative
straight lines. Often only small sections of the titration curve are approximately linear. It should
be noted that all measured values Ei must be corrected for dilution – multiplication of all Ei’s
by the factor (V0 + Vi)/V0. If this measure is omitted, even the linear sections are slightly curved.
The acidity constant Ka (see Section 3.1.4) or the pKa value is a measure of the strength of
an acid in the particular solvent. The pKa value is not only an important quantity in the
classification of an acid, but also determines the properties of the substance in nature or its
possible use as a drug.
The determination of exact values of the pKa by means of titration is a demanding task. The
correct procedure requires not only an exact knowledge of the electrode parameters, but also
the use of activities rather than concentrations.
From the law of mass action of the reaction of an acid HA with H2O (see Section 3.1.4)
+ -
H • A
Ka =
HA
it follows that
-
A
pH = pK a + log
HA
When the acid HA is half neutralized, i.e. V = VEQ/2, the concentration [HA] of the still
undissociated acid is approximately equal to the concentration [A-] of the base.
Hence
pH = pKa
This pH value at half consumption to the equivalence point is called the half neutralization
value. It can easily be shown that this half neutralization value is a good estimation of the pKa
value of weak acids.
The acidity constant Ka can be calculated at any point during the titration with a strong base
if all parameters are known [11]:
+ cT + d
Ka = H •
cA – cT – d
where
d = H
+ - CONC • V CONC • VEQ
– OH cT = cA =
V0 + V V0 + V
The quantity cT is the concentration of the added titrant and cA the concentration of acid in the
titration vessel during the titration. At the half neutralization value (V = VEQ/2)
cT = cHNV = 0.5 cA •
and hence
+ c HNV – d
Ka = H •
c HNV + d
or
c HNV + d
pH HNV = pK a + log
c HNV – d
If the condition d << cHNV is fulfilled, the half neutralization value pHHNV is equal to the pKa.
pH HNV - pK
a
-1
pK a
1 4 7 10 13
As the above figure shows, for pKa values between 4 and 10 the half neutralization value
represents an excellent approximation of the pKa value. For strong acids (pKa < 4) the half
neutralization value gives an overestimate of the pKa value.
[1] M. Kolthoff and N.H. Furmann, “Potentiometric Titrations”, Wiley, New York (1949)
[5] S. Ebel and S. Kalb, Fres. Z. Anal. Chem., 278, 109 (1976)
[7] S. Ebel, E. Glaser, R. Kantelberg and B. Reyer, Fres. Z. Anal, Chem., 312, 604 (1982)
[9] W. Schreiner, M. Kramer, S. Krischer and Y. Langsam, PC TECH JOURNAL, May 1985, p. 170 ff
[11] S. Ebel and W. Parzefall, “Experimentelle Einführung in die Potentiometrie”, Verlag Chemie, Weinheim,
Chapter 3 (1975).
Besides titration, direct measurement of the concentration with a suitable sensor is the most
frequently used method of determination in wet chemical analysis. This is especially true in the
case of water analysis where not only measurement of the pH and the redox potential, but also
the concentration determination with ion selective electrodes and the determination of the
conductivity, turbidity and oxygen content are important.
This section discusses the pH measurement, measurement of the redox potential, direct
measurement with ion selective electrodes and measurement of the conductivity. Particular
attention is paid to the special aspects of sensor calibration.
9.1 pH measurement
The pH is defined as the negative logarithm of the hydrogen ion activity (see Section 3.1.3).
The range of the pH value is given by the autoprotolysis of water and lies between pH 0 and
pH 14. The ionic product
Kw = aH+ aOH- •
and hence also the neutral point ([H+] = [OH-]) depend greatly on the temperature.
0 0.11 10-14
•
7.47
-14
25 1.0 10
•
7.00
50 5.5 10-14
•
6.63
100 54.0 10-14
•
6.13
For thermodynamic reasons it is not possible to calculate or measure the pH exactly. Various
approximation methods and conventions have thus been suggested [1]. The standard
acceptable today for the pH scale comprises different buffer solutions whose pH has been
fixed by convention. The pH values of these so-called NIST (NBS) buffer solutions have also
been accepted by DIN [2].
These buffer solutions are mixtures of substances with a stable hydrogen ion activity, that
show little change on dilution or in the presence of impurities. The pH values of these standard
buffer solutions are tabulated between 0 and 95°C (see appendix A). The buffer solutions
have the following pH values at a reference temperature of 25°C:
For routine calibration many chemical producers and electrode manufacturers offer so-called
technical buffer solutions with predominantly integral pH values. These are less susceptible
to dilution and have greater buffer capacities than the NIST (NBS)/DIN buffers. INGOLD [3]
and MERCK [4] are among the companies that offer such buffer solutions. The temperature
dependencies of the technical buffer solutions of both producers are very similar.
Tables of the temperature dependencies of these technical buffer solutions from INGOLD and
MERCK can be found in appendix A.
The buffer solutions with pH values 4.60 and 7.00 correspond to the electrode assembly zero
points of commercial glass electrodes. For routine calibration of the glass electrodes, buffer
solutions of pH 4, 7 and 10 are normally used.
E = E0 + S pH •
E0 is the standard potential at pH = 0. S is the slope and defines the change in potential per
pH unit. The slope is temperature dependent.
R•T
S = – 2.301
F
Temperature [°C] Slope S
0 –54.20
25 –59.16
50 –64.12
The electrode zero point pH0 (pH value at E = 0 mV) and the slope S are normally specified
as calibration parameters.
E0
+mV
pH0
0 pH
7 14
-mV
The combination glass electrodes used in routine analysis have an electrode zero point at pH
7. New electrodes should have a slope greater than 97% of the theoretical slope predicted by
the Nernst equation.
Both calibration parameters vary slightly from electrode to electrode. For accurate pH
measurements a calibration is thus necessary. The calibration can be performed as a single-
or multipoint measurement.
If the pH is measured over a narrow range, as a rule a single-point calibration suffices. Here,
only the electrode zero point is redetermined, the slope is not checked.
E
old electrode function
+mV
pH0
0 pH
6 7 8 9
Measurement point
-mV
The multipoint calibration utilizes two or more buffer solutions whose pH value must differ by
at least one pH unit. The new values for the calibration parameters pH0 and S are obtained by
linear regression through the measured points.
E
old electrode function
It becomes greater
T2 (measurement)
Error
+mV
pH0
0 pH
7 14
-mV
Error
T 2 + 273.16
S( T 2 ) = S( T 1 ) •
T 1 + 273.16
This correction is only approximately correct since the intersection point of the calibration lines
at different temperatures, the so-called isothermal intersection point Eiso does not lie at
exactly 0 mV [5]. Even though the slope has been corrected for temperature, an error still
arises. E T1 (calibration)
T2 (measurement)
Error
Slope: temperarure compensated
+mV
E iso pH0
0 pH
7 14
-mV
This error increases with increasing temperature difference between calibration and measure-
ment and with increasing value of Eiso. The error is independent of the pH of the solution.
The isothermal intersection point can be determined by calibration with two buffer solutions at
two different temperatures graphically or by calculation.
Temperature compensation which takes into account the isothermal intersection point is then
free from error when the temperature behavior of the slope follows the Nernst equation. This
may not necessarily be the case. For precise measurements a previous calibration at the
The potential of an ion selective electrode depends on the ionic activity in the solution and like
that of the pH electrode is described by the Nernst equation:
E = E0 + S pA
•
pA is the negative logarithm of the ionic concentration of entity A. The negative logarithm of
the ionic concentration of cations is designated pM, that of anions pX.
In this notation the sign of the slope S for cations is negative and that for anions positive.
R•T
S = ± 2.301 •
n F
The theoretical slope of an ion selective electrode at 25°C is 59.16 mV for monovalent ions and
29.58 mV for bivalent ions.
The following criteria must be heeded when working with ion selective electrodes:
Sample pretreatment
Sample pretreatment is the most important factor in direct measurement with ion selective
electrodes. For quantitative analyses each sample solution must be mixed with a certain
amount of electrolyte solution (so-called TISAB solution). These buffer solutions can perform
the following functions:
– Ensure constant ionic strength of all sample and calibration solutions. With this measure-
ment technique the ion selective electrode can be used for direct measurements of the
analytical concentration of interest and not just the ionic activity.
– pH buffering: Depending on the application, the medium must be acidic, neutral or basic.
– Destruction of all complexes with the analyte ion to determine the total ionic concentration.
Selectivity
The selectivity of an ion selective electrode is limited and is expressed by the selectivity
coefficient (see Section 4.1.5). Disturbing influences due to foreign ions can be suppressed
by appropriate sample pretreatment.
Precision
In the ideal case, the attainable precision for univalent ions is 1 to 2%, that for bivalent ions 2
to 4%.
Limit of detection
The lower limit of detection is determined by the ions released by the sensing electrode itself.
This minimum concentration depends on the solubility of the active membrane substance in
the medium. In the vicinity of the limit of detection the potential deviates from the linear behavior
predicted by the Nernst equation.
E [mV]
Limit of detection
Validity of the
Nernst equation
pM
1 3 5 7
Calibration
For the performance of a direct measurement the electrode must be calibrated with solutions
of known concentration. In the linear region two calibration points suffice, but in the nonlinear
The determination of the redox potential with a redox electrode is also a potentiometric
measurement. The redox potential is a measure of how easily a substance can accept or
donate electrons.
The measurable redox potential follows the Nernst equation. The correct equation is obtained
from the oxidation or reduction process.
3+
R•T Fe
Examples: E = E 0 + 2.301 log
F 2+
Fe
a. Fe2+ —> Fe3+ + e-
8
- +
R•T MnO 4 H
E = E 0 + 2.301 • log
5 F 2+
Mn
b. Mn2+ + 4H2O —> MnO4- + 8H+ + 5e-
These examples show that the redox potential is always determined by the ratio of the oxidizing
and reducing components and can also be pH dependent. The oxidizing or reducing action of
an analysis solution can thus also be a function of the pH.
Measurement of the redox potential is not carried out all that often for the following reasons:
In water analysis the redox potential is a frequently determined measured value (e.g. analysis
of the water quality of swimming pools).
The use of redox buffer solutions (e.g. INGOLD redox buffer solutions Nos. 9881 and 9883 [6])
is thus restricted to a purely operational test of the redox electrode.
The determination of the conductivity [7], [8], [9] as a measure of how well a solution conducts
an electric current has already been discussed in section 4.3.1. In addition to its use as an indi-
cation method for titrations, direct determination of the conductivity has achieved an important
standing. The most important application areas of the conductivity measurement are:
χ (T)
∆χ
χR
∆T
T
T T
R
For temperature ranges within ±10°C of the value of the reference temperature, the tempera-
ture coefficient is virtually constant. The measured conductivity is corrected to the reference
temperature by means of the following equation:
G( T ) • Z
χR =
α
1+ ( T – TR )
100
χ
where
R
: conductivity at the reference temperature TR [µS/cm, mS/cm]
G (T) : measured conductance at temperature T [µS, mS]
Z : cell constant [cm-1]
α : temperature coefficient [%/°C]
TR : reference temperature [°C]
T : measurement temperature [°C].
The temperature coefficient α has a typical value of 0 – 4%/°C. A mean value of 2%/°C is
frequently assumed.
For highly precise work the exact temperature dependence must be determined with the aid
of a reference solution.
The exact determination and checking of the cell constants is effected using calibration
solutions. The standards employed are KCl solutions of concentration 0.01, 0.1 and 1 mol/L.
[1] R.G. Bates, “Determination of pH”, John Wiley (1973)
[5] "Practice and Theory of pH Measurement", INGOLD brochure E-pH-TH-2-CH, Section 4.7, 1989
[8] F. Oehme and R. Bänninger, “ABC der Konduktometrie”, offprint, Chemische Rundschau (1979)
The aim of a titration is normally the determination of the content of a substance in a sample.
The result of the analysis is used to assess the test sample.
In practice, each analytical result is associated with random and systematic errors. Every
analyst must therefore always ask himself whether the quality of his titration results is good
enough. The methods offered by statistics are an excellent means to assess the accuracy of
the results ([1], [2],[3]).
For the practical application of statistics in the context of titration, a brief description of the most
important concepts and a listing of the associated calculation formulae will suffice here.
Arithmetic mean x
The arithmetic mean x is equal to the sum of the independent measurement results xi of a
series of measurements divided by the number of measurements N.
N
∑
1
x = xi
N i=1
Variance s2
The variance s2 of a series of measurements is the sum of the squares of the deviations of the
N individual values xi from the arithmetic mean x divided by the number of degrees of freedom
f (f = N-1).
N
1
∑ (x i – x )
2 2
s =
N–1 i=1
Standard deviation s
The standard deviation s of a series of individual values xi is a measure of the spread of the
individual values xi about the mean x. It is given by the positive square root of the variance s2.
The standard deviation has the unit of the individual values and should always be specified with
one significant figure more than the mean.
N
∑ (x i – x )
1 2
s =
N–1 i=1
s = x1 – x2 2
s
s rel =
x
s
s rel = • 100 %
x
Example: For the calculation of the standard deviation s of N independent individual values,
the number of degrees of freedom is N - 1 since the arithmetic mean x also enters
the calculation.
Confidence level P
The confidence level P denotes the probability that a certain statement is correct. In routine
analysis a confidence level of 95% is normally used, whereas scientific investigations employ
a level of 99%.
The value µ is the true mean value that identifies the position of the distribution. The scatter
σ of the population is a measure of the width of the distribution.
N
∑ xi
1
µ = lim
N
n→∞ i=1
N
∑
1 2
σ = lim ( xi – x )
N
n→∞ i=1
h (z)
2
- 1 z
h (z) = k • e 2
x–µ
z =
s
z
x: µ – 1.96 s µ– s µ µ+ s µ + 1.96 s
Within the limits of µ ± σ, 68.3% of all measured values are expected (hatched area), whereas
95% of all measured values lie between µ - 1.96 σ and µ + 1.96 σ.
The spread T signifies that P% of all measured values are to be expected in the range x + T
and x - T. For P = 95%, the spread is given by
T = ± 1.96 σ
or in general
T = ±µ σ •
The factor µ is known as the fractile of the normal distribution and depends only on the
selected confidence level P.
If a small number of measured values is present (typical in analytical practice), this is called
a random sample. If µ and σ of the population are unknown, the distribution of the random
sample values does not yet follow the normal distribution. The mean x is the best estimate of
the true value but is still associated with a statistical uncertainty that needs to be taken into
account. The distribution of the random sample will thus be broader than the normal
distribution. The theoretical function of this distribution is called the t or student distribution.
The spread T of the individual measurement and the confidence interval CI of the mean
value can thus be calculated.
T = ±s t •
s•t
CI = ±
N
The confidence interval CI (= spread of the mean) signifies that the true mean lies within the
range x + CI and x - CI with P% certainty.
The factor t is the fractile of the t distribution and depends on the confidence level P and the
number of measured values N; it can be taken from the t table (see appendix B).
Such outliers are uncovered by means of statistical tests, e.g. with the Grubbs test. Here, the
mean x and the standard deviation s of the analysis data are first calculated. From the
experimental data the value x* with the greatest deviation from the mean is sought and tested
using the following condition equation:
x* – x
TV =
s
The test variable TV is compared with the value in the Grubbs table G (N, P%), which in turn
depends on the number of analysis values N. A Grubbs table can be found in appendix B.
If the test variable TV is greater than G (N, P%), the experimental value under test is considered
an outlier and deleted from the series of measurements. The remaining data of the series are
used to calculate new values of the mean, and the standard deviation and the outlier test
repeated for another value suspected of being an outlier.
x1 = 30.38 %
x2 = 30.23 %
x3 = 30.34 %
x4 = 29.98 %
x5 = 30.29 %
x6 = 30.31 %
–––––––––––
x = 30.26 %
s = 0.144 %
outlier suspect: x4
x4 – x
TV = = 1.944
s
G(6,90%) = 1.822 —> x4 is an outlier.
The deviation from the correct value or that value accepted as correct of an analytical result
is known as the error.
Gross errors
Gross errors arise through failure to follow analytical procedures, through faulty analytical
instruments and through carelessness on the part of the analyst. From the statistics point of
view, gross errors are always avoidable.
Random errors
Random errors are as a rule unavoidable and are responsible for the spread in the individual
measured values about the mean. The magnitude of the random errors – determined using
the standard deviation s – is a measure of the precision of an analytical procedure.
Systematic errors
Systematic errors give rise to a difference between the expected value and the mean x
obtained from the measurement and thus determine the accuracy of an analytical procedure.
Systematic deviations have a definite cause and in principle can be rectified. A distinction is
made between proportional systematic and constant systematic deviations. It is apparent from
the standard formula for the calculation of titration results
that, for example, a wrong titrant concentration will lead to a proportional systematic error and
a wrong solvent blank value to a constant systematic error.
Errors
Deviations from the correct or assumed correct value
Correctness
The correctness of an analysis method is a qualitative concept. It describes the systematic and
random deviations of the measurement results from the true value. The accuracy (determined
by systematic errors) and the precision (determined by random errors) are subdivisions of the
concept of correctness.
A titration comprises a pretreatment part and the analytical process with the titrator. In the first
part the analyst has a great influence on the error by way of his method of working. In the actual
measurement process he has no influence on the correctness of the results. Even assuming
correct maintenance by the user, the quality of the analytical result of the titration is still greatly
dependent on the quality of the instrument. This is the responsibility of the manufacturer.
The limit of detection and the limit of determination are two concepts that are constantly
confused.
The limit of detection describes the smallest amount of substance or lowest concentration that
can be distinguished qualitatively from zero amount of substance with a specified confidence
level (e.g. 95%) in a single analysis.
The limit of determination is the smallest amount of substance or lowest concentration that can
be determined quantitatively with a specified confidence level (e.g. 95%) and can be
distinguished from zero amount of substance. The limit of determination provides no
information regarding the correctness of the analytical procedure.
A standard is a chemical substance that is used as a reference sample. It is used to check the
accuracy of an analytical method.
Standard samples are samples containing a component whose concentration is known with
sufficient accuracy and which can be used as standards.
Control samples are analysis samples whose matrix composition corresponds very closely to
that of real samples. They are used to check the accuracy and the precision of an analysis
method.
– Regular determination of the titer of the titrant used (also applies when commercial
volumetric solutions are used)
– Determination of any blank value of the solvent or the matrix
– Regular use of control samples.
For the standardization (titer determination) and control of titrimetric solutions, substances are
needed that within the limits of measurement accuracy of the titration (lower limit: 0.01%)
correspond exactly to the composition given by their formula. These so-called primary
standards must have the following properties:
In routine analysis (use of the same analysis procedure for many samples), the regular inser-
tion of control samples of known composition (e.g. primary standards) between the individual
series has proved useful. This provides a check not only on the titer but also of the correct
functioning of the analytical instrument. In addition to the result of the titration with the control
sample, its progress (titration time, titration curve, etc.) provides indications of any problems
(contamination of the electrode, etc.).
Note: Tables of primary standards for the most important titrants can be found in
appendix C.
[1] J.C. Miller and J.N. Miller, “Statistics for Analytical Chemistry”, second edition, Ellis Horwood, Chichester,
1988.
[2] R. Caulcutt and R. Boddy, “Statistics for Analytical Chemists”, Chapman and Hall, London, 1983.
[3] W. Funk, V. Damman, C. Vonderheid und G. Oehlmann, “Statistische Methoden der Wasseranalytik”, VCH
Verlagsgesellschaft, Weinheim, 1985
Appendices
Appendix C: Tables of primary standards for the most important titrants (see section 10.5)
Temper- Potas- Potas- Potas- Potas- Phos- Phos- Borax Sodium Calcium
ature sium sium sium sium phate phate carbo- hydrox-
°C tetra- hydrogen dihy- hydrogen nate/ ide
oxalate tartrate drogen phthalate Sodium
citrate bicar-
bonate
2. MERCK buffers
2. MERCK buffers
Temper- Phosphate Borate/ Boric acid/ Borate Boric acid/ Boric acid/ Phos- Potassium
ature hydro- potassium potassium potassium phate/ chloride/
°C chloric cloride- chloride- chloride- sodium sodium
acid sodium sodium sodium hydroxide hydroxide
hydroxide hydroxide hydroxide
3. INGOLD buffers
1. t table
ƒ P = 95 % P = 99 % P = 99.9 %
Alkalimetry
Sodium hydroxide NaOH deion. H2O 2, 5 DG111–SC Potassium hydrogen
Fundamentals of Titration
phthalate
Potassium hydroxide KOH C2H5OH or CH3OH or
(CH3)3COH or
i-C3H7OH 2, 5 DG112–SC/ Benzoic acid
DG113–SC
TBAH C16H37NO CH3OH/i-C3H7OH 2, 5 DG112–SC/ Benzoic acid
DG113–SC
Sodium methylate NaOCH3 CH3OH 1, 5 DG112–SC/ Benzoic acid
DG113–SC
Acidimetry
Hydrochloric acid HCl deion. H2O 3 DG111–SC Tris(hydroxymethyl)-
aminomethane
C2H5OH or C3H7OH 2 DG112–SC/ Tris(hydroxymethyl)-
DG113–SC aminomethane
Sulfuric acid H2SO4 deion. H2O 3 DG111–SC Tris(hydroxymethyl)-
aminomethane
Nitric acid HNO3 deion. H2O 3 DG111–SC Tris(hydroxymethyl)-
aminomethane
Perchloric acid HClO4 Glacial acetic acid 2 DG112–SC/ Tris(hydroxymethyl)-
DG113–SC aminomethane
137
138
Redox – Titrations
Reducing agents
Ammonium ferrous (ll)
sulfate (NH4)2Fe(SO4)2 60 mL 50% H2SO4
+ deion. H2O → 1 L 1, 4 DM140–SC Potassium dichromate
Sodium thiosulfate Na2S2O3 deion. H2O (+ 3 drops
CHCl3 + 0.1 g Na2CO3) 3 DM140–SC Potassium iodate
Hydroquinone C6H6O2 20 mL conc. H2SO4
+ deion. H2O → 1 L 2, 7 DM140–SC Potassium dichromate
Oxidizing agents
Potassium
permanganate KMnO4 deion. H2O 3, 7 DM140–SC Di-Sodium oxalate
Iodine I2 2.5% KI / deion. H2O 1, 7, 8, 9 DM140–SC Di-Sodium oxalate
Cerium (lV) sulfate Ce(SO4)2 58 mL 50% H2SO4
+ deion. H2O → 1 L 3 DM140–SC Di-Sodium oxalate
Potassium dichromate K2Cr2O7 deion. H2O 3 DM140–SC Di-Sodium oxalate
Iron (III) chloride FeCl3 deion. H2O 3 DM140–SC Ascorbic acid
Sodium nitrite NaNO2 deion. H2O 2 DM140–SC Sulfanilic acid
2,6 Dichlorophenol-
indophenol-Na-salt DPI deion. H2O 1, 7, 8, 9 DM140–SC/ Ascorbic acid
DP550
Fundamentals of Titration
4: Protect from O2
Precipitation Titrations
Argentometry
Silver nitrate AgNO3 deion. H2O 3, 7 DM141–SC Sodium chloride
Fundamentals of Titration
Sodium chloride NaCl deion. H2O 3 DM141–SC Silver nitrate
Potassium bromide KBr deion. H2O 3 DM141–SC Silver nitrate
Sulfate/Fluoride
Barium perchlorate Ba(ClO4)2 i-C3H7OH/H2O 2 DP550/
with Thorin Na2SO4 solution
Barium chloride BaCl2 deion. H2O 2 DP550/
with Thorin Na2SO4 solution
Lanthanum nitrate La(NO3)3 deion. H2O 3 Fluorid-Sensor NaF solution
Lead nitrate Pb(NO3)2 deion. H2O 3 Fluorid-Sensor NaF solution
139
140
Complexometric Titrations
Turbidimetric Titrations
N-cetylpyridinium-
chloride CPC deion. H2O 3 DP550/660 Sodium dodecyl
sulfate
Sodium dodecyl
sulphate SDS deion. H2O 3 DP550/660 N-cetylpyridinium
chloride
Water Titrations
Fundamentals of Titration
Index
Index
Titration curves 59
Titration reaction 12
Titrimetric analysis 5, 6
Titrimetric solution 7
Transmission 44
t table 123, 135
Tubbs procedure 93
Types of titration 50
Variance s2 119
Visual indication 27
Voltametry 42
*P704153*
© Mettler-Toledo GmbH 1993, 1996, 1997, 1998 ME-704153A Printed in Switzerland 9806/9.12
Mettler-Toledo GmbH, Analytical, Sonnenbergstrasse 74, CH-8603 Schwerzenbach, Tel. (01) 806 77 11, Fax (01) 806 73 50, Internet:http://www.mt.com
AT Mettler-Toledo Ges.m.b.H., A-1100 Wien AU Mettler-Toledo Ltd., Port Melbourne, Victoria 3207
Tel. (01) 604 19 80, Fax (01) 604 28 80 Tel. (03) 9 646 45 51, Fax (03) 9 645 39 35
BE n.v. Mettler-Toledo s.a., B-1651 Lot CH Mettler-Toledo (Schweiz) AG, CH-8606 Greifensee
Tel. (02) 334 02 11, Fax (02) 378 16 65 Tel. (01) 944 45 45, Fax (01) 944 45 10
CN Mettler-Toledo (Shanghai) Ltd., Shanghai 200233 CZ Mettler-Toledo, spol, s.r.o., CZ-12000 Praha 2
Tel. (21) 6485 04 35, Fax (21) 6485 33 51 Tel. (02) 25 15 55, Fax (02) 242 475 83
DE Mettler-Toledo GmbH, D-35353 Giessen DK Mettler-Toledo A/S, DK-2600 Glostrup
Tel. (0641) 50 70, Fax (0641) 507 128 Tel. (43) 270 800, Fax (43) 270 828
ES Mettler-Toledo S.A.E., E-08038 Barcelona FR Mettler-Toledo s.a., F-78222 Viroflay
Tel. (03) 223 22 22, Fax (03) 223 02 71 Tel. (01) 309 717 17, Fax (01) 309 716 16
HK Mettler-Toledo (HK) Ltd., Kowloon HR Mettler-Toledo, d.o.o., HR-10000 Zagreb
Tel. (02) 744 12 21, Fax (02) 744 68 78 Tel. (01)230 41 47, Fax (01) 233 63 17
HU Mettler-Toledo, KFT, H-1173 Budapest IT Mettler-Toledo S.p.A., I-20026 Novate Milanese
Tel. (01) 257 98 89, Fax (01) 257 70 30 Tel. (02) 333 321, Fax (02) 356 29 73
JP Mettler-Toledo K.K., Yokohama 231 KR Mettler-Toledo (Korea) Ltd., Seoul (135-090)
Tel. (45) 633 53 50, Fax (45) 664 96 50 Tel. (02) 518 20 04, Fax (02) 518 08 13
MY Mettler-Toledo (M), Sdn. Bhd. 47301 Petaling Jaya MY Mettler-Toledo (S.E.A.), 47301 Petaling Jaya
Tel. (03) 703 27 73, Fax (03) 703 17 72 Tel. (03) 704 17 73, Fax (03) 703 17 72
MX Mettler-Toledo S.A. de C.V., México C.P. 06430 NL Mettler-Toledo B.V., NL-4000 HA Tiel
Tel. (05) 547 57 00, Fax (05) 541 22 28 Tel. (0344) 638 363, Fax (0344) 638 390
PL Mettler-Toledo, Sp. z o.o., PL-02-929 Warszawa RU Mettler-Toledo C.I.S. AG, 10 1000 Moskau
Tel. (22) 651 92 32, Fax (22) 651 71 72 Tel. (95) 921 92 11, Fax (95) 921 63 53
SE Mettler-Toledo AB, S-12008 Stockholm SG Mettler-Toledo (S) Pte. Ltd., Singapore 139944
Tel. (08) 702 50 00, Fax (08) 642 45 62 Tel. 7 786 779, Fax 07 764 904
SK Mettler-Toledo, service, s.r.o., SK-83103 Bratislava SL Mettler-Toledo, d.o.o., SL-61111 Ljubljana
Tel. (07) 525 21 70, Fax (07) 525 21 73 Tel. (06) 112 35 764, Fax (06) 127 45 75
TH Mettler-Toledo (Thailand), Bangkok 10320 TW Mettler-Toledo Pac Rim AG, Taipei
Tel. (02) 719 64 80, Fax (02) 719 64 79 Tel. (62) 579 59 55, Fax (62) 579 59 77
UK Mettler-Toledo Ltd., Leicester, LE4 1AW US Mettler-Toledo Inc., Worthington, OH 43085
Tel. (0116) 235 70 70, Fax (0116) 236 63 99 Tel. (614) 438-4511, Fax (614) 438-4900