You are on page 1of 9

TITRIMETRY | Potentiometry☆

A Hulanicki, M Maj-Żurawska, and S Glab, University of Warsaw, Warsaw, Poland


ã 2013 Elsevier Inc. All rights reserved.

Introduction 1
Potentiometric Equipment 1
Indicator Electrodes 1
pH-sensitive electrodes for acid-base titrations 1
Electrodes for precipitation titrations 2
Electrodes for compleximetric titrations 2
Electrodes for redox titrations 2
Reference Electrodes 2
Instrumentation 2
Potentiometric Titration Curves 2
Endpoint Evaluation 3
Single Point Titration 4
Potentiometric Titrations with Polarized Indicator Electrodes 5
Automated Potentiometric Titrations 5
Applications 6
Acid-Base Titrations 6
Precipitation Titrations 6
Compleximetric Titrations 7
Redox Titrations 7
Potentiometric Coulometric Titrations 8

Introduction

Potentiometric titration belongs to chemical methods of analysis in which the endpoint of the titration is monitored with an
indicator electrode that records the change of the potential as a function of the amount (usually the volume) of the added titrant
of exactly known concentration. Potentiometric titrations are especially versatile because indicator electrodes suitable for the
study of almost every chemical reaction used in titrimetry are now available. This technique is also frequently used in the study
of operational conditions of visual titrimetric indicators proposed for general use in chemical analysis, as well as in the study of
numerous reactions, such as protonation and complexation, which find their application not particularly in analytical
measurements. The course of the potentiometric titration curve provides information not only about the titration end
point position, but also the position and shape of the curve may provide data about the processes accompanying the titration
reaction. Another advantage is that the necessary apparatus is generally not expensive, reliable and readily available in the
laboratories.

Potentiometric Equipment

Applications of potentiometry involve the use of an electrochemical cell consisting of a reference electrode of constant potential
and an indicator electrode that responds to the analyte studied and sample composition. The electromotive force (e.m.f.) of this cell
can be regarded as the difference of the potentials of the two electrodes (half-cells).

Indicator Electrodes
pH-sensitive electrodes for acid-base titrations
The glass electrode is most often used but occasionally metal–metal oxide (e.g., antimony electrode), and pH-sensitive polymeric
membrane electrodes are also used. These electrodes are also useful in investigation of chemical reactions through potentiometric
titrations, as in many of them their course can be traced by the changes of hydrogen ion activity (concentration).


Change History: March 2013. A Hulanicki added ‘Potentiometric Coulometric Titrations’ section; added new references in ‘Further reading’ section.

Reference Module in Chemistry, Molecular Sciences and Chemical Engineering http://dx.doi.org/10.1016/B978-0-12-409547-2.00546-1 1


2 TITRIMETRY | Potentiometry

Electrodes for precipitation titrations


The choice of the electrode depends on the character and properties of ions taking part in the titration reaction. Common is the use
of silver or mercury electrodes of the first order to trace the respective metal ions, or as the second-order electrodes: silver–silver
halide and mercury-mercury(I) halide electrodes, when the corresponding anions are involved in the reaction. Metal electrodes
with sulfide (or other sulfur(II) containing species) as co-anion are used less often. There is also a large group of ion-selective
electrodes (ISEs) sensitive to F, Cl, Br, I, S2, Cu2þ, Pb2þ, and Kþ, which are used to monitor precipitation titrations of both
inorganic and organic ions.

Electrodes for compleximetric titrations


Electrodes respond mainly to changes of metal ion concentration, similarly as electrodes used in precipitation reactions. Occa-
sionally, the electrode responds to the ligand, as is the case of the fluoride ISE. Electrodes with a similar scheme as the electrode of
third order (Ag/Ag2C2O4)/(CaC2O4/Ca2þ) can be used in compleximetric titration having as an intermediate step the equilibrium
with the appropriate ligand (e.g., AgEDTA/MEDTA), where M denotes the metal ion being titrated. There is additionally a large
group of ISEs that respond to the metal ions. As complexation with protonated ligands is accompanied by pH changes, pH-sensitive
electrodes are occasionally used.

Electrodes for redox titrations


For direct measurements of changes of the oxidation-reduction potentials, inert electrodes, such as bright platinum or gold, are
mainly used. Iridium and tungsten electrodes can also be used for measuring the redox potentials, the kinetic effects however may
influence their response. The practical application of mercury as an inert electrode is limited to regions of negative redox potentials,
but in this region it has a definite advantage over the noble metal electrodes. This is connected with the high hydrogen over-
potential on mercury.

Reference Electrodes
Most commonly used reference electrodes are based on mercury (calomel electrode) and silver, in equilibrium with the saturated
solution of the corresponding chlorides. Their potential will be constant if the chloride ion concentration around the electrode
is constant. The mercury-mercury(I) sulfate electrode is used instead of the calomel electrode when the presence of chloride is
undesirable.

Instrumentation
The experimental instrumentation for potentiometric titration consists primarily of a high-impedance electronic pH or millivolt
meter; a beaker and a magnetic stirrer, indicator and reference electrodes, and a burette for titrant delivery, which are needed for
manual titrations and point-by-point plotting. Automatic burettes and direct curve recording devices are increasingly used, in
particular when the decrease of the scale of the titration is expected to go down below 1 ml of the amount of titrant used up to the
end point. A variety of instruments are commercially available. In general pH (or potential) meters can be divided on the basis of
price and performance: utility (portability), general-purpose, expanded scale and research grades.

Potentiometric Titration Curves

Potentiometric titration curves are obtained by plotting the potential of the indicator electrode (in practice the e.m.f. of the cell)
against the volume of the titrant. In acid-base titrations, the pH value is usually plotted instead of potential. The logarithmic shape
of the titration curve is a consequence of the logarithmic dependence of activity or activity ratio of the species participating in the
titration reaction according to the Nernst equation:
RT aox
E ¼ E0 þ ln
zF ared
where aox represents the oxidant activity and ared the reductant activity. When the activity of one of those species is constant, the
activity ratio is simply replaced by the activity of the ion sensed by the indicator.
From the shape and position of the titration curve along the potential axis, information can be gained about the parameters of
the titration reaction. For example, in alkalimetric titration of an acid the pH value of the half-neutralization of the acid
corresponds numerically to the pKa value of the acid. This makes possible evaluation of the pKa value of the titrated species.
Similarly, in other types of titrations (precipitation, compleximetric, and redox) the titration curve may supply information about
the relevant titration reaction.
In the case of an ideal indicator electrode, such a curve should correspond to the change of the indicator ion activity. In practice,
the curve is often distorted because of changes of liquid junction potential and insufficient selectivity of the electrode (especially for
ISEs) and occurrence of processes not described by the Nernst equation. Nevertheless, in most cases it does not influence the
TITRIMETRY | Potentiometry 3

-100

-150

-200
Cu2+

E (mV)
-250 -250

E (mV)
-300 -300
Ni2+

-350 -350
0 1 2 3 4 5 6 7 8 0 1 2 3 4 5 6
(a) Volume (ml) (b) Volume (ml)

Figure 1 (a) Symmetric titration curve: compleximetric titration of copper ion, using a copper-selective electrode; (b) asymmetric titration curve:
compleximetric titration of nickel using a copper-selective electrode.

evaluation of the endpoint. A potentiometric titration curve should have a symmetrical ‘S’ shape, when the dilution effect is
neglected, and the indicator electrode is directly sensitive to the ion titrated. If the course is traced indirectly, for example, in a
compleximetric titration of nickel using a copper-selective electrode, the curve is asymmetric (Figure 1). Another reason for the
asymmetry of the curve is the asymmetry of the titration reaction, which mostly occurs in redox titrations when the number of
electrons in the reaction of the oxidant (e.g., MnO4 ! Mn2þ) is not equal to the number of electrons exchanged by the reductant
(e.g., Fe2þ ! Fe3þ). Unusual shapes of the titration curve may occur when the titrated substance is directly inactive but after the
addition of the titrant is gradually transformed into an active form in a kinetically restricted reaction. Such situations may occur also
when catalytic processes occur. An improved and rapid potential change is observed when an indicator reaction is catalyzed by the
small excess of the titrant. This can be exemplified by silver determination using iodide as titrant. The indicator reaction: Ce(IV) þ
As(III) ¼ Ce(III) þ As(V), followed by a redox electrode, is catalyzed by the presence of free iodide ions, which appear just after the
completing the reaction between silver and iodide.

Endpoint Evaluation
The endpoint of a titration is an experimentally estimated value, whereas the equivalence point corresponds to the theoretical
amount of the titrant that is exactly equivalent to the amount of the analyte. The difference between the endpoint and the
equivalence point should be minimized and is referred as the titration error.
The endpoint of the titration is often evaluated on the basis of a graphically represented titration curve. When the potential (or
pH) change is large and the curve is steep, the endpoint can be easily seen and evaluated. For less obvious cases of symmetrical
titrations, several graphical methods of finding the point of maximal slope were used. Another method is titration to a fixed
potential. The equivalence point concentration of the indicator ion may be calculated from the known reaction equilibrium
constants, and the corresponding potential values are subsequently evaluated. The equivalence point potential may also be
estimated empirically. Titration is continued until this potential is reached, and the volume of the titrant noted. The maximum
slope of the titration curve is easily found by plotting the first derivative curve, dE/dV, or the second derivative curve, d2E/dV2,
against the titrant volume, V (Figure 2). In practice this can be done for finite increments and the curve is approximated by plotting
(En  Enþ1)/(Vn  Vnþ1) against 0.5(Vn  Vnþ1), where n represents the successive additions of titrant. The numerical procedure

Figure 2 (a) Titration curve plotted in coordinates: potential, E, versus titrant volume, V; (b) first derivative curve, dE/dV, versus V; (c) second
derivative curve, d2E/dV 2, versus V.
4 TITRIMETRY | Potentiometry

termed the Hahn method is based on the same principle. Computerized titrations can achieve the same goal without calculation by
the analyst. It must be pointed out that such procedures are exactly valid for symmetrical curves only.
The common procedure of finding the endpoint, which should correspond to the equivalence volume, is based on the Gran
method. This procedure is especially useful for flat and unsymmetrical titration curves. The titration data are transformed in such a
way that the titration curve consists of two linear parts – before and after the equivalence point. The intersection of those lines, or
their intersection with the abscissa defines the equivalence volume of the titrant (Figure 3). The linearization is based on the
transformation of the Nernst equation into the form

antilog ðE=SÞ ¼ constant aX


0
where the constant is antilog (S /E), and S ¼ RT/2.303 F. The straight line is obtained when the increase of volume (V0 þ V) is taken
into account. Any deviation from linearity may indicate the occurrence of another reaction, either preceding or following the main
reaction.

Single Point Titration


This procedure is based on two potential measurements, one directly using the sample solution (E1), and another on the sample
solution with addition of a known amount of analyte ion (E2). This is also termed the known (or standard) addition method. For
successful application of this method certain conditions are required.

1. The activity coefficient of the indicator ion should not vary significantly between the two measurements.
2. The side reaction coefficient of the indicator ion, that is, the ratio of the free determinand to all its chemical forms, should be
constant.
3. The calibration slope of the electrode (S ¼ RT/F) must be known and constant. When these conditions are fulfilled,

E ¼ E2  E1 ¼ S logðVS CX þ VA CA Þ½CX ðVS þ VA Þ1

where VS and VA are volumes of sample and of the addition, respectively, CX and CA are unknown and addition concentrations,
respectively. This enables Cx to be calculated as
 
CA VA VA þ VS 1
CX ¼ 10E=S 
VA þ VS VA

which for negligible volume of addition (VA >>VS) simplifies to


 1
CX ¼ CA 10E=S  1

A reverse method is based on the removal of the determinand and is called the known (or standard) subtraction method. The
removal of the determinand is based on its precipitation or complexation. The necessary condition is that the reaction must be as
complete as possible.
Potentiometric titrations may be also performed in the flow injection mode. This is attractive because small amount of sample
are needed, the time of determination is short, the results for low concentration are reliable. Besides, the potentiometric signal
depends linearly on the concentration of the analyte, enabling determination in a broad concentration range. There are several
types of such titration, for example, a fixed sample volume is mixed with a titrant solution of constant concentration. The width of
the potentiometric signal, measured on the time scale as Dt, is proportional to the logarithm of the analyte concentration. The
quantitative evaluation is based on a calibration graph. The important factors are the linear range of determination and the time
constant of the potentiometric detector. The latter must be compatible to the flow rate of the solution in the system.

Figure 3 Titration of two component mixture; (a) titration curve; (b) linearization according to the Gran method.
TITRIMETRY | Potentiometry 5

Potentiometric Titrations with Polarized Indicator Electrodes


A number of oxidation-reduction systems (e.g., as Cr2O72/Cr3þ) are slow in establishing constant potentials at the platinum
indicator electrode when measurements are made without current flow. It is possible to avoid long waiting times by forcing slight
electrolysis to occur at the indicator electrode. Potentiometric titrations with polarized electrodes are divided in two classes,
depending on whether one or two electrodes are used. In the former, the potential of a single polarized platinum indicator
electrode against a reference electrode is measured. This indicator electrode may be polarized anodically or cathodically. If both
couples involved in titration are reversible, the potential change of one polarized indicator electrode at the endpoint will be the
same as when the electrode is not polarized. When one or both of the couples are irreversible, the potential change may be much
greater in the case of a polarized electrode.
In the case of potentiometry with two polarized electrodes (bipotentiometry, or differential electrolytic potentiometry), the
difference of potential between the two electrodes is measured. The shape of the titration curve depends on the reversibility of
the couples involved in the titration. If both the titrant and the analyte are reversible couples (e.g., Fe3þ þ Ce4þ in H2SO4), a curve
with a maximum at the endpoint is obtained (Figure 4(a)). When the analyte forms an irreversible couple and is titrated with a
reversible titrant couple (e.g., S2O32 þ I2) the potential decreases at the end point (Figure 4(b)). For the titration of a reversible
analyte with an irreversible titrant another curve is obtained (Figure 4(c)).

Automated Potentiometric Titrations

The most important component of an automated potentiometric titrator is the potentiometer. In simpler titrimeters, ordinary
burettes are used to deliver the titrant and are equipped with automatic valves. Usually piston (syringe) burettes allow delivery of
the titrant with predetermined constant or variable speed are employed. Most automatic titrators have built-in electronic switches
to perform various duties and include one or more derivatizing circuit enabling the location of the endpoint. The automated
potentiometric titrators available on the market may be divided into four groups: (1) automatic titration curve recording
instruments; (2) preset endpoint titrators, in which the flow of the titrant is stopped when the potential of the indicator electrode
reaches a preselected value; (3) first- or second-derivative titrators that continue titration until dE/dV ¼ max or d2E/dV2 ¼ 0;
(4) automatic titrimetric analysers, highly automated and suitable for routine analysis. These last instruments take a sample,
dispense the solution, refill the burette, and rinse the cell and electrodes automatically. They also calculate the result and produce a
printout or a graph.
Some automatic potentiometric titrators (continuous titrators) carry out the analysis continuously. The flow rate of the titrant
varies according to the signal obtained from the electrode. Once a steady state has been attained, the flow rate of the titrant becomes
the measure of the concentration of the analyte in the sample. If the potential of the indicator electrode and the flow rate of the
titrant are recorded as function of time one obtains the record of the variation of analyte concentration with time.

(a)
DE

(b)

(c)
0 1
Titration fraction
Figure 4 Titration curves for potentiometric titration using two polarized electrodes: (a) analyte and titrant reversible systems; (b) irreversible analyte,
reversible titrant; (c) reversible analyte, irreversible titrant.
6 TITRIMETRY | Potentiometry

Most automatic titrators can be utilized as pH-stats or potentiostats, with the purpose of maintaining the pH or redox potential
of a solution at a given value.
Automatic potentiometric titrations are widely used nowadays in industrial and research laboratories. Most of the procedures
used in normal titration need only slight alterations to make them suitable for automatic work.
Potentiometric endpoint detection is frequently used in automatic constant current coulometric titrators (coulometric
titrations).

Applications

Many different types of indicator electrodes can be used as endpoint indicators in potentiometric titrations. For example, an acid-
base titration can be performed with a glass electrode as endpoint detector instead of using colored indicators, or chloride ions can
be titrated with silver(I) using a chloride-ion- or silver-ion-selective electrode.

Acid-Base Titrations
Acid-base potentiometric titrimetry will always play an important role in quantitative analysis because of the extremely large
number of inorganic, organic and biochemical compounds that possess acidic or basic properties. Potentiometric titrations of
strong acids and bases with suitable strong titrants are relatively simple, since there is a relatively large and abrupt change in pH at
the endpoint. As the acid (or base) titrated becomes weaker, the sharpness and magnitude of the endpoint decrease. A similar
situation occurs when the concentration of the substance titrated (and the titrant) is decreased. If 0.1% accuracy is needed in
titration of strong acid or base, the concentration of the substance titrated should be greater than  3  104 mol l1, but if 1%
accuracy is sufficient, the concentration can be decreased by one order of magnitude. For the titration of a weak acid with a strong
base, the product of the acid concentration and its dissociation constant, Ka, should be greater than  107 for 0.1% accuracy, or
109 for 1% accuracy.
When the reaction equilibrium constants differ significantly (e.g., DpK >4) for comparable concentrations of two species,
determination of both is possible from one titration run. As the reaction constants depend on the solvent properties, the change of
solvent may be advantageous in subsequent titration of two analytes.
The range of applications of potentiometric titrations for determination of acids and bases is very wide, as illustrated by the
following examples. Carbonate, hydrogencarbonate and hydroxide ions are all bases that can be titrated with a strong acid such as
hydrochloric acid. The most popular method for determination of nitrogen, which is found in many important substances such as
proteins, fertilizers, drugs, pesticides, natural waters, is the Kjeldahl method, based on the conversion of the bound nitrogen to
ammonia, which is then separated by distillation and determined by titration with hydrochloric acid as the titrant. A large variety of
esters can be determined by a potentiometric acid-base back-titration procedure, where an excess of potassium hydroxide added for
saponification of the ester is determined. One very interesting acid-base titration is the determination of metal ions by the titration
of hydrogen ions released by a complexation reaction.
The apparent acidity or basicity of a compound is strongly dependent on the solvent used. For example, phenol is a too weak
acid for titration in aqueous solution. Very strong acids such as hydrochloric and perchloric acids cannot be titrated individually in
water. However, such titrations can be performed successfully when appropriate nonaqueos solvents are used. Bases too weak to be
titrated in water can be titrated in protogenic solvents (solvents more acidic than water), for example acetic acid. In this solvent
amines, medicinal sulfonamides, and most common alkaloids such as caffeine can be titrated with perchloric acid dissolved in the
same solvent. Acids too weak to be titrated in water appear much stronger in protophilic solvents (solvents more basic than water).
For example, ethylenediamine can be employed as a solvent for titration of phenols. Some solvents permit differentiation (or
stepwise titration) of a series of acidic or basic species that, in water, titrate either together or not at all. This group of solvents
includes dioxan, ketones and hydrocarbons. For example, perchloric, hydrochloric and acetic acids, and phenol, can be titrated in
4-methyl-2-pentanone solvent to obtain an endpoint for each compound, using tetrabutylammonium hydroxide in 2-propanol as
titrant. Examples of acid-base titrations in nonaqueous solvents are given in Table 1.

Precipitation Titrations
The potentiometric detection of the endpoint of precipitation titrations is very often used because not many visual indicators are
available, in particular when mixtures of analytes are titrated. Halides, cyanide, sulfide, chromate, mercaptans and thiols can be
titrated with silver nitrate, using the silver sulfide-based ISE. Also complex mixtures, such as sulfide, thiocyanide and chloride ions,
or chloride, bromide and iodide ions, can be titrated potentiometrically with silver(I) ions. When the solubility of a compound
formed during titration is too high, nonaqueous or mixed solvents are used, for example, when sulfates are titrated with lead(II)
using a lead ion-selective electrode. In this case water-ethanol solvent is used. Table 2 presents some examples of potentiometric
precipitation titrations.
TITRIMETRY | Potentiometry 7

Table 1 Examples of acid-base titrations in nonaqueous solvents

Solvent Analyte Titrant

Methanol, ethanol Inorganic and carboxylic acids Cyclohexylamine in methanol or 2-propanol, CH3OK in methanol
2-Propanol Carboxylic acids (C4H9)4NOH in 2-propanol
2-Methyl-2-propanol Acid mixtures (C4H9)4NOH in 2-methyl-2-propanol
Acetic acid Amines, aminoacids, salts, weak acids, sulfonamides HClO4 in dioxan or acetic acid
Dimethylformamide Carboxylic acids, phenols, ammonium salts CH3OK in methanol or benzene
Glycol þ hydrocarbons Salts of weak acids, amines HClO4 in dioxan
Benzene þ methanol Carboxylic acids, acid anhydrides KOH in methanol, CH3OK in benzene þ methanol
Ethylene diamine Weak acids, phenols (C4H9)4NOH in 2-methyl-2-propanol
Acetonitrile Amines, carboxylic acids, HClO4 in dioxan, CH3OK in acetonitrile
Chloroform Acids, alkaloids C2H5ONa in ethanol, HClO4 in acetic acid
Nitromethane Acid amides, urea, amine oxides HClO4 in dioxan
4-Methyl-2-pentanone Acids, acid mixtures (C4H9)4NOH in 2-propanol
Heterocyclic amines HClO4 in dioxan

Table 2 Examples of typical potentiometric precipitation titrations of inorganic and organic compounds

Analyte Titrant Indicator electrode Conditions


3 þ
AsO4 Ag Ag pH 9
Br Agþ Ag or Ag-ISE Dil. HNO3, 0.1 mol l1 KNO3
Ce3þ MoO42 Pt In presence of Ce3þ
Cl Agþ Ag or Ag-ISE Dil. HNO3, 0.1 mol l1 KNO3
Cl Hg22þ Hg Dil. HNO3
F La3þ F-ISE Neutral soln., 1:1 methanol
Fe(CN)64 Agþ Ag
Fe(CN)63 Agþ Ag
I Agþ Ag or Ag-ISE 0.1 mol l1 HNO3
Kþ Tetraphenylborate K-ISE
Liþ NH4F F-ISE 95% methanol
S2 Agþ S-ISE Alkaline soln.
SO42 Pb2þ Pb-ISE 50% dioxan or 1:1 ethanol
Zn2þ K4Fe(CN)6 Pt pH 3, in presence of Fe(CN)63
Barbiturates Agþ Ag or Ag-ISE Borate buffer, ethanol
Cysteine HgCl2 Hg pH 2–6
Dithioxamide Agþ Ag or Ag-ISE
Fatty acids salts Agþ Ag Aqueous soln.
Mercaptanes Hg(ClO4)2 Hg
Oxalates Agþ Ag or Ag-ISE 0.03 mol l1 NH3
Sulfonamides Agþ Ag pH 7–8
Thioacetamide Agþ Ag pH 8.5–9
Thiourea Agþ Ag 2 mol l1 NH3

Compleximetric Titrations
Many metal ions (e.g., calcium, cadmium, aluminum, lead) can be titrated with standard ethylenediaminetetraacetic acid (EDTA)
or other compleximetric titrants, using an appropriate indicator electrode. When no direct appropriate indicator electrode exists,
the addition of indicator metal ions can permit a determination. For example, barium may be titrated with EDTA in the presence of
silver-EDTA complex as an indicator reagent using a silver electrode. Examples of compleximetric titrations are given in Table 3.

Redox Titrations
Potentiometric titrations of oxidizing and reducing substances can be performed potentiometrically using inert electrodes, such as
platinum or gold electrodes. Typical examples of applications include the titration of iron(II) ion with permanganate, titration of As
(III) with bromine or iodine, and the determination of ascorbic acid with iodine. Many organic compounds such as azo, nitro, nitroso,
carbonyl compounds, and quinones may also be determined. Because of their limited solubility in water those compound are titrated
in mixed or nonaqueous solvents. The Karl Fischer method of water determination based on the redox reaction between sulfur dioxide
and iodine in the presence of a stoichiometric amount of water may be also mentioned. Water in liquid samples is titrated with titrant
solutions containing iodine, sulfur dioxide, pyridine and methanol as active solvent. Potentiometry, or more often bipotentiometry,
is used for detection of the endpoint of this titration. Some examples of potentiometric redox titrations are given in Table 4.
8 TITRIMETRY | Potentiometry

Table 3 Examples of typical compleximetric titrations

Indicator electrode Titrant Analyte Conditions

Ag EDTA 2þ
Ba , Ca , Mg2þ 2þ
pH 9–10.5, in presence of small amount of Agþ
Ca-ISE EDTA Ca2þ pH 10–12
Cu-ISE EDTA Cd2þ, Co2þ, Mn2þ, Ni2þ, Pb2þ, Zn2þ pH 7–8
Cu-ISE TETREN Cu2þ, Ni2þ, Zn2þ pH 4–6, acetate buffer
Cu-ISE CuSO4 Chelating agents pH 5–6
F-ISE NaF Al3þ pH 3.5–4.0
Hg I Hg2þ
Pt EDTA Bi3þ, In3þ, Th4þ 0.2 mol l1 chloroacetic acid, in presence of Fe3þ/Fe2þ
Pt EDTA Al3þ, Cd2þ, Cu2þ, Ni2þ, Pb2þ, Zn2þ Ammonium acetate, pH 5–6

TETREN, tetraethylenepentamine.

Table 4 Typical potentiometric redox titrations of inorganic and organic compounds

Analyte Titrant Conditions

As 3þ
Ce 4þ
4 mol l1 HCl, in presence of ICl
C2O42 MnO4 0.2–1 mol l1 H2SO4, 70  C
Ce3þ MnO4 1 mol l1 H2SO4
Co2þ Fe3þ Excess of 1,10-phenanthroline, pH 2–4
Cr2O72 As3þ 20% H2SO4
H2O2 MnO4 0.02 mol l1 H2SO4
I MnO4 0.1 mol l1 H2SO4
NO2 Pb(CH3COO)4 1 mol l1 NaCl
Ti4þ Cr2þ 2 mol l1 H2SO4, Hg electrode, absence of O2
Tlþ Chloramine T 3–4 mol l1 HCl, in presence of iodine chloride
Aldehydes KMnO4 1.5 mol l1 H2SO4
Alkenes Br2 in CH3COOH In CH3COOH
Amines NaNO2 Acidic solution
Ascorbic acid Br2 in CH3COOH In CH3COOH
Carboxylic acids KMnO4 0.1 mol l1 NaOH
Cysteine KIO4 9 mol l1 HCl
Hydroquinone Pb(CH3COO)4 or Br2 in CH3COOH In CH3COOH
Nitro-, Azo- compounds Cr(CH3COO)2 In CH3COOH
Phenols K2Cr2O7 H2SO4
Quinone Na2S2O3 CH3COOH þ H2O
Thiols Pb(CH3COO)4 or Br2 in CH3COOH In CH3COOH
Thiourea Chloramine T HCl þ KBr

In most cases a Pt electrode was used as the indicator electrode.

Potentiometric Coulometric Titrations


An important modification of conventional potentiometric titration is based on the fact that instead of adding a precisely measured
volume of a titrant of known concentration to the analyzed solution, it is possible to create such conditions that the analyte
undergoes an electrochemical reaction in strictly controlled conditions, of which completeness is potentiometrically controlled.
Such procedure in specially controlled conditions may be used for many procedures known in classical potentiometric titrations.
The success of coulometric process depends on the strict correlation between the amount of measured electric charge and the
amount of the analyte being determined. In most cases this may be simple when the current efficiency is 100% versus the
determined substance. This is however not the case when the concentration of the analyte drops down and the current is kept
constant due to the occurance of another reaction, for example with solvent species present in the solution. To avoid this a suitable
depolarizer should be added, which in turn will react with the analyte. As a result the total current efficiency will correspond to the
needed amount of the analyte.
There are many advantages of such procedure. The most basic one is the elimination of laborious and time-consuming careful
preparation, standardization and storage of standard solutions in particular when they are unstable in dilute solution. One most
important aspect of measuring the electric charge instead of volume and concentration of the solution is the fact that both charge
and time are fundamental quantities. In this sense the coulometric titration can be considered as an absolute procedure. The
charge consumed in that case can be estimated to 0.01% which is not possible in volumetric potentiometric titration. Therefore
such procedures are applicable for determination and studies of compounds available in minute quantities.
The acid-base coulometric potentiometric titrations may be used for the determination of both weak and strong bases and acids
in aqueous as well as in water-like solutions and also in some organic solvents. Such titrations should be often carried out in special
TITRIMETRY | Potentiometry 9

cells where the counter electrode is isolated by a diaphragm to avoid disturbing processes occurring at the auxiliary counter
electrode. Numerous redox titrations have been also carried out using the coulometric version in particular when the titrant is
unstable and cannot be used in classical titrations. Such titrant as Cuþ, Sn2þ, Cr2þ, and Fe2þ can be easily generated at the anode,
whereas strong and unstable oxidants as Br2, MnO4 or Ag2þ are generated in the cathodic compartment containing the sample.
The advantages of the coulometric potentiometric titrations result in the construction of commercial instruments for laboratory
and industrial applications.

Further Reading
1. Bakker, E.; Pretsch, E. Advances in Potentiometry. In Bard, A. J., Zoski, C. G., Eds.; Electroanalytical Chemistry; CRC Taylor and Francis Group: Boca Raton, FL, 2012; Vol. 24,
pp 1–73, Chapter 1.
2. Bard, A. J.; Faulkner, I. R. Electrochemical Methods, Fundamentals and Applications, 2nd ed.; Wiley: New York, 2000.
3. Brett, C.; Brett, A. M. O. Electroanalysis. Oxford Sciences: Oxford, 1998.
4. Gyenes, J. Titrationen in nichtwässrigen Lösungsmitteln. Enke: Stuttgart, 1970.
5. Henze, G.; Neeb, R. Electrochemische Analytik. Springer Verlag: Berlin, 1986.
6. Lingane, J. J. Electroanalytical Chemistry. Interscience Publishers: New York, 1958.
7. Mascini, M. Ion Sel. Electrode Rev. 1980, 2, 17–71.
8. Midgley, D. Potentiometric Water Analysis. Wiley: New York, 1974.
9. Serjeant, E. P. Potentiometry and Potentiometric Titrations. Wiley: New York, 1984.
10. Svehla, G. Automatic Potentiometric Titration. Pergamon Press: Oxford, 1978.
11. Vanysek, P. Modern Techniques in Electroanalysis. Wiley: Chichester, 1995.
12. Vytras, K. Ion Sel. Electrode Rev. 1985, 7, 77–164.
13. Wang, J. Analytical Electrochemistry, 3th ed.; Wiley: New York, 2006.

You might also like