Solvent Extraction Principles and Practice PDF
Solvent Extraction Principles and Practice PDF
Solvent Extraction Principles and Practice PDF
and Praciice
Second Edition, Revised and Expanded
edited by
Jan Rydberg
ChaZmers University of TechnoZogy
Goteborg, Sweden
Michael Cox
University of Hertfordshire
Hatjield, Hertfordshire, United Kingdom
Claude Musikas
Commissariat a I’ Energie Atomique
Paris, France
Gregory R. Choppin
Florida State University
Tallahassee, Florida, U.S.A.
MARCEL
MARCELDEKKER,
INC. NEWYORK * BASEL
DEKKER
Previous edition published as Principles and Practices of Solvent Extraction
(Rydberg, J. et al., Eds.), Marcel Dekker, 1992.
Although great care has been taken to provide accurate and current information,
neither the author(s) nor the publisher, nor anyone else associated with this pub-
lication, shall be liable for any loss, damage, or liability directly or indirectly caused
or alleged to be caused by this book. The material contained herein is not intended to
provide specific advice or recommendations for any specific situation.
Trademark notice: Product or corporate names may be trademarks or registered
trademarks and are used only for identification and explanation without intent to
infringe.
Library of Congress Cataloging-in-Publication Data
A catalog record for this book is available from the Library of Congress.
ISBN: 0-8247-5063-2
This book is printed on acid-free paper.
Headquarters
Marcel Dekker, Inc., 270 Madison Avenue, New York, NY 10016, U.S.A.
tel: 212-696-9000; fax: 212-685-4540
Distribution and Customer Service
Marcel Dekker, Inc., Cimarron Road, Monticello, New York 12701, U.S.A.
tel: 800-228-1160; fax: 845-796-1772
Eastern Hemisphere Distribution
Marcel Dekker AG, Hutgasse 4, Postfach 812, CH-4001 Basel, Switzerland
tel: 41-61-260-6300; fax: 41-61-260-6333
World Wide Web
http://www.dekker.com
The publisher offers discounts on this book when ordered in bulk quantities. For
more information, write to Special Sales/Professional Marketing at the headquarters
address above.
Copyright # 2004 by Marcel Dekker, Inc. All Rights Reserved.
Neither this book nor any part may be reproduced or transmitted in any form or by
any means, electronic or mechanical, including photocopying, microfilming, and
recording, or by any information storage and retrieval system, without permission in
writing from the publisher.
Current printing (last digit):
10 9 8 7 6 5 4 3 2 1
PRINTED IN THE UNITED STATES OF AMERICA
Preface
iii
iv Preface
Jan Rydberg
Michael Cox
Claude Musikas
Gregory R. Choppin
Contents
Preface iii
Contributors vii
v
vi Contents
Appendix 715
Index of Compounds 725
Subject Index 741
Contributors
*Retired.
vii
viii Contributors
*Retired.
{
Deceased.
Contributors ix
*Retired.
†The International Union of Pure and Applied Chemistry (IUPAC) recommends the use of the
term liquid-liquid distribution. However, more traditionally the term solvent extraction (sometimes
abbreviated SX) is used in this book.
1
2 Cox and Rydberg
Fig. 1.2 Model of a four-phase system consisting of two liquid phases (e.g., an aque-
ous and an organic phase) in equilibrium with a gas phase and a solid phase.
Fig. 1.3 Percentage of extraction of various metals from a solution of dissolved stain-
less steel scrap, at 40oC. The organic phase is 25% tertiary amine (Alamine 336), 15%
dodecanol (Loral C12) and 60% kerosene (Nysolvin 75A). The aqueous phase is a CaCl2
solution at pH 2.
↓↑
← +
Aqueous phase: HBz → H + Bz −
In 1891, Nernst realized that only if the solute has the same molecular weight
in the organic phase as in the aqueous phase, the distribution ratio would be
independent of the concentration of the solute, or distribuend. He proposed the
simple relation:
10 Cox and Rydberg
Although Kex neither describes the intermediate reaction steps (see Chapters 3
and 4) or kinetics of the reaction (Chapter 5), nor explains why or to what
extent HgCl2 dissolves in the organic solvent (Chapter 2), the extraction reaction
is a useful concept in applied solvent extraction, and Kex values are commonly
tabulated in reference works [3–4].
1.4.2 Extractants
During the years 1900 to 1940, solvent extraction was mainly used by the or-
ganic chemists for separating organic substances. Since in these systems, the
Introduction to Solvent Extraction 11
solute (or desired component) often exists in only one single molecular form,
such systems are referred to as nonreactive extraction systems; here the distribu-
tion ratio equals the distribution constant.
However, it was also discovered that many organic substances, mainly
weak acids, could complex metals in the aqueous phase to form a complex
soluble in organic solvents. A typical reaction can be written
← MA (org) + z H + (aq)
M z + (aq) + z HA(aq or org) → (1.6)
z
which indicates that the organic acid HA may be taken from the aqueous or the
organic phase. This is an example of reactive extraction. It became a tool for
the analytical chemist, when the extracted metal complex showed a specific
color that could be identified spectrometrically. The reagent responsible for
forming the extractable complex is termed the extractant.
The industrial use of solvent extraction of inorganic compounds grew out
of the analytical work. As both areas, analytical as well as industrial, needed
both better extractants and an understanding of the reaction steps in the solutions
in order to optimize the applications, theoretical interpretations of the molecu-
lar reactions in the solutions became a necessity, as will be described in later
chapters.
The increased use of computer graphics for modeling molecular structures
and chemical reactions has opened a path for the synthesis of tailor-made ex-
tractants. Thus the future promises new varieties of extractants with highly se-
lective properties for the desired process.
world production) is now produced annually. This and many other processes are
described in later chapters.
For these applications, the technique of solvent extraction had to be further
developed and with this a new terminology was also developed. This can be
illustrated by considering a process where a desired component in an aqueous
solution is extracted with an organic reagent (extractant) dissolved in another
organic liquid; note here that the term “organic solvent” is not used because of
possible confusion. The term “solvent” could be used for the whole organic
phase or for the organic liquid in which the organic extractant is dissolved. Thus
the term generally given to the latter is (organic) diluent.
While in laboratory experiments the extraction vessel may be a test tube,
or more conveniently some kind of separation funnel (Fig. 1.1), this is not suited
for industrial use. Industry prefers to use continuous processes. The simplest
separation unit is then the mixer-settler, or some clever development of the same
basic principle, as described in Chapter 9. Figure 1.5 pictures a simple mixer-
settler unit, here used for the removal of iron from an acid solution also contain-
ing nickel and cobalt (the same systems as in Fig. 1.3). The mixer (or contactor)
is here simply a vessel with a revolving paddle that produces small droplets of
one of the liquid phases in the other. This physical mixture flows into and
slowly through the separation vessel, which may be a long tank; through the
influence of gravity the two phases separate, so that the upper organic kerosene–
octanol–amine phase contains the Fe(III) and the lower aqueous CaCl2 phase
contains the Co(II) and Ni(II). Numerous variations of the construction of
mixer-settlers (or MS-units, as they are abbreviated) exist (Chapter 9), often
several joined together into MS-batteries. Many such will be described later on.
A diagram of a full basic process is given in Fig. 1.6 to illustrate the
common terminology. The incoming aqueous solution is called the feed. It is
Fig. 1.5 The principle of a mixer-settler unit, e.g., for separation of iron(III) from
nickel and cobalt.
Introduction to Solvent Extraction 13
Fig. 1.7 Arrangements of solvent extraction stages: (a) co-current; (b) cross-current;
and (c) counter-current.
Introduction to Solvent Extraction 15
ies have been made of the distribution or extraction constants, as enthalpy and
entropy values give a good indication of the driving force of the extraction
and also may indicate the structure of the molecular species in the organic
solvent.
Separation of metals by solvent extraction is usually based on the various
complexing properties of the metals (Chapter 3). Separation systems may be
chosen on the basis of complexity constants obtained from the literature. How-
ever, the literature often shows different values for “same systems” causing
considerable concern for process design chemists. There is an obvious need for
an objective presentation of the uncertainty in the published equilibrium con-
stants, however conditional they may be.
Although theories of solution (Chapter 2) and formation of extractable
complexes (Chapters 3 and 4) now are well advanced, predictions of distribution
ratios are mainly done by comparison with known similar systems (Chapter
3). Solvatochromic parameters, solubility parameters, and donor numbers, as
discussed in Chapters 2–4, are so far mainly empirical factors. Continuous ef-
forts are made to predict such numbers, often resulting in good values for sys-
tems within limited ranges of conditions. It is likely that these efforts will suc-
cessively encompass greater ranges of conditions for more systems, but much
still has to be done. In the future, theory may allow the assignment of exact
numbers to the solvatochromic parameters, thus also permitting theoretical pre-
dictions of distribution constants for known as well as for hypothetical (not yet
synthesized) extractants.
Chemical quantum mechanics (Chapter 16) are now promising to contrib-
ute to prediction of distribution ratios, particularly when used in interactive com-
puter modeling of chemical structures and reactions. This will provide better
understanding of solute–solvent interactions and the theory of solubility, which
is the foundation for prediction of distribution ratios. Progress is likely to speed
up as efficient computing programs with large data bases become more easily
available on the market at more attainable prices. Such interactive computer
research will be able to explain poor extractability due to steric hindrance, hy-
dration, the synergistic effect of specific adducts, etc., and, as a consequence,
give clues to better extraction reagents and conditions.
A combination of known extractants can increase the net extraction both
in degree and kinetics through synergistic extraction. This subject also is poorly
explored, though it is of considerable importance to industrial solvent extraction
processes.
1.5.2 Kinetics
All chemical reactions occur with a certain rate. For aqueous systems, in which
no redox reactions occur (e.g., simple complex formation), equilibrium is often
Introduction to Solvent Extraction 17
attained rapidly, which is also true for adduct reactions in the organic phase, all
at normal temperatures. However, the transfer of a solute from an aqueous solu-
tion to an organic solvent occurs via a phase boundary; such reactions depend
on several parameters and may be quite slow. Therefore, the kinetics of solvent
extraction, which is of paramount importance to all industrial applications, is
largely determined by the interfacial chemistry. To speed up phase transfer, the
engineer tries to maximize the interfacial surface by, e.g., violent stirring, pro-
ducing billions of small droplets. Though the theory is well advanced (see Chap-
ters 5 and 9), the interface is very complicated and its properties are difficult to
investigate, particularly at the molecular level. Few applicable techniques are
available, and the results are often difficult to interpret. The coalescence of the
droplets (to produce two clean phases that are easy to separate) presents another
interfacial surface problem, and so-called phase reversals may further compli-
cate the separation. This is particularly important in the development of new
types of contactor. Much progress is required here and as such presents a chal-
lenge to future scientists and engineers.
The largest research effort in extraction kinetics is likely to be in the
development of solvent extraction related techniques, such as various versions
of liquid chromatography, liquid membranes, etc. These techniques require a
detailed knowledge of the kinetics of the system to predict the degree of separa-
tion.
It is a practical fact that most industrial solvent extractions are carried out
under nonequilibrium conditions, however close the approach may be; for exam-
ple, centrifugal contactor-separators (Chapter 9) rarely operate at distribution
equilibrium. An interesting possibility is to expand this into extractions further
from equilibrium, if the kinetics of the desired and nondesired products are
different. Such operations offer a real technlogical challenge.
taking place, to even smaller sizes (volumes and amounts)—the nano experi-
ments. Progress in this field requires a high degree of ingenuity.
Although single-stage laboratory techniques provide the first step toward
multistage industrial processes, such process development usually requires
small-scale multistage and pilot-plant scale equipment. A large number of excel-
lent designs are available, and we consider further fundamental improvements
unlikely.
The industrial application of solvent extraction is a mature technique, and
it is now possible to move from laboratory experiments on a new extraction
system to full industrial practice with little technological risk. There is a suffi-
cient variety of large-scale equipment available to cope with most problems
encountered in application, although much of the equipment remains rather mas-
sive. Attempts to miniaturize, for instance, by using centrifugal forces to mix
and separate phases, still has to be developed further.
Many industrial processes begin with a leaching step, yielding a slurry
that must be clarified before solvent extraction. The solid–liquid separation is a
costly step. The solvent extraction of unclarified liquids (“solvent-in-pulp”) has
been proposed to eliminate solid–liquid separation. The increased revenue and
reduced energy cost make this an attractive process, but many problems remain
to be solved: loss of metals and extractants to the solid phase, optimization of
equipment design, effluent disposal, etc.
An essential step in industrial solvent extraction is the regeneration of the
extractant. This can be done in many ways, e.g., by distillation, evaporation, or
stripping (back-extraction). While distillation and evaporation do not discrimi-
nate between solutes (the diluent is simply removed by heating), stripping, by
careful choice of strip solution and conditions, can be made highly selective.
Alternatively, all the solutes can be stripped and then subjected to a selective
extraction by changing the extractant; examples of both types of process will be
found in Chapter 13. The possibilities are many, and it may be worthwhile to
explore new paths.
Membrane extraction is a relatively new technique for solvent extraction,
in which a solute is transferred from one aqueous phase to another through a
membrane holding an extractant dissolved in a diluent. This ingenious scheme
has been only slightly explored, though it offers great potential for the future,
e.g., for waste water cleaning.
The step from laboratory experiments via pilot plants to industrial scale
requires serious consideration of all the points here; practical experience is in-
valuable in order to avoid mistakes and excess costs, as indicated in Chapter 7.
1.5.7 Spin-off
The principle of solvent extraction—the distribution of chemical species be-
tween two immiscible liquid phases—has been applied to many areas of chemis-
try. A typical one is liquid partition chromatography, where the principle of
solvent extraction provides the most efficient separation process available to
organic chemistry today; its huge application has become a field (and an indus-
try!) of its own. The design of ion selective electrodes is another application of
the solvent extraction principle; it also has become an independent field. Both
these applications are only briefly touched upon in the chapter of this book on
analytical applications (Chapter 14), as we consider them outside the scope of
Introduction to Solvent Extraction 21
1.5.8 Conclusion
Solvent extraction is a mature technique in that extensive experience has led to
a good understanding of the fundamental chemical reactions. At the same time,
compared to many other chemical separation processes like precipitation, distill-
ation, or pyrometallurgical treatment, the large-scale application of solvent ex-
traction is, nevertheless, a young technique. New reagents are continually being
developed, spurred on by computer modeling, and more efficient contacting
equipment is coming into use. Considering such factors as demands for higher
product purity, less pollution, and the need for recovering substances from more
complex matrices and more dilute resources, the efficiency and high selectivity
of solvent extraction should make it an increasingly competitive separation pro-
cess both in research and in industry.
REFERENCES
1. Freiser, H.; and Nancollas, G. H.; Compendium of Analytical Nomenclature. Defini-
tive Rules 1987. IUPAC. Blackwell Scientific Publications, Oxford (1987).
2. Blass, E.; Liebl, T.; Häberl, M.; Solvent Extraction—A Historical Review, Proc. Int.
Solv. Extr. Conf. Melbourne, 1996.
Introduction to Solvent Extraction 25
2.1 INTRODUCTION
Solvent extraction is another name for liquid–liquid distribution, that is, the
distribution of a solute between two liquids that must not be completely mutu-
ally miscible. Therefore, the liquid state of aggregation of matter and the essen-
tial forces that keep certain types of liquids from being completely miscible are
proper introductory subjects in a study of solvent extraction. Furthermore, the
distribution of a solute depends on its preference for one or the other liquid,
which is closely related to its solubility in each one of them. Thus, the general
subject of solubilities is highly relevant to solvent extraction.
In a solution, the solute particles (molecules, ions) interact with solvent
molecules and also, provided the concentration of the solute is sufficiently high,
with other solute particles. These interactions play the major role in the distribu-
tion of a solute between the two liquid layers in liquid–liquid distribution sys-
tems. Consequently, the understanding of the physical chemistry of liquids and
solutions is important to master the rich and varied field of solvent extraction.
Solvent extraction commonly takes place with an aqueous solution as one
liquid and an organic solvent as the other. Obviously, the extraction process is
limited to the liquid range of these substances. Since solvent extraction is gener-
ally carried out at ambient pressures, the liquid range extends from about the
freezing temperature up to about the normal boiling temperature. If, however,
high pressures are applied (as they are in some solvent extraction processes),
then the liquid range can extend up to the critical temperature of the substance.
Supercritical fluid extraction beyond the critical temperature (such as decaffei-
nation of coffee with supercritical carbon dioxide) is a growing field of applica-
*Retired.
27
28 Marcus
tion of solvent extraction. It has the advantages that the properties of the super-
critical fluid can be fine-tuned by variation of the pressure, and that this
“supercritical solvent” can be readily removed by a drastic diminution of the
pressure, but has drawbacks related to the high temperatures and pressures often
needed.
Numerous solvents are used in solvent extraction. They can be divided in
the context of solvent extraction into different classes as follows (but see Refs.
[1] and [2] for some other classification schemes):
Class 1: Liquids capable of forming three-dimensional networks of strong
hydrogen bonds, e.g., water, poly- and amino-alcohols, hydroxy-acids, etc.
Class 2: Other liquids that have both active hydrogen atoms and donor
atoms (O, N, F; see Chapter. 3), but do not form three-dimensional networks
(rather forming chainlike oligomers), e.g., primary alcohols, carboxylic acids,
primary and secondary amines, nitro compounds with α-positioned hydrogen
atoms, liquified ammonia, etc. They are generally called protic or protogenic
substances.
Class 3: Liquids composed of molecules containing donor atoms, but no
active hydrogen atoms, e.g., ethers, ketones, aldehydes, esters, tertiary amines,
nitro compounds without α-hydrogen, phosphoryl-group containing solvents,
etc. (see Table 4.3). They are generally called dipolar aprotic substances.
Class 4: Liquids composed of molecules containing active hydrogen
atoms but no donor atoms, e.g., chloroform and some other aliphatic halides.
Class 5: Liquids with no hydrogen-bond forming capability and no donor
atoms, e.g., hydrocarbons, carbon disulfide, carbon tetrachloride, supercritical
carbon dioxide, etc.
This diversity in solvent properties results in large differences in the distri-
bution ratios of extracted solutes. Some solvents, particularly those of class 3,
readily react directly (due to their strong donor properties) with inorganic com-
pounds and extract them without need for any additional extractant, while others
(classes 4 and 5) do not dissolve salts without the aid of other extractants. These
last are generally used as diluents for extractants, required for improving their
physical properties, such as density, viscosity, etc., or to bring solid extractants
into solution in a liquid phase. The class 1 type of solvents are very soluble in
water and are useless for extraction of metal species, although they may find
use in separations in biochemical systems (see Chapter 9).
cal, its viscosity. For water, the viscosity is only 0.89 mPa s−1 (nearly 1 centi-
poise) at 25°C [1,2], and water is considered a highly fluid liquid.
Contrary to a fluid in the gaseous state, a liquid has a surface, and is
characterized by a surface tension. For water, the surface tension is 72 mN m−1
at 25°C [1,2]. Again, contrary to the fluid in the gaseous state, the volume of a
liquid does not change appreciably under pressure; it has a low compressibility
and shares this property with matter in the solid (crystalline, glassy, or amor-
phous) state. For water, the compressibility is 0.452 (GPa)−1 at 25°C [1,2]. These
are macroscopic, or bulk, properties that single out the liquid state from other
states of aggregation of matter.
There are also some general microscopic, or molecular, properties that are
peculiar to the liquid state. Contrary to the crystalline solid, the particles of a
liquid do not possess long-range order. Although over a short range—2 to 4
molecular diameters—there is some order in the liquid, this order dissipates at
longer distances. A particle in the liquid is free to diffuse and, in time, may
occupy any position in the volume of the liquid, rather than being confined at
or near a lattice position, as in the crystalline solid. Contrary to particles in the
gaseous state, however, the particles in a liquid are in close proximity to each
other (closely packed) and exert strong forces on their neighbors [3].
The close packing of the molecules of a substance in the liquid state re-
sults in a density much higher than in the gaseous state and approaching that in
the solid state. The density, ρ, is the mass per unit volume, and can be expressed
as the ratio of the molar mass M to the molar volume V of a liquid. Table 2.1
lists the values of the properties M and V of representative liquids that are
important in the field of solution chemistry and solvent extraction. The densities
and molar volumes depend on the temperature, and the latter are given for 25°C.
(For a discussion of industrial solvents, see Chapter 12.)
Many liquids used in solvent extraction are polar. Their polarity is mani-
fested by a permanent electric dipole in their molecules, since their atoms have
differing electronegativities. Oxygen and nitrogen atoms, for instance, generally
confer such dipolarity on a molecule, acting as the negative pole relative to
carbon or hydrogen atoms bonded to them. The dipole moment µ characterizes
such polar molecules and ranges from about 1.15 D (Debye unit = 3.336 10−30
C ⴢ m) for diethyl ether or chloroform, to 4.03 D for nitrobenzene, and 5.54 D
for hexamethyl phosphoric triamide [1,2]. Water is also a polar liquid, with a
moderate dipole moment, µ = 1.83 D. A list of dipole moments of some solvents
that are important in solvent extraction is presented in Table 2.1. Substances
that do not have a permanent dipole moment (i.e., µ = 0) are called nonpolar.
Many hydrocarbons belong to this category, but not all (e.g., toluene has µ =
0.31 D).
When nonpolar liquids are placed in an electric field, only the electrons
in their atoms respond to the external electric forces, resulting in some atomic
30 Marcus
Fig. 2.1 Dipole and hydrogen bond interactions. A schematic representation of (a)
“head-to-tail” dipole-dipole attractive interactions (e.g., in tri-n-octylamine); (b) “head-
to-head” dipole-dipole repulsive interactions caused by steric hindrance (e.g., in dibutyl
sulfoxide); (c) chainlike dipole-dipole interactions (e.g., in 1-octanol); (d) a cyclic, hy-
drogen-bonded dimer (e.g., in hexanoic acid).
Principles of Solubility and Solutions 33
and, at all points, is under the same pressure. Such a region of space is called a
phase. At equilibrium, at a constant temperature and pressure, a liquid phase is
homogeneous and is isotropic: it has the same properties in all directions. In
solvent extraction, we are interested in two liquid phases in contact, normally
under the same temperature and pressure, which eventually reach mass transfer
equilibrium. Such an equilibrium is a dynamic process: as many molecules of
one solvent enter the other as leave it and return to their original phase as do
the molecules (ions) of the solute(s). If each of the two liquids is originally a
pure chemical substance and there is (as yet) no solute to distribute, we speak
of a two-component, two-phase system, but the net effect is that an amount of
the substance from one liquid phase is transferred to the other and vice versa.
For two practically immiscible liquids this transferred amount is small relative
to the total amount of the two phases. There may or may not be a vapor phase
present, but this is generally disregarded in the context of solvent extraction.
Fig. 2.2 Liquid immiscibility. The guaiacol (A) + glycerol (B) system happens to have
a closed miscibility loop. The (phase) coexistence curves are shown on the left-hand side
(a) for lower temperatures, at which a lower critical solution temperature (LCST), Tcs =
40°C, is seen, and on the right-hand side (b) for higher temperatures, where a UCST,
Tcs = 82°C, is seen. The compositions of the A-rich phases′′ and the B-rich phases′ are
shown at 50°C and 70°C, respectively.
36 Marcus
ln x A ′ ≈ −VB* (δ A − δ B ) 2 / RT (2.8)
where x′A is the mole fraction of component A in the B-rich phase marked by
′. A similar equation holds for x″B, in the A-rich phase, marked ″, with V *A
replacing V *B . Thus, the composition of the two phases can also be estimated
by the application of the solubility parameters. These considerations can also
be used for an estimate of the distribution of a solute between the immiscible
liquids A and B, assuming the conditions for regular mixing to be fulfilled for
all the components (see section 2.8.2).
As an example consider the ethylene glycol (A)–benzene (B) system at
25°C, assuming it to be a regular mixture (actually it is not). The relevant
quantities are V *A = 55.9 and V *B = 89.9 cm3 mol−1, δA = 32.4 and δB = 16.0 J1/2
cm−3/2. Hence for the equimolar mixture xA = xB = 0.5, according to Eq. (2.7),
bAB = 54.7 kJ mol−1 Ⰷ RT/ 2xAxB = 5 kJ mol−1, so that this system ought to split
into two liquid phases, as it does. In the benzene-rich phase (′) the ethylene
glycol content should be 0.005 wt% according to Eq. (2.8) and the benzene
content in the ethylene glycol–rich phase (′′) should be 0.31 wt%. Most regular
mixtures, however, would form a single homogeneous liquid phase, since the
condition of Eq. (2.6) is rarely met.
amines, up to a total of about five carbon atoms in all the chains, including
pyridine, but not with nitro compounds. Water is also miscible with many difun-
ctional liquid compounds, such as diamines, alcohol amines, amides, and others,
and with such compounds as dimethylsulfoxide [1–3].
The higher members of homologous series based on these compounds
become increasingly immiscible with water as the chain length increases or as
aromatic rings are added. Consequently, these higher members can be used as
solvents in extraction systems. Table 2.2 lists the mutual solubilities, in weight
percent (wt%), of water and representative organic solvents at 25°C, unless
noted otherwise.
Table 2.2 Mutual Solubility of Water and Some Organic Solvents at 25°C
Solvent in water Water in solvent
Solvent wt% wt%
solute and solvent have suitable electron pair donation and acceptance proper-
ties. Hydrogen bonds between the solute and neighboring solvent molecules
may form if one or the other or both are protic, i.e., have hydrogen atoms
bonded to electronegative atoms. The solute particle may also undergo changes
in the solvation process, its internal structure, if it has one, being affected by
the strengthening or weakening of certain bonds, by a redistribution of the par-
tial electrical charges on its atoms, or by the favoring of a certain conformation
of a flexible solute molecule.
Spectroscopic measurements may, in certain cases, yield direct informa-
tion on these interactions. On the other hand, thermodynamic values, obtained
by measuring certain bulk properties of the system, require the aid of statistical
mechanical methods to be related to specific interactions between the solute and
the solvent. However, the thermodynamic aspects of the solute–solvent interac-
tions reflect the preference of the solute for one solvent over another and,
thereby, determine distribution of the solute in a solvent extraction system.
assumed for the processes discussed elsewhere in this chapter, and is only em-
phasized here for Eq. (2.9) by the subscript eq for ∆solnGB and the ratio of con-
centrations, cB(l)/cB(g). The concentration in the gas phase is, of course, given
by the pressure according to the ideal gas law cB(g) = nB/V = P/RT. The ratio of
the concentrations is known as the Bunsen coefficient, KB(B,A), for the solubility
of the gaseous solute B in liquid A. At low pressures and concentrations KB(B,A)
depends only on the temperature and is independent of the pressure and the
concentration.
Whether obtained from an actual experimentally feasible process or from
a thought process, ∆solnG°,B which is obtained from Eq. (2.9) by re-arrangement,
pertains to the solvation of the solute and expresses the totality of the solute–
solvent interactions. It is a thermodynamic function of state, and so are its deriv-
atives with respect to the temperature (the standard molar entropy of solvation)
or pressure. This means that it is immaterial how the process is carried out, and
only the initial state (the ideal gaseous solute B and the pure liquid solvent) and
the final state (the dilute solution of B in the liquid) must be specified.
Because it is a function of state, ∆solnG°B may be considered to be made up
additively of the contributions from the various stages in which the transfer of
the solute particle from the gaseous state into the liquid solvent has been envis-
aged by the foregoing to take place.
The first stage is the creation of the cavity in the liquid to accommodate
the solute. Obviously, work must be done against the cohesive forces of the
liquid that hold its molecules together. This work should be proportional to the
required size of the cavity, and increase as the volume of the solute increases.
An expression for this work would be
∆ cav G = Acav VBδ A2 (2.10)
where Acav is a proportionality coefficient, VB is the molar volume of the solute
B, and δA2 is the cohesive energy density of the solvent A. Thus, in a series
of solvents for a given solute, the positive contribution of cavity formation to
∆solnG°B increases with the squares of the solubility parameters of the solvents.
For a series of solutes and a given solvent, it increases with the molar volumes
of the solutes.
It is more difficult to estimate the contribution from the dispersion forces
to the solute–solvent interactions. Their energy increases with the product of
the polarizabilities of the partners, but decreases strongly with the distance
between them (being proportional to the inverse sixth power of the distance).
The polarizability is related to the molar volume, hence, to the third power of
the linear dimension of the solute or the solvent. Hence, the product of the
polarizabilities depends on the sixth power of the distance between the centers
of the interacting molecules. Consequently, these tendencies balance each
other. For large molecules (e.g., metal chelates, liquid hydrocarbons), it is bet-
42 Marcus
Source: Ref. 2.
44 Marcus
interactions being, therefore, absent), and that a reaction such as ionic dissocia-
tion of the phenol in water is ignored in this example.
Extensive tables of solute parameters are beyond the scope of this book.
Equations (2.10) to (2.12) are meant to show the nature of the dependencies of
the additive terms on various quantities. They enable the prediction of tenden-
cies of solute–solvent interactions for a given solute with a series of solvents or
for a series of solutes with a given solvent.
Table 2.4 Crystal Ionic Radii and Standard Molar Gibbs Free Energies
of Hydration of Ions
−∆hydG° −∆hydG° −∆hydG°
Ion r (nm) (kJ/mol) Ion r (nm) (kJ/mol) Ion r (nm) (kJ/mol)
Source: Ref. 6.
Fig. 2.3 A schematic representation of the hydration layer near a small ion (left) and
a large ion (right), showing the region where the water is dielectrically saturated (with a
low relative permittivity ε′), hence electrostricted (squeezed) and immobilized. The
thickness of this layer, ∆r, depends reciprocally on the size of the ion.
tions that lead to consistent results have been proposed. One of them involves
a reference electrolyte, of which the cation and anion are large, spherical, univa-
lent, and similar in all respects except the sign of the charge. Such a reference
electrolyte is tetraphenylarsonium tetraphenylborate (TATB); the standard molar
Gibbs energy of hydration of each of its ions is ∆hydG° = −38 ± 6 kJ mol−1. On
this basis, values of ∆hydG° of other ions can be obtained, consistent with certain
other extrathermodynamic assumptions, and are listed in Table 2.4 [6].
The differences in the solvation abilities of ions by various solvents are
seen, in principle, when the corresponding values of ∆solvG° of the ions are
compared. However, such differences are brought out better by a consideration
of the standard molar Gibbs energies of transfer, ∆tG° of the ions from a refer-
ence solvent into the solvents in question (see further section 2.6.1). In view of
the extensive information shown in Table 2.4, it is natural that water is selected
as the reference solvent. The TATB reference electrolyte is again employed to
split experimental values of ∆tG° of electrolytes into the values for individual
ions. Tables of such values have been published [5–7], but are outside the scope
of this text. The notion of the standard molar Gibbs energy of transfer is not
limited to electrolytes or ions and can be applied to other kinds of solutes as
well. This is further discussed in connection with solubilities in section 2.7.
µ B = µ ∞B + RT ln xB (2.19)
over a certain (low) concentration range of B in A is said to obey Raoult’s law
for the solvent and Henry’s law for the solute in this concentration range (Fig.
2.4).
The vapor pressure of the solvent in such a case is given by
pA = xA pA* (2.20)
where p*A is the vapor pressure of pure liquid A at the given temperature. Con-
versely, when Eq. (2.20) is followed by the vapor pressure of the solvent,
Raoult’s law is said to be valid and Eq. (2.17) is obeyed. The vapor pressure of
the solute is also proportional to its mole fraction, if Henry’s law is obeyed.
However, the proportionality constant is not its vapor pressure in the pure liquid
state, which may not be attainable at the given temperature. Instead, the expres-
sion
Fig. 2.4 The vapor pressure diagram of a dilute solution of the solute B in the solvent
A. The region of ideal dilute solutions, where Raoult’s and Henry’s laws are obeyed by
the solvent and solute, respectively, is indicated. Deviations from the ideal at higher
concentrations of the solute are shown. (From Ref. 3.)
50 Marcus
pB = K B(A) xB (2.21)
is followed, where KB(A) is called the Henry’s law constant, and depends on the
chemical natures of both B and A, as the molecular description of the foregoing
solutions requires. These considerations apply also when there are several sol-
utes present at low concentrations, so that Raoult’s law may apply to the solvent
and Henry’s law to each of the solutes.
2.4.1.2 Two Liquid Phases
Consider now two practically immiscible solvents that form two phases, desig-
nated by ′ and ″. Let the solute B form a dilute ideal solution in each, so that
Eq. (2.19) applies in each phase. When these two liquid phases are brought into
contact, the concentrations (mole fractions) of the solute adjust by mass transfer
between the phases until equilibrium is established and the chemical potential
of the solute is the same in the two phases:
µ B′′ = µ B∞′′ + RT ln x B′′ = µ B′ = µ B∞′ + RT ln x B′ (2.22)
(It is the difference in the chemical potentials of the solute that is the driving
force for the mass transfer.) This equation can be rewritten in the form:
x B′ / x B′′ = exp[( µ B∞′′ − µ B∞′ ) / RT ] = PB (2.23)
where xB′ /xB″ = DB is the distribution ratio of the solute B (on the mole fraction
scale) and the exponential expression is a constant (at a given temperature),
called the distribution constant of B,* PB. Equation (2.23) is an expression of
Nernst’s distribution law that states:
The distribution ratio of a solute between two liquid phases at equilibrium is
a constant, provided that the solute forms a dilute ideal solution in each phase.
*There is no internationally recommended symbol for the distribution constant, but the symbol P
(for partition constant) is common usage in this context. Coordination (complex) chemists (as in
Chapter 3) prefer the symbol KD (for the German “Konstante” for “constant”); sometimes other
symbols are used. The important point is that the symbol should be properly defined in the text.
Principles of Solubility and Solutions 51
well as the activity a of each substance are the same in all the phases. It is the
inequality of the activities of a substance in two phases that causes some of the
substance to transfer from the one (where it is higher) to the other phase, until
equality is achieved. The activity of a pure liquid or solid substance is defined
as unity. In any mixture, whether ideal or not, the activity of a component A,
aA, is related to its chemical potential by [3]:
µ A = µ°A + RT ln aA (2.24)
Therefore, if component A is the solvent in a solution, and it obeys Raoult’s
law when the solution is dilute in all solutes, then aA = xA under these conditions.
Otherwise, the solution is nonideal, and the deviation from the ideal is described
by means of the activity coefficient
f A = aA / xA (2.25)
where aA is given by Eq. (2.24) or can also be expressed by the ratio of the
vapor pressures, pA/p*A (neglecting interactions in the vapor phase, otherwise the
fugacities have to be employed). As the solution is made more dilute and ap-
proaches the dilute ideal solution and the pure solvent, the activity coefficient
fA approaches unity:
lim f A = 1 (2.26)
xA →1
so that
µ B = µ ∞B + RT ln xB + µ EB = µ ∞B + RT ln xB + RT ln f B (2.29)
The activity coefficients fA and fB may be smaller or larger than unity;
hence, the excess chemical potentials µAE and µBE may be negative or positive.
Depending on the sign of µAE, the solution is said to exhibit positive or negative
deviations from Raoult’s law. Similarly, positive or negative deviations from
Henry’s law can be noted (see Fig. 2.4).
52 Marcus
The considerations that lead to Eq. (2.22) also apply to the case of non-
ideal solutions, except that the terms RT ln fB′ and RT ln fB″ must be added to
the left- and right-hand sides. When rearranged into the form of Eq. (2.23), the
result is
DB = x B′ / x B′′ = PB ( fB′′ / fB′ ) (2.30)
Therefore, the distribution ratio of B remains constant only if the ratio of the
activity coefficients is independent of the total concentration of B in the sys-
tem, which holds approximately in dilute solutions. Thus, although solutions
of metal chelates in water or nonpolar organic solvents may be quite nonideal,
Nernst’s law may still be practically obeyed for them if their concentrations
are very low (xchelate < 10−4). Deviations from Nernst’s law (constant DB) will in
general take place in moderately concentrated solutions, which are of particular
importance for industrial solvent extraction (see Chapter 12).
When no distinction between the solvent and the solute in a liquid mix-
ture is made, then nonideal mixtures can still be described by means of an
expression similar to Eq. (2.16), but with the addition of a term that is the
excess molar Gibbs energy of mixing, ∆GAB E
(see also section 2.2). Thus the
Gibbs energy per mole of mixture is:
G = x B GA* + x B GB* + RT [ x A ln x A + x B ln x B ] + ∆GAB
E
(2.31)
where ∆G = xAµ + xBµ = RT [xA ln fA + xB ln fB] is a function of the compo-
E
AB
E
A
E
B
sition. It is this quantity that has been used in Eq. (2.5) to define the variable
bAB(T), which describes the tendency of the mixture to separate into two liquid
phases. The excess Gibbs energy is positive (>2RT) for mixtures that tend to
separate, but is negative when the components have a strong mutual attraction
of their molecules. For some mixtures, ∆G AB E
is positive over a part of the
composition range and negative over another part.
When a partial derivative of ∆G AB E
relative to the number of moles of a
component, say, A, is taken, the result is RT times the (natural) logarithm of
the activity coefficient of this component, RT lnf′A. When the partial derivative
relative to the temperature is taken, the negative of the excess entropy of mix-
ing, −∆S AB
E
, results. For a regular solution (see section 2.2) this quantity is zero,
so that the (excess) heat (enthalpy) of mixing is ∆H AB E
= ∆G AB
E
. The qualifier
“excess” for the enthalpy of mixing is put in parentheses, since ideal mixtures
have no enthalpies of mixing, so that if heat is evolved or absorbed when two
liquids are mixed, the mixture is not ideal and the observed heat (enthalpy) of
mixing is always an excess quantity. The enthalpy of mixing, ∆H AB E
, may or
may not be temperature dependent itself, and may be positive, negative, or
may change sign within the composition range. A negative ∆H AB E
is, again,
indicative of strong mutual interactions of the molecules of the components.
An example of such behavior is met with in aqueous ethanol (EtOH) at,
say, 75°C. At xEtOH < 0.24, the heat of mixing is negative (heat is evolved on
mixing the components), but at higher ethanol contents it is positive (heat
is absorbed and the mixture cools). In the equimolar mixture ∆H AB E
= 220
−1
but ∆G AB = 960 J mol , due to a negative entropy of mixing, but since
E
Principles of Solubility and Solutions 53
2xEtOHxwaterRT = 1446 J mol−1 for this mixture, a single liquid phase is stable.
However, for aqueous 1-butanol (BuOH) at 25°C, ∆G AB E
≥ 2xBuOHxwaterRT =
1220 J mol for the near equimolar mixture, although ∆H AB = 500 J mol−1 only,
−1 E
so the mixture is on the verge of splitting into two liquid phases (it splits when
xBuOH = 0.485).
Fig. 2.5 Concentration scales. The molality mB [in mol/(kg solvent)] and molarity cB
[in mol (L solution)−1] of solute B, which has a molar mass MB 0.100 kg mol−1 and a
molar volume VB = 0.050 L mol−1, in the solvent A, which has MA = 0.018 kg mol−1 and
VA = 0.018 L mol−1 (water), are shown as a function of the mole fraction xB of the solute.
Note that the molarity tends toward a maximal value (1/VB), whereas the molality tends
toward infinity as xB increases toward unity.
where, as before, µ∞B(c) = lim(cB→0) (µB − RT ln cB) is the standard chemical poten-
tial of B on the molar scale, lim(cB→0) yB = 1.
As an example for such concentration scale conversions, consider 20 wt%
tri-n-butyl phosphate (TBP, B) in toluene (A). The mole fraction of the TBP
is xB = (20/MB)/[(20/MB) + (80/MA)] = 0.080, its molality is mB = (0.200/MB)/
0.800 = 0.94 m, and its molarity is cB = 0.94 ρA/(1 + 0.94 MB) = 0.65 M (MB is
taken in kg mol−1).
The quantity µB is independent of the concentration scale used, being a
true property of the solution, but the three standard chemical potentials µB∞(x),
µB∞(m), and µB∞(c) are not equal. Consequently, differences between the standard
chemical potentials of a solute in the two liquid phases employed in solvent
extraction also depend on the concentration scale used. Thus, PB defined in
Eq. (2.23) is specific for the rational concentration scale, and does not equal
corresponding quantities pertaining to the other scales. Therefore, Eq. (2.30)
might be rewritten with a subscript (x), to designate the rational scale, [i.e., with
DB(x) and PB(x)]. Similar expressions would then be
DB(m) = mB′ / mB′′ = PB(m) (γB′′/γB′) (2.36a )
for the molal scale, and
DB(c ) = cB′ / cB′′ = PB(c ) ( yB′′/yB′) (2.36b)
for the molar scale. The subscripts are often omitted, if the context unequivo-
cally defines the concentration scale employed.
∑ c+ z+ + ∑ c− z− = 0 (2.37)
(Note that z− is a negative number, so that ν+ z+ + ν− z− = 0.) The electroneutral-
ity principle is equivalent to stating that it is impossible to produce a solution
that contains, for example, only cations or an excess of positive charge (i.e.,
that ions cannot be considered as independent components of solutions). Only
entire electrolytes are components that can be added to a solution.
It is expedient to consider the mean ionic properties of the electrolytes.
An electrolyte that dissociates into ν ions has a mean molality m± = (ν+ m+ +
ν− m−)/ν and a mean molarity c± = (ν+ c+ + ν− c−)/ν.
A useful concept that is used when the activities of electrolytes are calcu-
lated is that of the ionic strength of the solution. This is defined (on the molar
scale) as:
1
I= ∑ ci zi2 (2.38)
2
where the summation extends over all the cations and anions present in the
solution. Since the charges appear to the second power in the expression for the
ionic strength, all the terms are positive. An electrolyte that is completely disso-
ciated to univalent ions is designated as a 1:1 electrolyte, examples being aque-
ous HCl and LiNO3. If only a single 1:1 electrolyte is present at a molar concen-
tration cCA, then I = cCA. According to this notation, MgCl2, and (NH4)2SO4 are
2:1 and 1:2 electrolytes, respectively, for which I = 3 cCA. For 1:1, 2:2, . . . (i.e.,
symmetric) electrolytes, m± = mCA, and I = cCA, 4 cCA, . . . The relationship be-
tween mCA and cCA is given by Eq. (2.34.)
The mole fraction concentration scale is generally used for the solvent water,
designated by subscript w and having molar mass Mw.
The activity of water is related formally to the molality of the electrolyte
by means of the osmotic coefficient, ϕ, of the solution:
ln aW = ln ( pW / pW* ) = − v m± M Wϕ (2.41)
As the solution becomes more dilute in the electrolyte (or, in all electrolytes
present, where ν m± is replaced by Σ νi m±i), ln aw approaches zero (aw ap-
proaches unity), and ϕ approaches unity.
A definite relationship exists between γ± and ϕ (the Gibbs–Duhem rela-
tionship), which may be expressed by the excess Gibbs energy of the solution
of an electrolyte:
∆G E = v m± RT [1 − ϕ + ln γ ± ] (2.42)
This expression is analogous to Eq. (2.3), in that (1 − ϕ) expresses the contri-
bution of the solvent and ln γ± that of the electrolyte to the excess Gibbs energy
of the solution. The calculation of the mean ionic activity coefficient of an
electrolyte in solution is required for its activity and the effects of the latter in
solvent extraction systems to be estimated. The osmotic coefficient or the activ-
ity of the water is also an important quantity related to the ability of the solu-
tion to dissolve other electrolytes and nonelectrolytes.
Electrostatic and statistical mechanics theories were used by Debye and
Hückel to deduce an expression for the mean ionic activity (and osmotic) coeffi-
cient of a dilute electrolyte solution. Empirical extensions have subsequently
been applied to the Debye–Hückel approximation so that the expression remains
approximately valid up to molal concentrations of 0.5 m (actually, to ionic
strengths of about 0.5 mol L−1). The expression that is often used for a solution
of a single aqueous 1:1, 2:1, or 1:2 electrolyte is
log γ ± = − Az+ z−I 1 / 2 /[1 + 1.5 I 1 / 2 ] + bI (2.43)
A is a constant that depends on the temperature, and at 25°C A = 0.509, and b
has the value 0.2. (Note that although I is given on the molar scale, the γ±
obtained from the semi-empirical Eq. (2.43) is on the molal scale.) However,
Eq. (2.43) does not seem to be valid for 2:2, 3:1, or higher electrolytes (i.e.,
those with highly charged ions) at practical concentrations. A much more com-
plicated expression results from the Debye–Hückel theory for the osmotic coef-
ficient of the solution, but in the range of validity of Eq. (2.43), ϕ differs from
unity by no more that about −5%. Other modifications of the theory have been
suggested to extend its validity beyond the I = 0.5 mol L−1 of Eq. (2.43), but
their consideration [3,8] is outside the scope of this book. If I for the solution
is >0.5 mol L−1, and for higher-charge-type electrolytes (2:2, 3:1, etc.), experi-
mental values of γ± and ϕ can be obtained from suitable compilations [8].
58 Marcus
if the relative permittivity of the solvent (see Table 2.1 for values) is roughly
ε > 40, electrolytes are, more or less, completely dissociated into ions. On the
other hand, if ε < 10, only slight or practically no ionic dissociation takes place.
Furthermore, solutions in organic or mixed aqueous–organic solvents lack the
three-dimensional cooperative hydrogen bond network that characterizes aque-
ous solutions. Many “anomalous” properties of aqueous solutions that depend
on this structured nature of water become normal properties in organic solvents.
The superscript zero designates the standard state of infinite dilution, where
only solute–solvent but no solute–solute interactions take place. As mentioned
in section 2.3, tables of ∆tG°(ion,aq → org) [5–7] may be used to obtain the
values for ions, which can then be combined to give the value of ∆tG°(CA,aq
→ org), for the desired electrolyte. For example, for the solvent org = nitro-
benzene (PhNO2), ∆tG°(Na+,H2O → PhNO2) = 36 kJ mol−1, ∆tG°(Cs+, H2O →
PhNO2) = 18 kJ mol−1, and ∆tG°(I3−, H2O → PhNO2) = −23 kJ mol−1. Therefore,
∆tG° = 13 kJ mol−1 for NaI3 and −5 kJ mol−1 for CsI3 transferring from water to
nitrobenzene, signifying the preference of the former salt for water (positive
∆tG° and of the latter salt for nitrobenzene (negative ∆tG°).
Principles of Solubility and Solutions 61
charges of the cation and the anion and inversely proportional to the absolute
temperature, T, the relative permittivity (dielectric constant) of the solvent, ε,
and to the distance a. At 25°C
log b = log z + z − + 1.746 − log ε − log a(in nm) (2.49)
b −4 t
The function Q(b) = 2 ∫ t e dt (t is an auxiliary variable) mentioned earlier can
be approximated by eb/b4 at high values of b, or be read from a figure [3]. The
value of Kass at 25°C is given by
log K ass (in L mol -1 ) = 3 log z + z − + 6.120 − 3 log ε + log Q(b) (2.50)
Finally, values of α can be obtained from Kass by means of Eq. (2.48), provided
that y± is calculated iteratively [use Eqs. (2.34), (2.38), and (2.43)].
Solvent-separated ion pairs, in which the first solvation shells of both ions
remain intact on pairing may be distinguished from solvent-shared ion pairs,
where only one solvent molecule separates the cation and the anion, and contact
ion pairs, where no solvent separates them (Fig. 2.6). The parameter a reflects
the minimum distance by which the oppositely charged ions can approach each
other. This equals the sum of the radii of the bare cation and anion plus 2, 1,
and 0 diameters of the solvent, respectively, for the three categories of ion pairs.
Since a appears in Eq. (2.49), and hence, also in Q(b), it affects the value of
the equilibrium constant, Kass. The other important variable that affects Kass is
the product Tε and, at a given temperature, the value of the relative permittivity,
ε. The lower it is, the larger b is and, hence, also Kass.
When the relative permittivity of the organic solvent or solvent mixture is
ε < 10, then ionic dissociation can generally be entirely neglected, and potential
electrolytes behave as if they were nonelectrolytes. This is most clearly demon-
strated experimentally by the negligible electrical conductivity of the solution,
which is about as small as that of the pure organic solvent. The interactions
between solute and solvent in such solutions have been discussed in section 2.3,
and the concern here is with solute–solute interactions only. These take place
mainly by dipole–dipole interactions, hydrogen bonding, or adduct formation.
64 Marcus
phate esters. Cyclic dimers are formed (see Fig. 2.1d), bonded by two hydrogen
bridges, the energy of each being about 20 kJ mol−1. Such cyclic dimers are
quite stable. If the numbers of ionizable hydrogen atoms and electronegative
accepting sites do not match, then noncyclic aggregates, possibly larger than
dimers, are formed by hydrogen bonding. One aspect of such an aggregation is
a drastic increase in the viscosity of the solution with increasing concentra-
tions.
The mean aggregation number of a solute at molality m, forming various
oligomers (i-mers, i = 1, 2, 3, . . . ), is
n% = m / ∑i mi = ∑i imi / ∑i mi (2.51)
the molality of each oligomeric species being weighted by the number of mono-
mers it contains. The equilibrium constants for the aggregation reactions can be
obtained from the dependence of ñ on m by the methods discussed elsewhere
in this book (see Chapter 3).
The third kind of solute–solute interaction, donor–acceptor adduct forma-
tion, tends to form 1:1 adducts between the molecules of two different kinds of
solute, rather than to lead to self-aggregation. Adduct formation results when
one partner has a donor atom (i.e., an atom with an exposed pair of unbonded
electrons), such as the nitrogen atom in trioctylamine, and the other has an
acceptor atom (i.e., one with an orbital that can take up such a pair of electrons),
such as antimony in antimony pentachloride or even the hydroxyl hydrogen
atom in octanol. Although adduct formation formally can be said to take place
in dipole–dipole interactions between two different kinds of molecules, this is
rare, since the strong dipoles required naturally participate also in self-aggrega-
tion. But if one kind of molecule is a donor only and the other is an acceptor
only, as occurs when octanol (the oxygen atom of which has two exposed non-
bonded pairs of electrons) in the foregoing example is replaced by antimony
pentachloride, then self-aggregation is precluded, and only mutual adduct for-
mation can take place (see following).
Hydrogen bond formation between dissimilar molecules is an example
of adduct formation, since the hydrogen atom that is bonded to an electronega-
tive atom, such as oxygen or nitrogen, is a typical acceptor atom. The ability
of molecules to donate a hydrogen bond is measured by their Taft–Kamlet
solvatochromic parameter, α, (or αm. for the monomer of self-associating sol-
utes) (see Table 2.3). This is also a measure of their acidity (in the Lewis
sense, see later, or the Brønsted sense, if protic). Acetic acid, for instance,
has α = 1.12, compared with 0.61 for phenol. However, this parameter is not
necessarily correlated with the acid dissociation constant in aqueous solutions.
The ability of molecules to accept a hydrogen bond is measured by the Taft–
Kamlet solvatochromic parameter, β, (or βm for the monomer of self-associat-
ing solutes) (see Table 2.3). This, too, is a measure of their basicity (in the
Lewis sense), also measured by the Gutmann donor number DN (discussed
later). Thus, pyridine has β = 0.64, compared with 0.40 for acetonitrile, but
Rydberg_5063-2_Ch02_R4_02-09-04 11:07:43
again, this measure is not directly correlated to the base dissociation constant
(protonation constant) in aqueous solutions. The tendency for a hydrogen bond
to be formed between two dissimilar solute molecules A and B increases with
the sum of the products of their donating and accepting parameters: αAβB + αBβA.
Adduct formation does not require formation of a hydrogen bond, and
other acceptor atoms (or molecules) are known (e.g., transition metal cations in
general, SbCl5, I2, and so on). They are collectively called Lewis acids, and they
react with electron pair donors, that are collectively called Lewis bases.
The enthalpy change during donor–acceptor adduct formation has been re-
lated by Drago to the sum of two terms: (1) the product of the electrostatic proper-
ties of the acid and the base, EA and EB; and (2) the product of their tendency
toward covalent bonding, CA and CB [10]. For the particular case, where the ac-
ceptor is specified to be SbCl5 (and the inert solvent is 1,2-dichloroethane), the
negative of this enthalpy change (in kcal mol−1, 1 cal = 4.184 J) is the Gutmann
donor number, DN [2,11]. These concepts are further discussed in Chapter 3.
of the saturated solution), the molal (mol kg−1; i.e., per kilogram of the pure
solvent), the mole fraction, the mass fraction (wt%; see Table 2.2), or the vol-
ume fraction scales. Only if the solute is a gas is the pressure of any signifi-
cance. For liquid and solid solutes, however, the temperature is the only variable
that ordinarily needs to be specified.
cant, but have not as yet been explained. However, the qualitative trends are
the same.
Hg0 and Hg(II) (this is strongly complexed by the Cl− ions present). The precau-
tion of taking into account any side reactions must be followed when it is de-
sired to deduce the actual molar solubilities from solubility products; in particu-
lar, in higher-charge-type salts.
Solubility products can be derived indirectly from standard electrode po-
tentials and other thermochemical data, and directly from tabulated standard
Gibbs energies of formation, ∆fG°, of the ions in aqueous solution [12]. Thus,
the use of
∆ soln G° = ν + ∆ f G°(C z + , aq ) + ν − ∆ f G°(A z − , aq ) − ∆ f G°(Cν + A ν - ) (2.58)
and Eq. (2.56) gives the corresponding Ksp value in aqueous solutions.
When solubility products in nonaqueous solvents are desired, tables [5–7]
of the Gibbs energies of transfer of the ions from water to the desired solvent,
org, must be consulted. For any ion
∆ f G°(ion, org) = ∆ f G°(ion, aq ) + ∆ t G°(ion, aq → org) (2.59)
In other words, values of ∆tG°(ion,aq → org), weighted by the stoichiometric
coefficients for both cation and anion, must be added to ∆solnG°(aq) from Eq.
(2.58) to obtain ∆solnG°(org), which in turn gives Ksp and sCA in the solvent org
from Eqs. (2.56) and (2.58) [12].
The lower the relative permittivity of the solvent org, the greater is the
deviation of the activity coefficient of the electrolyte from unity at a given low
ionic strength, and the smaller must Ksp be for the approximation [Eq. (2.57)]
to remain valid. On the other hand, the values of ∆tG°(ion,aq → org) are posi-
tive for the halide and many other anions, and are positive or negative, but
smaller absolutely than the former, for univalent cations. This means that for
1:1 electrolytes, the value of ∆solnG°(org) is more positive than that of ∆solnG°(aq)
and Ksp in the organic solvent org is smaller than that in water. This trend is
generally enhanced with decreasing values of ε of the solvents. Most alkali
metal halides are much less soluble in common organic solvents than in water,
although there are exceptions (some lithium salts and some iodides).
Divalent or higher-valent cations and, in particular, transition metal cat-
ions, are likely to be covalently solvated by solvents that are strong electron
pair donors (have large solvatochromic β values). This solvation often persists
in crystals, so that the salt that is in equilibrium with the saturated solution in
such solvents may not be the anhydrous salt (nor the salt hydrate). Equation
(2.56) omits any consideration of the solvent of crystallization and pertains to
the solventless (anhydrous) salt. For a salt hydrated by n water molecules in the
crystal, the activity of water raised to the nth power must multiply the right-
hand side of Eq. (2.56) for it to remain valid. A similar consideration applies
for salts crystallizing with other kinds of solvent molecules, the activity of the
solvent in the saturated solution replacing that of water. Such situations must be
70 Marcus
avoided when Eq. (2.59) is to be applied for the calculation of solubilities. For
instance, the solubility of lead iodide, PbI2, in water is only 1.65 × 10−3 mol L−1,
but those of the solvates PbI2 ⴢ DMF and PbI2 ⴢ 2DMSO are 0.71 and 0.99 mol
L−1 in N,N-dimethylformamide (DMF) and dimethylsulfoxide (DMSO), respec-
tively.
Fig. 2.7 Ternary phase diagrams with components A, B, and C. (a) A homogeneous
system, the composition at point P being given by the lengths of the lines marked with
Xs; the dashed line shows the changes occurring when solvent A is added (upward) or
removed (downward). (b) A heterogeneous system, splitting into two liquid phases, with
compositions marked with r, q, p, p′, q′, and r′, along the phase boundary; r and r′ denote
the BC miscibility gap; at nominal composition P there are Pp′ moles of a phase of
composition p for every pP moles of one of composition p′, at the two ends of the tie-
line. (c) A heterogeneous system with two solid solutes, B and C, and solvent A; r and
q denote the solubilities in the binaries AC and AB; at composition r′ the solution con-
tains both B and C and is saturated with solid C; at the invariant point P the solution is
saturated with both B and C.
Principles of Solubility and Solutions 73
plus solvent) system, the composition changes along a saturation line, the satu-
rated solution becoming poorer in this solute. The other solute has its own
saturation line. The point at which the two saturation lines meet is, again, an
invariant point of the system, since now there are two solids in equilibrium with
a saturated solution (i.e., three phases in all). The composition of the solution
that is saturated relative to the two solid solutes is thus given uniquely by the
fixed temperature (see Fig. 2.7c).
General solvent extraction practice involves only systems that are unsatu-
rated relative to the solute(s). In such a ternary system, there would be two
almost immiscible liquid phases (one that is generally aqueous) and a solute at
a relatively low concentration that is distributed between them. The single de-
gree of freedom available in such instances (at a given temperature) can be
construed as the free choice of the concentration of the solute in one of the
phases, provided it is below the saturation value (i.e., its solubility in that
phase). Its concentration in the other phase is fixed by the equilibrium condition.
The question arises of whether or not its distribution between the two liquid
phases can be predicted.
Given a nonionic solute that has a relatively low solubility in each of the
two liquids, and given equations that permit estimates of its solubility in each
liquid to be made, the distribution ratio would be approximately the ratio of
these solubilities. The approximation arises from several sources. One is that,
in the ternary (solvent extraction) system, the two liquid phases are not the pure
liquid solvents where the solubilities have been measured or estimated, but
rather, their mutually saturated solutions. The lower the mutual solubility of the
two solvents, the better can the approximation be made. Even at low concentra-
tions, however, the solute may not obey Henry’s law in one or both of the
solvents (i.e., not form a dilute ideal solution with it). It may, for instance,
dimerize or form a regular solution with an appreciable value of b(T) (see sec-
tion 2.2). Such complications become negligible at very low concentrations, but
not necessarily in the saturated solutions.
The logarithm of the distribution constant, log P, (i.e., of the distribution ratios
D extrapolated to infinite dilution) in such 1-octanol/water systems was shown
by Hansch and Leo to be additive relative to the contributions of a number Ni
of certain fragments of the solute molecules and a number Nj of certain struc-
ture factors. Each hydrocarbon fragment contributes some hydrophobicity (its
tendency to avoid water) or lipophilicity (tendency to prefer nonpolar organic
solvents) to the solute, and adds a positive quantity fi to log P; polar fragments
provide negative fi values, and certain structural features, Fj, may change the
hydrophobicity (e.g., by the shielding of fragments at branching points from
water).
log P (1-octanol/water ) = ∑ N i f i + ∑ N j Fj (2.61)
Table 2.5 lists the additive values for representative fragments f and structural
factors F. As an example, the infinite dilution distribution ratio of C2H5
C(O)OC2H5, ethyl propionate, between 1-octanol and water is obtained as fol-
lows. The carboxylate group contributes −1.49 to log P, each ethyl group con-
tributes 0.89 + 0.66 − 0.12 = 1.43 (for the methyl and methylene groups and
Table 2.5 Selected Fragment Factors f and Structure Factors F for the Calculation
of log P(1-octanol/water)
Fragment f Fragment f
the chain structural feature, respectively), and altogether the calculated log P
is −1.49 + 2 × 1.43 − 0.12 = 1.25 or the distribution ratio is 17.8. The experi-
mental value of log P is 1.21 and of the distribution ratio 16.2, and it is seen
to be estimated with sufficient accuracy.
The entries in Table 2.5 pertain to the distribution between water and 1-
octanol; for any other organic solvent other solute f and F values should be
used. However, 1-octanol is a fair representative of organic solvents in general,
and the effect of using some other solvent can be dealt with by the use of a
“solvent effect” correction, which is much smaller than the solute dependency
given by the log P values. An approximation to such solvent effect corrections
may be made by using the solubility parameters, accounting in effect for the
differing amounts of work done against the cohesive energies of the solvents.
Thus, for the distributions of a solute B (also an electrolyte CA) between water
and two solvents A1 and A2 (where 1 might be 1-octanol):
log ( DB(A1 ) / DB(A2 ) ) = VB (δ 2A1 − δ 2A2 ) / RT (2.62)
If DB(A1) can be equated with P calculated from the entries in Table 2.5, then
DB(A2) in any other solvent A2 can be estimated from Eq. (2.62). Equation (2.62)
is actually a combination of four expressions of the form of Eq. (2.8) (see
section 2.2.2), two for water and solvent Al and two for water and solvent A2,
presuming them to be immiscible pairs of liquids. It employs concentrations
on the mole fraction scale, and assumes that the systems behave as regular
solutions (which they hardly do). This eliminates the use of the solubility param-
eter δ of water, which is a troublesome quantity (see Table 2.1). Solvent Al
need not, of course, be 1-octanol for Eq. (2.62) to be employed, and it suggests
the general trends encountered if different solvents are used in solvent extrac-
tion.
As an illustration of the application of Eq. (2.62), consider the distribution of
acetylacetone, having a molar volume of 103 cm3 mol−1, between various sol-
vents and water, compared with the distribution between 1-octanol and water,
for which log P = 1.04. The values of log D calculated from Eq. (2.62) at
infinite dilution of the acetylacetone are 1.29, 0.80, 0.69, 0.51, and 0.14 for 1-
pentanol, chloroform, benzene, carbon tetrachloride, and n-hexane, respec-
tively. The experimental values of log D are 1.13, 0.77, 0.76, 0.50, and −0.05
for these solvents. The order is the same, and even the numerical values are
not very different, considering that solute–solvent interactions other than the
work required for cavity formation have not been taken into account.
organic phases, provided the solutions are sufficiently dilute for solute–solute
interactions to be negligible. In such cases, the energetics of the relevant interac-
tions consist of the work required for the creation of a cavity to accommodate
the solute, and the Gibbs energy for the dispersion force interactions and for the
donation and acceptance of hydrogen bonds. For a given solute distributing
between water (W) and a series of water-immiscible solvents:
log D = AVVsolute ∆δ 2 + Aπ π*solute ∆π* + Aα α solute ∆β + Aββsolute ∆α (2.63)
where the difference factors are ∆δ2 = δsolvent2 − δW2, ∆π* = π*solvent − π*
W, ∆α = αsol-
vent − αW, and ∆β = βsolvent − βW, i.e., the differences between the relevant proper-
ties of the organic solvent and those of water. The first term on the right-hand
side is related to Eqs. (2.10) and (2.62) and the next three terms to Eq. (2.12).
Interestingly, it was established [13] that the coefficients AV, Aπ, Aα, and Aβ are
independent of the solutes and solvents employed—i.e., are “universal”—pro-
vided that the water content of the organic solvent at equilibrium is approxi-
mately xW < 0.13. Such water-saturated solvents can be considered to be still
“dry” and to have properties the same as those of the neat solvents. The values
of the coefficients are AV = 2.14 (for δ2 in J cm−3 and Vsolute = 0.01VX, where VX
is the additive McGowan–Abraham intrinsic volume [2]), Aπ ⬇ 0, Aα = −7.67,
and Aβ = −4.62. The substantial absence of a term with the polarity/polarizability
values π*solute and π*solvent is explained by the fact that the cavity formation term
in Vsolute and ∆δ2 effectively takes care also of the polar interactions. For “wet”
solvents, i.e., those with xW > 0.13 when saturated with water, such as aliphatic
alcohols from butanol to octanol and tri-n-butyl phosphate, Eq. (2.63) is still
valid with practically the same AV, Aπ, and Aβ, but with Aα = 1.04, provided that
δ2, α, and β are taken as those measured (or calculated) for the water-saturated
solvents.
The applicability of Eq. (2.63) was tested for some 180 individual solutes (with
up to 10 carbon atoms) and 25 “dry” solvents by stepwise multivariable linear
regression and for 28 solutes and 9 dry solvents by target factor analysis; essen-
tially the same conclusions and universal coefficients were obtained by both
methods. As an example of the application of Eq. (2.63), the distribution of
succinic acid between water and chloroform (a “dry” solvent with xW = 0.0048)
and tri-n-butyl phosphate (a “wet” solvent with xW = 0.497) may be cited [13].
For the former solvent log D = −1.92 (experimental) vs. −1.98 (calculated)
and for the latter log D = 0.76 (experimental) vs. 0.88 (calculated). Obviously,
succinic acid with two carboxylic groups that strongly donate hydrogen bonds
(assigned α = 1.12 as for acetic acid) prefers the basic (in the Lewis basicity,
hydrogen-bond-accepting sense) tri-n-butyl phosphate (β = 0.82, measured for
the “wet” solvent) over water (β = 0.47) and naturally also chloroform (β =
0.10).
Principles of Solubility and Solutions 77
salt to be extracted into nitrobenzene from water, contrary to NaI3, which cannot
(having a positive ∆tG°).
A common situation is that the electrolyte is completely dissociated in the
aqueous phase and incompletely, or hardly at all, in the organic phase of a
ternary solvent extraction system (cf. Chapter 3), since solvents that are practi-
cally immiscible with water tend to have low values for their relative permittivi-
ties ε. At low solute concentrations, at which nearly ideal mixing is to be ex-
pected for the completely dissociated ions in the aqueous phase and the
undissociated electrolyte in the organic phase (i.e., the activity coefficients in
each phase are approximately unity), the distribution constant is given by
log PCA = log cCA (org ) − ν log cCA (aq) (2.65)
The distribution ratio for the electrolyte, DCA, however, is given by
log DCA = log[cCA (org ) / cCA (aq )] = log PCA − (ν − 1) log cCA (aq) (2.66)
which is seen to be a concentration-dependent quantity. Even if the constant log
PCA can be predicted, log DCA cannot, unless some arbitrary choice is made for
the aqueous concentration.
As examples, take the extraction of nitric acid by tri-n-butyl phosphate or
of hydrochloric acid by diisopropyl ether. For the former, at low acid concentra-
tions, the acid is completely dissociated in the aqueous phase, but is associated
in the organic phase. Therefore, DHNO3 increases with the nitric acid concentra-
tion according to Eq. (2.65), up to such concentrations at which the dissociation
of this acid in the aqueous phases becomes significantly lower than complete,
and then DHNO3 flattens out. For hydrochloric acid, the situation is the opposite:
it is dissociated in the ether phase up to about 3 M aqueous acid, so that DHCl is
constant, but at higher concentrations it associates in the organic phase, and DHCl
starts to increase, again according to Eq. (2.65), when the nonidealities in both
phases are taken into account.
2.9 CONCLUSION
This chapter provides the groundwork of solution chemistry that is relevant to
solvent extraction. Some of the concepts are rather elementary, but are necessary
for the comprehension of the rather complicated relationships encountered when
the solubilities of organic solutes or electrolytes in water or in nonaqueous sol-
vents are considered. They are also relevant in the context of complex and
adduct formation in aqueous solutions, dealt with in Chapter 3 and of the distri-
bution of solutes of diverse kinds between aqueous and immiscible organic
phases dealt with in Chapter 4.
For this purpose it is necessary to become acquainted, at least in a cursory
fashion, with the physical and chemical properties of liquids (section 2.1.1)
Principles of Solubility and Solutions 79
and the forces operating between their molecules (section 2.1.2). Since solvent
extraction depends on the existence of two immiscible liquid phases, solvent
miscibility, i.e., their mutual solubility, is an important issue (section 2.2). The
solvation of a solute that is introduced into a solvent, i.e., its interactions with
the solvent, is described (section 2.3.1) and the thermodynamics thereof are
elaborated (section 2.3.2), the case of electrolytes that dissociate into ions being
given special attention (section 2.3.3).
Solution chemistry depends strongly on thermodynamic relationships that
have to be mastered in order to make full use of most other knowledge concern-
ing solvent extraction. Therefore, a comprehensive section (2.4) is devoted to
this subject, dealing with ideal (section 2.4.1) as well as nonideal (section 2.4.2)
mixtures and solutions. Then again, as solvent extraction in, e.g., hydrometal-
lurgy, deals with electrolytes and ions in aqueous solutions, the relevant thermo-
dynamics of single electrolytes (section 2.5.1) and their mixtures (section 2.5.2)
has to be understood. In the organic phase of solvent extraction systems, compli-
cations, such as ion pairing, set in that are also described (section 2.6.2).
The amount of a solute that can be introduced into a solvent depends on
its solubility, be it a gas (section 2.7.1), a solid nonelectrolyte (section 2.7.2),
or an electrolyte (section 2.7.3). Ternary systems, which are the basic form of
solvent extraction systems (a solute and two immiscible solvents), have their
own characteristic solubility relationships (section 2.8.1).
In conclusion, therefore, it should be perceived that this groundwork of
solution chemistry ultimately leads to the ability to predict at least semi-quanti-
tatively the solubility and two-phase distribution in terms of some simple prop-
erties of the solute and the solvents involved. For this purpose, Eqs. (2.12),
(2.53), (2.61), and (2.63) and the entries in Tables 2.3 and 2.5 should be particu-
larly useful. In the case of inorganic ions, the entries in Table 2.4 are a rough
guide to the relative extractabilities of the ions from aqueous solutions. The
more negative the values of ∆hydG° of the ions, the more difficult their removal
from water becomes, unless complexation (see Chapter 3) compensates for the
ion hydration.
Although theories of solution (this chapter) and formation of extractable
complexes (see Chapters 3 and 4) now are well advanced, predictions of distri-
bution ratios are mainly done by comparison with known similar systems. Sol-
vatochromic parameters, solubility parameters, and donor numbers, as discussed
in Chapters 2–4, are so far mainly empirical factors. Continuous efforts are
made to predict such numbers, often resulting in good values for systems within
limited ranges of conditions. It is likely that these efforts will successively en-
compass greater ranges of conditions for more systems, but much still has to be
done.
Because a 10% change in the distribution ratio, which can be measured
easily and accurately in the distribution range 0.01–100, corresponds to an en-
80 Marcus
ergy change of only 0.2 kJ/mole, distribution ratio measurements offer a method
to investigate low energy reactions in solutions, such as weak solute–solute and
solute–solvent interactions in the organic phase. So far, this technique has been
only slightly exploited for this purpose. It is particularly noteworthy to find how
few thermodynamic studies have been made of the distribution or extraction
constants, as enthalpy and entropy values give a good indication of the driving
force of the extraction and indicate the structure of the molecular species in the
organic solvent.
REFERENCES
1. Riddick, J. A.; Bunger, W. B.; Sasano, T. K. Organic Solvents, Fourth Edition,
John Wiley and Sons, New York, 1986.
2. Marcus, Y. The Properties of Solvents, John Wiley and Sons, Chichester, 1998.
3. Marcus, Y. Introduction to Liquid State Chemistry, John Wiley and Sons, Chiches-
ter, 1977; 204–207.
4. Kamlet, J. M.; Abboud, J.-L.; Abraham, M. H.; Taft, R. W. J. Org. Chem., 1983,
48: 2877.
5. Marcus, Y. Ion Solvation, John Wiley and Sons, Chichester, 1985.
6. Marcus, Y. Ion Properties, M. Dekker, New York, 1997.
7. Marcus, Y. Pure Appl. Chem., 1983, 55:977.
8. Robinson, R. A.; Stokes, R. H. Electrolyte Solutions, Second Revised Ed., Butter-
worth, London, 1970.
9. Harned, H. S.; Robinson, R. A. Multicomponent Electrolyte Solutions, Pergamon
Press, Oxford, 1968.
10. Drago, R. S. Struct. Bonding, 1973; 15:73.
11. Marcus, Y. J. Solution Chem., 1984; 13:599.
12. Wagman, D. D.; et al., NBS Tables of Chemical Thermodynamic Properties, J.
Phys. Chem. Ref. Data, 1982, Vol. 11, Suppl. 2.
13. Marcus, Y. J. Phys. Chem., 1991;95:8886; Marcus, Y. Solvent Extract. Ion Exch.,
1992, 10: 527; Migron, Y.; Marcus, Y.; Dodu, M. Chemomet. Intell. Lab. Syst.,
1994, 22: 191.
3
Complexation of Metal Ions
GREGORY R. CHOPPIN* Florida State University, Tallahassee, Florida,
U.S.A.
*Retired.
81
82 Choppin
background for the material in the chapters that deal with the distribution equi-
libria, kinetics, and practices of solvent extraction systems.
and ML in the solution since β1 ⴢ[L] = 102 10−2 = 1. This relationship is indepen-
84 Choppin
dent of the amount of metal in the solution and the equality [M] = [ML] = 1 is
valid whether the total metal concentration is 10−9 M or is 1M as long as the
free ligand concentration [L] = 0.01M. For example, if the ligand is present only
as L and ML, and if the total metal concentration, [M] + [ML], is 10−9 M, the total
ligand concentration is [L] + ML = 0.01 + 0.5 × 10−9 = 0.01 M. However, if the
total metal is 1.00 M and [M] = [ML], the total ligand is 0.01 + 0.50 = 0.51 M.
Consider a system in which β1 = 102, β2 = 103 and β3 = 2 × 103. The con-
centrations of each complex can be calculated relative to that of the free (un-
complexed) metal for any free ligand concentration. If [L] = 0.1 M,
[ ML] /[ M ] = β1 ⋅[ L] = 102 × 10−1 = 10
[ ML2 ] /[ M ] = β2 ⋅ [ L]2 = 103 × 10−2 = 10
[ ML3 ] /[ M ] = β3 ⋅[ L]3 = 2 × 103 × 10−3 = 2
Thus, abbreviating the total metal concentration Σ[MLn] = [MT], and defining
the mole fraction of each metal species (i.e., of M, ML, ML2, and ML3) as
the ratio of the concentration of that species to the total metal concentration by
KML,i = [MLi]/[MT], then
K M = 1 /(1 + 10 + 10 + 2) = 0.43
K ML = 10 /(1 + 10 + 10 + 2) = 0.43
K ML = 10 /(1 + 10 + 10 + 2) = 0.43
K ML = 2 /(1 + 10 + 10 + 2) = 0.087
Such calculations can be done for a series of free ligand concentrations to
generate a family of formation curves of concentration or mole fraction of metal
complex species as a function of the concentration of the free ligand. Such curves
are shown in Figs. 3.1 and 3.2. These calculations are particularly useful for trace
level values of metal as they require only knowledge of the free ligand concentra-
tions and the βn. Values of stability constants can be found in Refs. [1,2].
Fig. 3.1 Formation curves of the fraction of Cd(II), Xi, in the species Cd(NH3)2i +, where
i = 0 to 4, as a function of the uncomplexed concentration of NH3.
Fig. 3.2 Formation curves of the fraction of Eu(III), Xi, in the species Eu(OAc)3−i
i ,
where i = 0 to 4, as a function of the uncomplexed acetate anion concentration.
the concentrations of the complexing metal and ligand may be so low (e.g., 0.01
M) that if an inert (i.e., noncomplexing) electrolyte is added at higher
concentration (e.g., 1.00 M), it can be assumed that the ionic strength remains
constant during the reaction. However, such a practice, referred to as a constant
ionic medium, is not always possible if the complexation is so weak that rela-
tively high concentrations of metal and ligand are required.
Consider the complexation of a trivalent metal by a dinegative anion. As-
sume M(ClO4)3 is 0.01 M and Na2SO4 is 0.05 M. To obtain an ionic strength of
1.00 M, if no complexation occurs, requires a concentration of NaClO4 calcu-
lated to be:
[ NaClO 4 ] = 1.00 − 1 / 2〈 0.01 × (3) 2 + 3 × 0.01( −1) 2 + 2 × 0.05 × (1) 2
+ 0.05 × (2)2 〉 = 1.00 − 0.21 = 0.79 M
If all the M3+ is complexed,
[ MSO +4 ] = 0.01 M,
[SO 24- ] = 0.05 − 0.1 = 0.4 M
[Na + ] = 2 × 0.05 = 0.10 M,
[ClO-4 ] = 3 × 0.01 = 0.3 M and
[NaClO 4 ] = 0.79
Complexation of Metal Ions 87
*The literature may contain several differing equilibrium constants for the same reaction [1,2] as
measured in different laboratories or by different techniques.
90 Choppin
Fig. 3.3 Correlation of log β1 of fluoride complexation vs. z+/r+ of the metal.
Fig. 3.4 Relationship between the stability constant, β101, for formation of SmL2+ and
the acid constant, pKa, of HL: (1) propionic acid; (2) acetic acid; (3) iodoacetic acid; (4)
chloroacetic acid; (5) benzoic acid; (6) 4-fluorobenzoic acid; (7) 3-fluorobenzoic acid;
and (8) 3-nitrobenzoic acid.
2 Linear ≤0.15
3 Triangular 0.15 − 0.22
4 Tetrahedral 0.22 − 0.41
4 Planar 0.41 − 0.73
6 Octahedral 0.41 − 0.73
8 Cubic >0.73
Fig. 3.8 Geometric structures of tetrahedral NiCl24−, planar Ni(CN)24− and octahedral
Ni(H2O)26+.
Fig. 3.9 Variation of log β101 for complexation by ammonia (NH3) and ethylenediamine
(en) of some first-row divalent transition metal. The electron occupations of the 3d orbit-
als are listed for each metal.
extra stability is due to the Jahn–Teller effect by which d9 systems have excep-
tional stability when only two or four of the six coordinate sites of the metal
96 Choppin
are occupied by the ligand. These sites are in the square plane of the octahedron.
A loss of this extra stabilization is shown for Cu(en)32+ formation when the three
ethylenediamine (a bidentate ligand) molecules occupy all six coordination sites.
As expected, the ligand field stabilization (the difference between log β103 and
the dashed line between Mn2+ and Zn2+) is greater for Ni(en)32+ than for Cu(en)32+.
some hard acid/hard base pairs do not interact strongly and neither do some
of the soft acid/soft base pairs. The exceptions to the general rule have fac-
tors other than inherent acidity and/or basicity, which are more important in
the interaction. Nevertheless, the HSAB principle has proven a very useful
model for a large variety of complexation reactions, partially because of its
simplicity.
Commonly, soft acids have low-lying acceptor levels while soft bases
have high-lying donor orbitals. As a consequence, the bond in the complex
results from sharing the electron pair. Softness is thus associated with a tendency
to covalency and soft species generally have large polarizabilities. In hard
(metal) acids, the acceptor levels are high and in hard bases, the donor levels
are low. The large energy difference is so large that the cation acid cannot share
the electron pair of the base. The lack of such covalent sharing results in a
strongly electrostatic bond.
The principal features of these species can be listed as follows:
1. Hard species: difficult to oxidize (bases) or reduce (acids); low polarizabili-
ties; small radii; higher oxidation states (acids); high pKa (bases); more
positive (acids) or more negative (bases) electronegativities; high charge
densities at acceptor (acid) or donor (base) sites.
2. Soft species: easy to oxidize (bases) or reduce (acids); high polarizability;
large radii; small differences in electronegativities between the acceptor and
donor atoms; low charge densities at acceptor and donor sites; often have
low-lying empty orbitals (bases); often have a number of d electrons (acids).
Following these guidelines, cations of the same metal would be softer for lower
oxidation states, harder for higher ones. In fact, Cu+ and Tl+ are soft while Cu2+
is borderline and Tl3+ is hard. Even though Cs+ has a large radius and low
charge density, the low ionization potential is sufficient to give it hard acid
characteristics. This illustrates that the properties listed may not be possessed
by all hard (or soft) species, but can serve as guides as to what characteristics
can be used to predict the acid-base nature of species.
Table 3.2 Partial List of Hard and Soft Acids and Bases
A. Acids:
Hard
+1 ions H+, Li to Cs
+2 ions Mg to Ba, Fe(II), Co, Mn
+3 ions Fe(III), Cr(III), Ga, In, Sc, all Ln (III), all actinides (III)
+4 ions Ti, Zr, Hf, all actinides (IV)
−y1 ions Cr (VI), VO, MoO3, AnO2, Mn(VII)
Borderline +2 ions Fe, Co, Ni, Cu, Zn, Sn, Pb
+3 ions Sb, Bi, Rh, Ir, Ru, Os, R3C+, C6H+5
Soft BH3
+1 ions Cu, Ag, Au, Hg, CH3Hg, l
+2 ions Cd, Hg, Pd, Pt
B. Bases:
Hard H2O, ROH, NH3, RNH2, N2H4, R2O, R3PO, (RO)3PO,
−1 ions OH, RO, RCO2, NO3, ClO4, F, Cl,
−2 ions O, R(CO2)2, CO3, SO4
−3 ions PO4
Borderline C6H5NH2, C5H5N
−1 ions N3, NO2, Br
−2 ions SO3,
Soft C2H4, C6H6, CO, R3P(RO)3P, R3As, R2S
−1 ions H, CN, SCN, RS, I.
−2 ions S2O3
Fe2+, Co2+, Ni2+, Cu2+, and Zn2+ are borderline, and, therefore, their complexing
trends are less easily predicted. For F−, we find the order of log β1 to be
Cu > Zn ∼ Fe > Mn > Ni > Co
whereas for thiocyanate, SCN−, it is
Cu >> Ni > Co > Fe > Zn > Mn
For borderline metals and/or borderline ligands, the order is determined by other
factors such as dehydration (removal of hydrate water), steric effects, etc.
Table 3.3 shows some stability constants for hard and soft types of metal
ligand complexes. The constants are experimental values [1,2] and, as such,
contain some uncertainties. Nevertheless they support well the HSAB principle.
The HSAB principle is useful in solvent extraction, as it provides a guide
to choosing ligands that react strongly (high log β) to give extractable com-
plexes. For example, the actinide elements would be expected to, and do, com-
plex strongly with the β-diketonates, R3C−CO−CH2−CO−CR3 (where R = H
100 Choppin
Table 3.3 Stability Constants for Hard and Soft Types of Metal Complexes
and Halide Ions
Hard type ion: In3+ Soft type ion: Hg2+
Liganda log K1 log K2 log K3 log K4 log K1 log K2 log K3 log K4
−
F (hardest) 3.70 2.56 2.34 1.10 1.03
Cl− ↓ 2.20 1.36 6.74 6.48 0.85 1.00
Br− ↓ 1.93 0.67 9.05 8.28 2.41 1.26
I − (softest) 1.00 1.26 12.87 10.95 3.78 2.23
a
For In3+ the ionic medium is 1M NaClO4 + F−, Cl−, and Br−, and 2M NaClO4 + I−. For Hg2+
the ionic medium is 0.5M Na(X,ClO4).
or an organic group), as the bonding is through the oxygens (hard base sites) of
the enolate isomer anion. The formation of a chelate structure by the metal–
enolate complex results in a relatively strong complex, and if the R groups on
the β-diketonates are hydrophobic, the complex would be expected to have good
solubility in organic solvents. In fact, many β-diketonate complexes show excel-
lent solvent extraction properties.
The solvent extraction of the β-diketonate system can be used to illustrate
the importance of the high coordination numbers of the actinides and lanthan-
ides (see section 3.2) in which hydrophobic adducts are involved. The charge
on a trivalent actinide or lanthanide cation is satisfied by three β-diketonate
ligands (each is −1). However, the three β-diketonate ligands would occupy
only six coordination sites of the metal while the coordination number of these
metal cations can be 8 or 9. Thus, the neutral ML3 species can coordinate other
ligand bases. The alkyl phosphates, (RO)3PO, are neutral hard bases that react
with the ML3 species to form ML3ⴢSn (n = 1 to 3). Tributyl phosphate is widely
used as such a neutral adduct-forming ligand of actinides and provides en-
hanced extraction as the ML3ⴢSn species is more hydrophobic than the hydrated
ML3 and, thus, more soluble in the organic phase. However, the role of coordina-
tion sphere saturation by hydrophobic adducts is not limited to synergic systems.
In the Purex process for processing irradiated nuclear fuel, uranium and pluto-
nium are extracted from nitric acid solution into kerosene as UO2(NO3)2 ⴢ (TBP)2
and Pu(NO3)4 ⴢ (TBP)2. In these compounds the nitrate can be bidentate and so
fills four or fewer of the coordinate sites for the uranium and eight or fewer for
the plutonium. Uranium (U) in UO22+ usually has a maximum coordination num-
ber of 6, while that of Pu4+ can be 9 or even 10. The addition of the two TBP
adduct molecules in each case causes the compound to be soluble and extract-
able in the kerosene solvent.
Complexation of Metal Ions 101
step reactions above for complexation. The net reaction has the thermodynamic
relations:
∆Gr = ∆Gc + ∆Gh = (∆H c + ∆H h ) − T (∆Sc + ∆Sh ) (3.20)
The compensation effect assumes ∆Gh ⬃ 0; the positive values of ∆Hr (=∆Hc +
∆Hh) and ∆Sr (=∆Sc + ∆Sh) mean *∆Hh* > *∆Hc* and *∆Sh* > *∆Sc*. So Eq. (3.20)
can be rewritten:
∆Gr ≈ ∆Gc ≈ ∆Hr − T ⋅ ∆Sr ≈ ∆Hc + T ⋅ ∆Sc (3.21)
The complexation, interpreted in this fashion, implies that:
1. The free energy change of the total complexation reaction, ∆G°, is related
principally to step 2 [Eq. (3.18b)], the combination subreaction;
2. The enthalpy and entropy changes of the total complexation reaction, ∆H°
and ∆S°, reflect, primarily, step 1 [Eq. (3.18a)], the dehydration subreaction.
This discussion is important to solvent extraction systems, as it provides further
insight into the aqueous phase complexation. It also has significance for the
Complexation of Metal Ions 105
3.5 SUMMARY
The important role of thermodynamics in complex formation, ionic medium
effects, hydration, solvation, Lewis acid-base interactions, and chelation has
been presented in this chapter. Knowledge of these factors are of great value in
understanding solvent extraction and designing new and better extraction sys-
tems.
Complexation of Metal Ions 107
REFERENCES
1. Martell, A. E.; Smith, R. M.; Critical Stability Constants, Vol. 1–5, Plenum Press,
New York, (1974–1982).
2. Högfeldt, E. Stability Constants of Metal-Ion Complexes, Part A: Inorganic ligands.
IUPAC Chemical Data Series No. 22, Pergamon Press, New York (1982).
3. Robinson, R. A.; Stokes, R. H.; Electrolyte Solutions, Second Revised Edition, But-
terworth, London (1970).
4. Bjerrum, J.; Metal Amine Formation in Aqueous Solution, P. Haase and Son, Copen-
hagen (1941).
5. Moeller, T.; Inorganic Chemistry, John Wiley and Sons, New York (1952).
6. Lewis, G. N.; Valence and the Structures of Atoms and Molecules, The Chemical
Catalog Co., New York (1923).
7. Pearson, R. G.; Ed. Hard and Soft Bases, Dowden, Huchinson and Ross, East
Stroudsburg, PA, (1973).
8. Gutmann, V.; The Donor-Acceptor Approach to Molecular Interactions, Plenum
Press, New York (1980).
4
Solvent Extraction Equilibria
JAN RYDBERG* Chalmers University of Technology, Göteborg, Sweden
GREGORY R. CHOPPIN* Florida State University, Tallahassee, Florida,
U.S.A.
CLAUDE MUSIKAS* Commissariat à l’Energie Atomique, Paris, France
TATSUYA SEKINE† Science University of Tokyo, Tokyo, Japan
4.1 INTRODUCTION
The ability of a solute (inorganic or organic) to distribute itself between an
aqueous solution and an immiscible organic solvent has long been applied to
separation and purification of solutes either by extraction into the organic phase,
leaving undesirable substances in the aqueous phase; or by extraction of the
undesirable substances into the organic phase, leaving the desirable solute in the
aqueous phase. The properties of the organic solvent, described in Chapter 2,
require that the dissolved species be electrically neutral. Species that prefer the
organic phase (e.g., most organic compounds) are said to be lipophilic (“liking
fat”) or hydrophobic (“disliking water”), while the species that prefer water
(e.g., electrolytes) are said to be hydrophilic (“liking water”), or lipophobic
(“disliking fat”). Because of this, a hydrophilic inorganic solute must be ren-
dered hydrophobic and lipophilic in order to enter the organic phase.
Optimization of separation processes to produce the purest possible prod-
uct at the highest yield and lowest possible cost, and under the most favorable
environmental conditions, requires detailed knowledge about the solute reac-
tions in the aqueous and the organic phases. In Chapter 2 we described physical
factors that govern the solubility of a solute in a solvent phase; and in Chapter
3, we presented the interactions in water between metal cations and anions by
This chapter is a revised and expanded synthesis of Chapters 4 (by Rydberg and Sekine) and 6 (by
Allard, Choppin, Musikas, and Rydberg) of the first edition of this book (1992).
*Retired.
†Deceased.
109
110 Rydberg et al.
which neutral metal complexes are formed. This chapter discusses the equations
that explain the extraction data for inorganic as well as organic complexes in a
quantitative manner; i.e., the measured solute distribution ratio, Dsolute, to the
concentration of the reactants in the two phases. It presents chemical modeling
of solvent extraction processes, particularly for metal complexes, as well as a
description of how such models can be tested and used to obtain equilibrium
constants.
The subject of this chapter is broad and it is possible to discuss only the
simpler—though fundamental—aspects, using examples that are representative.
The goal is to provide the reader with the necessary insight to engage in solvent
extraction research and process development with good hope of success.
like NaClO4. Under such conditions the activity factor ratio of Eq. (4.2) is as-
sumed to be constant, and KD is used as in Eq. (4.1) as conditions are varied at
a constant ionic strength value. In the following derivations, we assume that the
activity factors for the solute in the aqueous and organic solvents are constant.
Effects due to variations of activity factors in the aqueous phase are treated in
Chapter 6, but no such simple treatment is available for species in the organic
phase (see Chapter 2).
The assumption that the activity factor ratio is constant has been found to
be valid over large solute concentration ranges for some solutes even at high
total ionic strengths. For example, the distribution of radioactively labeled GaCl3
between diethyl ether and 6M HCl was found to be constant (KD,Ga ⬇ 18) at all
Ga concentrations between 10−3 and 10−12 M [1].
In the following relations, tables, and figures, the temperature of the sys-
tems is always assumed to be 25°C, if not specified (temperature effects are
discussed in Chapters 3 and 6, and section 4.13.6). We use org to define species
in the organic phase, and no symbol for species in the aqueous phase (see Ap-
pendix C).
Fig. 4.1 Liquid-liquid distribution plots. (a) The distribution ratios D for three different
substances A, B, and C, plotted against the variable Z of the aqueous phase. Z may
represent pH, concentration of extractant in organic phase ([HA]org), free ligand ion con-
centration in the aqueous phase ([A−]), aqueous salt concentration, etc. (b) Same systems
showing percentage extraction %E as a function of Z. D and Z are usually plotted on
logarithmic scale.
range of D is best measured from about 0.1–10, though ranges from about
10−5 –104 can be measured with special techniques (see section 4.15).
In many practical situations, a plot like Fig. 4.1a is less informative than
one of percentage extraction, %E, where:
% E = 100 D /(1 + D) (4.4)
Such a plot is shown in Fig. 4.1b for the same system as in Fig. 4.1a. Percentage
extraction curves are particularly useful for designing separation schemes. A
series of such curves has already been presented in Fig. 1.3.
A convenient way to characterize the S-shaped curves in Figs. 1.3 or 4.1b,
where the extraction depends on the variable Z, is to use the log Z value of 50%
extraction, e.g., log[Cl−]50. The pH50-value indicates −log[H+] for 50% extrac-
tion. This is shown in Fig. 4.1 for distribuends A and B.
Solvent Extraction Equilibria 113
Very efficient separations are often needed in industry, and a single ex-
traction stage may be insufficient. The desired purity, yield, etc. can be achieved
by multiple extractions, as discussed in Chapter 7 (see also section 1.2). In
the design of separation processes using multistage extractions, other extraction
diagrams are preferred. Only single stage extraction is discussed in this chapter,
while multistage extraction is discussed in the second part (Chapters 7–14) of
this book.
4.1, the sign of the free energy change, ∆G0, in each step is given by qualita-
tively known chemical affinities (see Chapter 2). The reaction path is chosen
beginning with the complexation of U(VI) by NO−3 in the aqueous phase to form
the uncharged UO2(NO3)2 complex (Step 1). Although it is known that the free
uranyl ion is surrounded by water of hydration, forming UO2(H2O)2+ 6 , and the
nitrate complex formed has the stoichiometry UO2(H2O)6(NO3)2, water of hydra-
tion is not listed in Eq. (4.5) or Table 4.1, which is common practice, in order
to simplify formula writing. However, in aqueous reactions, water of hydration
can play a significant role. As the reactive oxygen (bold) of tributylphosphate,
OP(OC4H9)3, is more basic than the reactive oxygen of water, TBP, which
slightly dissolves in water (Step 2), replaces water in the UO2(H2O)6(NO3)2 com-
plex to form the adduct complex UO2(TBP)2(NO3)2. This reaction is assumed to
take place in the aqueous phase (Step 3). Adduct formation is one of the most
commonly used reactions in solvent extraction of inorganic as well as organic
compounds. (Note: the term adduct is often used both for the donor molecule
and for its product with the solute.) The next process is the extraction of the
complex (Step 4). Even if the solubility of the adduct former TBP in the aqueous
phase is quite small (i.e., DTBP very large), it is common to assume that the
replacement of hydrate water by the adduct former takes place in the aqueous
phase, as shown in the third step of Table 4.1; further, the solubility of the
adduct UO2(TBP)2(NO3)2 must be much larger in the organic than in the aqueous
phase (i.e., DUO2 (TBP)2(NO3)2 Ⰷ 1), to make the process useful. Other intermediate
reaction paths may be contemplated, but this is of little significance as ∆G0ex
depends only on the starting and final states of the system. The use of such a
thermodynamic representation depends on the knowledge of the ∆G0i values as
they are necessary for valid calculations of the process.
The relation between ∆G0ex and Kex is given by
∆Gexo = Σ∆Gio = −RT ln K ex ( 4.6)
Solvent Extraction Equilibria 115
Omitting water of hydration, the equilibrium constant for the net extraction pro-
cess in Eq. (4.5) is Kex, where
[UO 2 (NO3 )2 (TBP)2 ]org
K ex = ( 4.7)
[UO 22 + ][HNO3 ]2 [TBP]2org
The extraction constant, Kex, can be expressed as the product of several equilib-
rium constants for other assumed equilibria in the net reaction:
−2
K ex = ΠK i = β2, NO3 K DR β2, TBP K DC (4.8)
where β2,NO3 is the complex formation constant of UO2(NO3)2, and β2,TBP the
formation constant of the extractable UO2(NO3)2(TBP)2 complex from from
UO2(NO3)2 and TBP. KDR and KDC are the distribution constants of the un-
charged species, the reagent and the extractable complex, respectively.
Kex determines the efficiency of an extraction process. It depends on the
“internal chemical parameters” of the system, i.e., the chemical reactions and
the concentration of reactants of both phases. The latter determine the numerical
value of the distribution factor for the solute, which for our example is
the large negative value of the fourth step makes the overall reaction ∆G0ex nega-
tive, thus favoring the extraction of the complex. The first step can be measured
by the determination of the dinitrato complex in the aqueous phase. The second
is related to the distribution constant KD,TBP in the solvent system. Also, the
formation constant of the aqueous UO2(NO3)2(TBP)2 can be measured (for ex-
ample by NMR on 31P of TBP in the aqueous phase). Thus, ∆G04 can be derived.
Fig. 4.2 Synergistic extraction: Distribution of U(VI) between 0.01 M HNO3 and mix-
tures of thenoyltrifluoroacetone (TTA) and tributylphosphate (TBP), or tributylphos-
phineoxide (TBPO), at constant total molarity ([TTA]org plus [TBP]org or [TBPO]org =
0.02 M) in cyclohexane. (From Ref. 2.)
Solvent Extraction Equilibria 117
four are occupied in this complex, the uranyl group is coordinativaly unsatu-
rated. At the left vertical axes of Fig. 4.2, the free coordination sites are occu-
pied by water and/or NO−3, only; and the U(VI) complex is poorly extracted, log
DU about −1. When TBP or TBPO (tributylphosphine oxide*) [both indicated
by B] are added while [HTTA] + [B] is kept constant, the DU value increases to
about 60 for TBP and to about 1000 for TBPO. At the peak value, the complex
is assumed to be UO2(TTA)2B1 or 2. The decrease of DU at even higher [B] is due
to the corresponding decrease in [TTA−], so that at the right vertical axes of
Fig. 4.2 no U(VI)—TTA complex is formed. For this particular case, at much
higher nitrate concentrations, the U(VI) is complexed by NO−3 and is extracted
as an adduct complex of the composition UO2(NO3)2 B1–2, as discussed earlier
for Case I.
The primary cause for synergism in solvent extraction is an increase in
hydrophobic character of the extracted metal complex upon addition of the ad-
duct former. Three mechanisms have been proposed to explain the synergism
for metal + cheland† + adduct former. In the first suggested mechanism, the
chelate rings do not coordinately saturate the metal ion, which retains residual
waters in the remaining coordination sites and these waters are replaced by other
adduct-forming molecules. The second involves an opening of one or more of
the chelate rings and occupation by the adduct formers of the vacated metal
coordination sites. The third mechanism involves an expansion of the coordina-
tion sphere of the metal ion upon addition of adduct formers so no replacement
of waters is necessary to accommodate the adduct former. As pointed out before,
it is not possible from the extraction constants to choose between these alterna-
tive mechanisms, but enthalpy and entropy data of the reactions can be used to
provide more definitive arguments.
The HTTA + TBP system can serve to illustrate the main points of ther-
modynamics of synergism. The overall extraction reaction is written as:
M n + + n HTTA(org) + p TBP(org) ← +
→ M(TTA) n (TBP) p (org) + p H ( 4.10a )
We assume that the first step in the extraction equation is complexation in the
aqueous phase
M z + + n TTA − ← M(TTA) nz − n (aq) + n H + ( 4.10 b)
→
*TBPO = (C4H9)3PO, see Appendix D, example 16, at the end of this book.
†Cheland or chelator is the chelating ligand.
118 Rydberg et al.
The adduct formation reaction in the organic phase (the “synergistic reaction”)
is obtained by subtracting Eqs. (4.10b) and (4.10c) from Eq. (4.10a):
M(TTA) n (org) + pTBP(org) ←
→ M(TTA) n (TBP) p (org) (4.10d )
Thermodynamic data for the extraction reactions of Eqs. (4.10a) and (4.10c)
allow calculation of the corresponding values for the synergistic reaction of Eq.
(4.10d). Measurements of the reaction
UO 2 (TTA) 2 (org) + TBP(org) ←
→ UO 2 (TTA) 2 ⋅ TBP(org) (4.11)
at different temperatures gives log K = 5.10, ∆H0 = −9.3 kJ ⴢ mol−1, T∆S o = 20.0
kJ ⴢ mol−1.
In another experiment, it was found for Th(TTA)4
Th(TTA) 4 (org) + TBP(org) ← Th(TTA) 4 ⋅ TBP(org) ( 4.12)
→
−1
the corresponding values: log K = 4.94, ∆H = −14.4 kJ.mol , T∆S = 13.7 kJ.
o o
mol−1.
Both UO2(TTA)2 and Th(TTA)4 have two molecules of hydrate water
when extracted in benzene, and these are released when TBP is added in reac-
tions Eqs. (4.11) and (4.12). The release of water means that two reactant mole-
cules (e.g., UO2(TTA)2 ⴢ 2H2O and TBP) formed three product molecules (e.g.,
UO2(TTA)2 ⴢ TBP and 2H2O). Therefore, ∆S is positive. Since TBP is more
basic than H2O, it forms stronger adduct bonds, and, as a consequence, the
enthalpy is exothermic. Hence, both the enthalpy and entropy changes favor the
reaction, resulting in large values of log K.
(c) for the (mixed) hydroxide MAn(OH)p see Fig. 4.3. It should be noted that the
mixed MAn(OH)p complexes include the MAn + M(OH)p complexes. As mixed
complexes are more difficult to determine, they are less often described. How-
ever, it is important to realize that if metal hydroxy complexes are formed and
not corrected for, the result of the investigation can be misleading. A test of the
system according to Fig. 4.3 rapidly establishes the type of metal complexation.
Because metals differ in size, charge, and electronic structure, no two
metals behave exactly the same in the same solvent extraction system, not even
for the same class of solutes. Nevertheless, there are systematic trends in the
formation and extraction of these complexes, as described in Chapter 3. Here,
the emphasis is on models that give a quantitative description of the extraction
within each type or class.
In the subsequent discussion, the following simplifications are made:
1. The systems behave “ideally,” i.e., the activity factors are assumed to be
unity, unless specifically discussed;
2. The metal extracted is in trace concentration: [M]t Ⰶ [Extractant]t, as this
simplifies the equations;
3. The reactants are at very low concentrations in both phases.
These are great simplifications in comparison with the industrial solvent
extraction systems described in later chapters. Nevertheless, the same basic reac-
tions occur also in the industrial systems, although activity factors must be intro-
duced or other adjustments made to fit the data, and the calculation of free
Fig. 4.3 Extraction curves for various types of metal chelate complexes, when log DM
is plotted against free ligand ion concentration, pA = −log[A−], or against [HA][H+]−1.
From such plots, the general type of metal chelate complex may be identified: (a) type
MAn, (see also Fig. 4.10); (b) type MAn(OH)p(HA)r, (see also Figs. 4.14 and 4.30); (c)
type MAn(OH)p, (see also Fig. 4.19). (From Refs. 3a and 3b.)
124 Rydberg et al.
ligand concentrations are more complex. Some of these simplifications are not
used in later chapters.
Source: Ref. 4.
Solvent Extraction Equilibria 125
Rn is extracted more easily than the smaller Xe, because the work to produce a
cavity in the water structure is larger for the larger molecule. The energy to
produce a cavity in nonpolar solvents is much less, because of the weaker inter-
actions between neighboring solvent molecules. Energy is released when the
solute leaves the aqueous phase, allowing the cavity to be filled by the hydro-
gen-bonded water structure. Thus the distribution constant increases with in-
creasing inertness of the solvent, which is measured by the dielectric constant
(or relative permittivity). The halogens Br2 and I2 show an opposite order due
to some low reactivity of halogens with organic solvents. Very inert solvents
with low permittivity, such as the pure hydrocarbons, extract inert compounds
better than solvents of higher permittivity; conversely, liquids of higher permit-
tivity are better solvents for less inert compounds. Molar volumes should be
used for accurate comparisons; such data are found in Table 2.1 and in Ref. [6].
In benzene, the distribution constant depends on specific interactions be-
tween the solute and the benzene pi-electrons. Table 4.4 shows the importance
of the volume effect for the mercury halide benzene system (Cl<Br<I).
Undissociated fatty acids (HA) behave like inert molecules. Figure 4.4
shows the distribution (DHA = KD,HA) between benzene and 0.1 M NaClO4 of
fatty acids of different alkyl chain lengths (Cn, n = 1 to 5); the distribution con-
stant for an acid with chain length n is given by the expression log KD,HA = −2.6
+ 0.6n. Similar correlations between KD,HA and molecular size or chain length
are observed also for other reagents (e.g., normal alcohols).
For organic solutes, not only the size but also the structure is of impor-
tance. Table 4.5 gives distribution constants for substituted oxines. When the
substitution increases the size of the molecule, the distribution constant in-
creases. The variations within Table 4.5 and position of the substitution in the
oxine molecule reflect structural effects. Table 4.6 shows distribution constants
for β-diketones. The increasing KD with molecular size for the series acetylace-
tone, benzoylacetone, and dibenzoylmethane, reflects the decreasing solubility
in the water phase, mainly governed by the increased energy necessary to over-
come the solute-solvent interactions for the larger extractant molecules in water.
Thenoyltrifluoroactetone has a greater hydrophilic character than the other β-
diketones due to its O− and F− atoms, which interact with the water molecules,
Source: Ref. 5.
126 Rydberg et al.
Fig. 4.4 Distribution constants KD,HA of fatty acids as a function of the number n of
carbon atoms in the alkyl chain (C1 is acetic acid) in the system 0.1 M NaClO4/benzene.
(From Ref. 7.)
R is C CH 2 C R ′ is thenoyl, H 3C 3SC
|| ||
O O
a
Organic phases 0.1 M in solute; 25°C.
Source: Ref. 4.
leading to a reduction in the distribution constant. Thus, either the size effect
related to the water structure or the presence of hydrophilic groups in the solute
determines the general level of its distribution constant.
A(aq) + B(org) ←
→ AB(org) K ex = [ AB]org /[ A ]aq [ B]org
= K D,AB K ad K D,B (4.15d )
and also
DA = K ex [B]org (4.15e)
For the extraction reaction it may suffice to write the reaction of Eq.
(4.15d), though it consists of a number of more or less hypothetical steps. As
mentioned, equilibrium studies of this system cannot define the individual steps,
but supplementary studies by other techniques may reveal the valid ones. Equa-
tion (4.15) indicates that the reaction takes place at the boundary (interface)
between the aqueous and organic phases. However, it is common to assume that
a small amount of B dissolves in the aqueous phase, and the reaction takes place
in the steps
A(aq) + B(aq) → AB(aq) → AB(org)
These equations allow definition of a distribution constant for the species AB,
KD,AB [see Eq. (4.15c)]. Distribution constants can also be defined for each of
the species A, B and AB (KD,A, etc.) but this is of little interest as the concentra-
tion of these species is related through Kex. A large Kex for the system indicates
that large distribution ratios DA can be obtained in practice. As shown in Eq.
(4.15), the concentration of B influences the distribution ratio DA.
Consider first the extraction of hexafluoroacetylacetone (HFA) by TOPO
by Example 1, and, second, the extraction of nitric acid by TBP (Example 7).
The principles of volume and water-structure effects, discussed for the solute A
in section 4.4, are also important in the distribution of the adducts.
Example 1: Extraction of hexafluoroacetylacetone (HFA) by trioctylphospine
oxide (TOPO).
Abbreviating HFA (comp. structure 5e, Appendix D) by HA, and TOPO
by B, we can write the relevant reactions
HA(aq) ←
→ HA(org) K D = [HA]org /[HA] = D0 (4.16a )
←
HA(org) + B(org) → HAB(org) K ad1 = [HAB]org /[HA]org [ B ]org (4.16b)
HA(org) + 2B(org) ← 2
→ HAB2 (org) K ad2 = [HAB2 ]org /[HA]org [ B ]org (4.16c)
assuming that 2 adducts are formed, HAB and HAB2, the latter containing 2
TOPO molecules. Equation (4.16a) denotes the distribution of “uncomplexed
HA” by Do. Combining these equations yields
D ⋅ D0−1 = 1 + Kad1[B] + Kad 2 [B]2 ( 4.16d )
−1
Figure 4.5 shows the relative distribution, log D ⴢ D , of hexafluoroace-
o
tylacetone as a function of the concentration of the adduct former TOPO. HFA
Solvent Extraction Equilibria 129
HA is used for the extraction of a metal, KD,HA is abbreviated KDR, for the distri-
bution constant (of the unmodified) reagent (or extractant).
Figure 4.7a shows the effect of aqueous salt concentrations on the DHA
value of acetylacetone at constant total HA concentration and pH. The salt has
two effects: (1) it ties up H2O molecules in the aqueous phase (forming hydrated
ions) so that less free water is available for solvation of HA; and (2) it breaks
down the hydrogen bond structure of the water, making it easier for HA to
dissolve in the aqueous phase. Figure 4.7 shows that the former effect dominates
for NH4Cl while for NaClO4 the latter dominates. We describe the increase of
the distribution ratio with increasing aqueous salt concentration as a salting-out
effect, and the reverse as a salting-in effect.
Figure 4.7b shows DHA for the extraction of acetylacetone into CHCl3 and
C6H6 for two constant aqueous NaClO4 concentrations at pH 3, but with varying
concentrations of HA. Acetylacetone is infinitely soluble in both CHCl3 and
C6H6; at [HA]org = 9 M, about 90% of the organic phase is acetylacetone (Mw
100), so the figure depicts a case for a changing organic phase. Figure 4.7b also
Solvent Extraction Equilibria 131
Fig. 4.6 Distribution ratios calculated by Eq. (4.22) for acetylacetone (HAA); benzoyl-
acetone (HBA); bezoyltrifluoroacetone (HBTFA); and oxine (8-hydroxyquinoline,
HOQ), in the system 0.1 M NaClO4 /CCl4, using the following constants. (From Refs.
8a, b.)
indicates different interactions between the acetylacetone and the two solvents.
It is assumed that the polar CHCl3 interacts with HA, making it more soluble in
the organic phase; it is also understandable why the distribution of HA decreases
with decreasing concentration (mole fraction) of CHCl3. C6H6 and aromatic sol-
vents do not behave as do most aliphatic solvents: in some cases the aromatics
seem to be inert or even antagonistic to the extracted organic species, while in
other cases their pi-electrons interact in a favorable way with the solute. For
acetylacetone, the interaction seems to be very weak. The salting-in effect is
shown both in Figs. 4.7a. and 4.7b.
4.6.1 Dissociation
Acids dissociate in the aqueous phase with a dissociation constant Ka
132 Rydberg et al.
Solvent Extraction Equilibria 133
HA ←
→ H +A
+ −
K a = [H + ][A − ]/[HA] (4.18)
The distribution ratio incorporates the Ka for extraction of acids, HA, as:
DA = [HA]org ([HA] + [A − ]) −1 = K D ,HA (1 + K a [H + ]−1 ) (4.19)
Index A indicates that the distribution ratio refers to the concentration of all
species of A in the organic and in the aqueous phase. In Fig. 4.6 the distribution
of the β-diketones is constant in the higher hydrogen ion concentration range
(lower pH) where they are undissociated. In the higher pH region, DA becomes
inversely proportional to the hydrogen ion concentration due to increase in the
concentration of the dissociated form of the acid A−, in agreement with Eq.
(4.19.)
The free ligand concentration, [A−], is an important parameter in the for-
mation of metal complexes (see Chapter 3 and section 4.8). In a solvent extrac-
tion system with the volumes V and Vorg of the aqueous and organic phases,
respectively, [A−] is calculated from the material balance:
log [A − ] = log Ka − log[H + ] + log ( mHA, t / Vorg ) − log F ( 4.20a )
where
−1 −1
F = K D ,HA + V Vorg + K a [H + ] V Vorg (4.20b)
mHA,t is the total amount (in moles) of HA (reagent) added to the system. Often
−1 o
mHA,t V org is abbreviated [HA]org, indicating the original concentration of HA in
the organic phase at the beginning of the experiment (when [HA]aq = 0). When
Vorg = V and pH Ⰶ pKa, F = 1 + KD,HA. From Eq. (4.20) it can be deduced that
[A−] increases with increasing pH, but tends to become constant as the pH value
approaches that of the pKa value. In the equations relating to the extraction of
metal complexes, HA is often identical with reagent R; the indexes may be
changed accordingly, thus e.g., KD,HA ≡ KDR. (Note: Various authors use slightly
different nomenclature; here we follow reference Appendix C.)
4.6.2 Protonation
At low pH, some organic acids accept an extra proton to form the H2A+ com-
plex. This leads to a decrease in the DA value at pH < 6, as shown in Fig. 4.6:
䉳
Fig. 4.7 Distribution ratio DHA of undissociated acetylacetone. (a) Distribution between
benzene and aqueous phase containing different inorganic salts; 25°C. (b) Distribution
between CHCl3 (upper curves) or C6H6 (lower curves) and aqueous phase 0.1 and 1.0 M
in NaClO4 as a function of [HA]org. The uncertainty at the lowest D values is ±1 for
CHCl3 and ±0.2 for C6H6. (From Ref. 10.)
134 Rydberg et al.
4.6.3 Dimerization
Figure 2.1 illustrates a number of orientations by which two linear acids may
form a dimer. The partial neutralization of the hydrophilic groups leads to in-
creased solubility of the acid in the organic solvent, but is not observed in the
aqueous phase. The dimerization can be written as:
2HA(org) ←
→ H 2 A 2 (org) K di = [H 2 A 2 ]org /[HA]org
2
(4.23)
The distribution ratio for the extraction of the acid becomes:
DA = ([HA]org + 2[H 2 A 2 ]org )([HA] + [A − ]) −1
= K D ,HA (1 + 2 K di K D ,HA [HA]aq (1 + K a [H]−1 ) −1 (4.24)
The last term can be expressed in several different ways. Because the distribu-
tion ratio DA reflects the analytical concentration of A in the organic phase, the
dimer concentration is given as 2[H2A2], although it is a single species (one
molecule). Figure 4.8 illustrates how the dimerization leads to an increase of
acid distribution ratio with increasing aqueous acid concentration. For propionic
acid logKa = −4.87, logKD,HA = −1.90 and logKdi = 3.14 in the system. The ex-
traction increases as the size of the acid increases. A dimeric acid may form
monobasic complexes with metal ions, as is illustrated by the formulas in Ap-
pendix D:14 b–d, for the M(H(DEHP)2)3 and UO2((DEHP)2)2 complexes. The
situation may be rather complex. For example, at very low concentrations in
inert solvents, dialkylphosphates (RO)2POOH act as a monbasic acid, but at
concentrations >0.05 M they polymerize, while still acting as monobasic acids
(i.e., like a cation exchanger). The degree of dimerization/polymerization de-
pends on the polarity of the solvent [11b].
4.6.4 Hydration
In solvent extraction, the organic phase is always saturated with water, and the
organic extractant may become hydrated. In the extraction of benzoic acid, HBz
(Appendix D:2), it was found that the organic phase contained four different
Solvent Extraction Equilibria 135
Fig. 4.8 Distribution ratios (from bottom to top) of acetic (C2 ●), propionic (C3 O),
butyric (C4 ∆), and valeric (C5 ▲) acids (carbon chain length Cn) between carbon tetra-
chloride and water as a function of the acid concentration in the aqueous phase, [HA]aq.
(From Ref. 11a.)
species: the monomer HBz, the monomer hydrate HBz ⴢ H2O, the dimer H2Bz2,
and the dimer hydrate H2Bz2(H2O)2. Only by considering all these species is it
possible to explain the extraction of some metal complexes with this extractant.
← B + HA)
Acid adduct (and acid and adduct HAB (→
former) in organic phase ↓↑ (↓↑) (↓↑)
Acid HA dissociating and forming HAB →← B + HA →← H + + A−
adduct in aqueous phase
The solubility of organic acids in water is due to the hydrophilic oxo- and
hydroxo-groups of the acid that form hydrogen bonds with water molecules. If
the hydrogen ion of the acid is solvated by a donor organic base, B, in the
136 Rydberg et al.
organic phase, the adduct B–HA is likely to have much greater solubility in the
organic phase.
再
Phosphine compounds R3P
Phosphoryls (RO)3POd < R′(RO)2POe < R2′(RO)POf < R3′POg
Arsenyls R3AsOh
Oxo-compounds Carbonyls RCHO < R2CO (≤R2O < ROH < H2O)i
Sulfuryls (RO)2SO2j < R2SOk2 < (RO)2SOl < R2SOm
Nitrosyls RNOn2 < RNOo
Substitutions causing basicity decrease of oxo compounds
(CH3)2CH— < CH3(CH2)n — < CH3 — < CH3O— < ClCH2 —
a–c
Tertiary, secondary, and primary amines.
d
tri-R phosphate.
e
di-R-R′ phosphonate.
f
R-di-R′ phosphinate.
g
tri-R′ phosphine oxide.
h
arsine oxide.
i
ether and hydroxo compounds.
j
sulfates.
k
sulfones.
l
sulfites.
m
sulfoxides.
n
nitro compounds.
o
nitroso compounds.
Source: Ref. 12.
Since [HA]aq and [B]org are easily measurable quantities, it is common to define
the extraction constant Kex for this model:
K ex = [HABb ]org [HA]−1 [B]org
−b
(4.25b)
2. The organic phase reaction model assumes all reactions take place in the
organic phase. Thus one assumes
HA(org) + bB(org) ←
→ HABb (org) (4.26a )
The equilibrium constant for this reaction is
−1 −b
Kad, bB = [HABb ]org [HA]org [B]org ( 4.26 b)
Rydberg_5063-2_Ch05_R3_02-09-04 11:29:21
where Kad,bB is the (organic phase) adduct formation constant. The distribution
ratio of the acid in this system becomes
DA = ([HA]org + [HABb ]org )([HA] + [A − ]−1
= K D ,HA (1 + ΣK ad,bB [B]borg ) /(1 + K a [H + ]) (4.26c)
Equation (4.25b) becomes identical to Eq. (4.26b) if Kex is replaced by KD,HA
Kad,bB. Equilibrium measurements do not allow a decision between the two reac-
tion paths.
Example 2: Extraction of nitric acid by pure TBP.
Many metals can be extracted from nitrate solutions by TBP. In those
systems it is important to account for the HNO3-TBP interactions. The next set
of equations were derived by [13] and are believed to be valid for the extrac-
tion of HNO3 at various nitrate concentrations into 30% TBP in kerosene. Ab-
breviating HNO3 as HL, and TBP as B, and including hydration for all species
without specification, one derives
1. The formation of an acid monoadduct:
H + + L− + B(org) ←
→ HLB(org) (4.27a )
For simplicity, we write the adduct HLB, instead of HB+L−. The extraction
constant is
K ex1 = [HLB]org [H]−1 [L]−1 [B]org
−1
(4.27 b)
H + + L− ← + −
→HL (4.29a )
K ass = [H + ] [L− ]/[H + L− ] (4.29b)
This reaction only occurs under strong acid conditions, and the equilibrium
constant may be <1.
4. The distribution of nitric acid is then given by
DNO3 = ([HLB]org + 2[(HL) 2 B]org + . . . )([L− ] + [H + L− ])−1 (4.30)
6. The distribution ratio in terms of only [H+] and monomeric [B]org can then
be expressed by
DNO3 = K ass−1 [B]org ( K ex1 + 2 K ex2 [H + ]2 ) (4.33)
Fig. 4.9 Test of the equations in Example 2 for extraction of 0.01–0.5 M nitric acid
with 30% TBP in kerosene at temperatures 20–60oC. (From Ref. 13.)
140 Rydberg et al.
Section 3.2 describes an important class of organic ligands that are able
to complex a metal ion through two or more binding sites of “basic” atoms, like
O, N, or S, to form metal chelates. Table 4.9 presents the various types, their
number of acidic groups, the chelate ring size, and the coordinating atoms. Neu-
tral chelate compounds are illustrated in Appendix D: 5h, 14b and 14d. In Ap-
pendix D:5h the ring size is 6 for the complex between Cu2+ and each of the
two acetylacetonate anions, Aa−, while for the dimeric HDEHP ligand in the
figures in Appendix D:14b and 14c the ring size becomes 7 for the M3+ and
UO2+2 complexes (see the structures in Appendix D). As discussed in Chapter 3,
chelation provides extra stability to the metal complex. The formation and extrac-
tion of metal chelates are discussed extensively also in references [14–16].
In Chapter 3 we described how an uncharged metal complex MAz is
formed from a metal ion Mz+ (central atom) through a stepwise reaction with
the anion A− (ligand ) of a monobasic organic acid, HA, defining a stepwise
formation constant kn, and an overall formation constant βn, where
βn = [MA nz −n ]/[M z + ][A − ]n (4.34)
The MAz complex is lipophilic and dissolves in organic solvents and the distri-
bution constant KDC is defined (index C for complex):
K DC = [MA z ]org /[MA z ] (4.35)
Taking all metal species in the aqueous phase into account, the distribution
of the metal can be written (omitting the index aq for water)
[MA z ]org K DCβz [A − ]z
DM = = (4.36)
Σ [MA nz −n ] Σ β z [ A − ]z
The distribution ratio depends only on the free ligand concentration, which
may be calculated by Eq. (4.20). Most coordinatively saturated neutral metal
complexes behave just like stable organic solutes, because their outer molecular
structure is almost entirely of the hydrocarbon type, and can therefore be ex-
tracted by all solvent classes 2–5 of Chapter 2. The rules for the size of the
distribution constants of these coordinatively saturated neutral metal complexes
are then in principle the same as for the inert organic solutes of section 4.4.
However, such complexes may still be amphilic due to the presence of electro-
negative donor oxygen atoms (of the chelating ligand) in the chelate molecule.
In aqueous solution such complexes then behave like polyethers rather than
hydrocarbons. Narbutt [17] has studied such outer-sphere hydrated complexes
and shown that the dehydration in the transfer of the complex from water to the
organic solvent determines the distribution constant of the complex. This is
further elaborated in Chapter 16.
Example 3: Extraction of Cu(II) by acetylacetone.
Simple β-diketones, like acetylacetone (Appendix D: 5d) can coordinate
in two ways to a metal atom, either in the uncharged keto form (through two
keto oxygens), or in dissociated anionic enol form (through the same oxygens)
as shown in Appendix D: 5c, 5h. It acts as an acid only in the enolic form.
Figure 4.10 shows the extraction of Cu(II) from 1 M NaClO4 into benzene at
various concentrations of the extractant acetylacetone (HA) [18]. Acetylace-
tone reacts with Cu(II) in aqueous solutions to form the complexes CuA+ and
CuA2. Because acetylacetone binds through two oxygens, the neutral complex
CuA2 contains two six-membered chelate rings; thus four coordination posi-
tions are taken up, forming a planar complex (Appendix D: 5f). This complex
is usually considered to be coordinatively saturated, but two additional very
142 Rydberg et al.
weak bonds can be formed perpendicular to the plane; we can neglect them
here.
The distribution of copper, DCu, between the organic phase and water is
then described by
[CuA 2 ]org
DCu = (4.37)
[Cu 2+ ] + [CuA + ] + [CuA 2 ]
One then derives
K DC β2 [A − ]2 K β [A − ]2
DCu = = DC 2 − n (4.38)
1 + β1 [A ] + β2 [A ]
− − 2
Σ βn [A ]
where KDC refers to the distribution constant of the uncharged complex CuA2.
In Eq. (4.37), log D is a function of [A−], the free ligand concentration,
only, and some constants. In Fig. 4.10, log Dcu is plotted vs. log [H+] (= −pH).
Through Eqs. (4.17) and (4.18) it can be shown that log DCu is a function of
pH only at constant [HA]org (or [HA]aq), while at constant pH the log DCu de-
pends only on [HA]org (or [HA]aq).
Solvent Extraction Equilibria 143
The horizontal asymptote equals the distribution constant of CuA2, i.e., KDC.
From the curvature between the two asymptotes, the stability constants β1 and
β2 can be calculated.
This example indicates that in solvent extraction of metal complexes with
acidic ligands, it can be more advantageous to plot log D vs. log[A−], rather
than against pH, which is the more common (and easy) technique.
In order to calculate DM from Eq. (4.36), several equilibrium constants as
well as the concentration of free A− are needed. Though many reference works
report stability constants [19, 20] and distribution constants [4, 21], for practical
purposes it is simpler to use the extraction constant Kex for the reaction
M z+ (aq) + zHA(org) ←
→ MA z (org) + zH (aq)
+
(4.41a )
z-n
in which case the MA complexes in the aqueous phase are neglected. The
n
relevant extraction equations are
K ex = [MA z ]org [H + ]z [M z+ ]−1 [HA]org
−z
(4.41b)
and
DM = K ex [HA]org
z
[H + ]− z (4.41c)
Thus only one constant, Kex, is needed to predict the metal extraction for given
concentrations [H+] and [HA]org. Tables of Kex values are found in the literature
(see references given).
Equation (4.41) is valid only when the complexes MAnz−n can be neglected
in the aqueous phase. Comparing Eqs. (4.37b) and (4.41c), it is seen that no
horizontal asymptote is obtained even at high concentrations of A−, or HA and
H. Thus, for very large distribution constant of the uncharged complex (i.e.,
Ⰷ1000) a straight line with slope −z is experimentally observed, as in the case
for the Cu(II)-thenoyltrifluoroacetone (HTTA) system (Appendix D: 5g).
144 Rydberg et al.
Fig. 4.11 Distribution of Cu(II) between hexone (∆), carbon tetrachloride (●), or chlo-
roform (䊊) and 0.1 M NaClO4 in the presence of isopropyltropolone (IPT) or thenoyl-
trifluoroacetone (TTA); (a) as a function of −log[H+] at constant [TTA]org = 0.1 M; (b)
as a function of [TTA]org at constant [H+] = 0.1M. (From Ref. 22.)
Solvent Extraction Equilibria 145
Cu 2+ + 2HA(org) ←
→ CuA 2 (org) + 2 H
+
(4.43a )
The extraction constant is
K ex = [CuA 2 ]org [H + ]2 /[Cu 2+ ] [HA]org
2
(4.43b)
Table 4.10 Distribution Constants for Acetylacetone (HA) and Some Metal
Acetylacetonates Between Various Organic Solvents and 1 M NaClO4 at 25°C
log KDC
ε log KDR
Organic solventa Solvent HA ZnA2 CuA2 NpA4
Fig. 4.12 Enhancement of Zn(II) extraction, D D−o1, from 1 M NaClO4 into carbon
tetrachloride containing the complexing extractants acetylacetone (䊊), trifluoroacetone
(∆), or hexafluoroacetone (䊐) as a function of the concentration of the adduct former
trioctyl phosphine oxide (B). The curves are fitted with Eq. (4.50) using the constants
log Kad1 = 3.07 (AA), 6.70 (TFA), 7.0 (TFA), and Kad2 = 4.66 (AA), nil (TFA), 11.6
(HFA). (From Ref. 24.)
148 Rydberg et al.
TBP and TOPO (B) [12]. The extracted neutral complex is ZnA2Bb. The distri-
bution ratio becomes
[ ZnA 2 ]org + [ ZnA 2 B]org + [ ZnA 2 B2 ]org + L
DZn = (4.48a )
[ Zn] + [ ZnA] + [ ZnA 2 ] + L
To analyze these systems, the overall extraction reaction must be broken into
its partial reactions, or by introducing Eq. (4.47), to obtain
K DCβ2 [A]2 (1 + K ad,1[B]org + K ad,2 [B]org
2
+ L)
DZn = (4.48b)
Σ βn [A]n
where the adduct formation constant is defined by
K ad,b = [ ZnA 2 Bb ]org [ZnA 2 ]org
−1 −b
[B]org (4.49)
In the absence of any adduct former, DZn is given as a function of the free
ligand concentration by Eq. (4.36), i.e.,0 the parentheses in Eq. (4.48b) equals
1; denoting this DZn-value as Do, and introducing it into Eq. (4.48) gives
DZn′ = Do (1 + K ad ,1[B]org + K ad,2 [B]org
2
+ L) (4.50)
DZn′ (instead of DZ′ n) indicates that this expression is valid only at constant
[A−], or, better, constant [H+] and [HA]org [see Eqs. (4.36) and (4.41c)]. In Fig.
4.12, log DZ′ n D−o1 is plotted as a function of log [B]org. The distribution ratio
proceeds from almost zero, when almost no adduct is formed, towards a limit-
ing slope of 2, indicating that the extracted complex has added two molecules
of B to form ZnA2B2. From the curvature and slope the Kad,b-values were deter-
mined (see section 4.10). The calculation of the equilibrium constants is further
discussed under Example 13.
Tables 4.11–4.13 presents adduct formation constants according to Eq.
(4.47). For the alkaline earths TTA complexes in carbon tetrachloride in Table
4.11, the TBP molecules bond perpendicular to the square plane of the two TTA
rings, producing an octahedral complex. The higher the charge density of the
Source: Ref. 4.
Solvent Extraction Equilibria 149
Source: Ref. 4.
central atom, the stronger is the adduct complex with TBP. Note: Charge density
refers to the electrostatic interaction between ions of opposite charge according
to the Born equation [see (2.13), and (3.13)] (based on the Coulomb interac-
tion). It is mostly given as the ratio between the ionic charge, z+ (or z+z−) and
the ionic radius r+ (or r+r−). Table 4.12 compares the adduct formation of the
europium β-diketone complexes with TBP in chloroform. In Eu(TTA)3 the
TTAs only occupy six of the eight coordination positions available; the two
empty positions have been shown to be occupied by water. Though the tendency
is not strong, the stronger the acid (i.e., the larger its electronegativity), the
larger is the adduct formation constant. Table 4.13 compares the adduct forma-
tion tendency of the Eu(TTA)3 complex with various adductants. The basicity
of the donor oxygen atom increases in the order as shown in Table 4.13, as does
the adduct formation constant; this is in agreement with the order of basicity in
Table 4.8 and the donor numbers of Table 3.3; see also section 4.2.
Source: Ref. 4.
150 Rydberg et al.
The difference between the two solvent systems is likely a result of CHCl3
solvating the Eu complex to some extent, while CCl4 is inert. This has two
effects: the KDC value increases due to the solvation by CHCl3 (not shown in the
table), while the adduct formation constant Kad decreases as the solvation hinders
the adduct formation; the more inert solvent CCl4 causes an opposite effect, a
lower KDC and a larger Kad.
Fig. 4.13 Distribution ratio of Zn(II) when extracted from 1 M Na(SCN−,ClO−4 ) into
0.001 M TOPO in hexane, as a function of aqueous SCN− concentration. The following
equilibrium constants were obtained with Eq. (4.53): β1 3.7, β2 21, β3 15, Kex 2.5 107 for
b = 2. (From Ref. 26.)
ZnL 2 + b B(org) ←
→ ZnL 2 Bb (org) (4.51)
for which we may define an equilibrium constant Kex,bB.
Because more than one solvated species may be extracted, the distribution ratio
becomes
[ ZnL 2 B]org + [ ZnL 2 B2 ]org + L
DZn = (4.52)
[ Zn] + [ ZnL] + [ ZnL 2 ] + [ ZnL3 ] + L
Inserting the partial equilibrium constants [see Eq. (4.46)],
β2 [L]2 Σ K ex,bB [B]borg
DZn = (4.53)
1 + Σ βn [L]n
where both summations are taken from 1. The solvation number b can be deter-
mined from the dependence of D on [B]org while [L] is kept constant. From the
slope of the line in Fig. 4.13 at low SCN − concentrations, it follows that b =
2; thus only one adduct complex is identified: Zn(SCN)2(TOPO)2. The authors
were able to calculate the formation constants βn from the deviation of the
curve from the straight the line at constant [B]org, assuming b constant. With
152 Rydberg et al.
these equilibrium constants, the line through the points was calculated with Eq.
(4.53).
Fig. 4.14 Extraction of Pm(III) by acetylacetone (HAa) from 1 M NaClO4 into ben-
zene at three different original concentrations of HAa in the organic phase. pA =
−log[Aa−] is calculated according to Eq. (4.20). The analysis of the system yielded the
constants log β1 5.35, log β2 9.20, log β3 13.22, log β4 14.06, Kad1 7, Kad2 3, and KDC
0.008, shown for Pm in Fig. 4.15. (From Refs. 27a,b.)
Kad,b is the equilibrium constant for the self-adduct formation in the organic
phase, i.e.,
PmA 3 B(org) + b HA(org) ←
→ PmA 3 (HA) b (org) (4.57a )
K ad,b = [PmA 3 (HA) b ]org [PmA ]−1
3 org
−b
[HA]
org
(4.57 b)
For pedagogic reasons we rewrite Eq. (4.55b)
154 Rydberg et al.
βn = Π K n (4.59)
according to Eq. (3.5).
The maximum distribution ratio for the Pm-HA system (DPm about 0.1)
is reached in the pH range 6–7. It is well known that the lanthanides hydrolyze
in this pH region, but it can be shown that the concentration of hydrolyzed
species is <1% of the concentration of PmAn species for the conditions of Fig.
4.14, and can thus be neglected.
Self-adducts are rather common, and have been identified for complexes of Ca,
Sr, Ba, Ni, Co, Zn, Cd, Sc, Ln, and U(VI) with acetylacetone, thenoyltrifluoro-
acetone, tropolone, and oxine. Table 4.14 lists some self-adduct constants. The
fact that the constants vary with the organic solvent indicates that the self-adduct
䉴
Fig. 4.15 The system La(III) acetylacetone (HA) − 1M NaClO4/benzene at 25°C as a
function of lanthanide atomic number Z. (a) The distribution ratio DLn (stars, right axis)
at [A−] = 10−3 and [HA]org = 0.1 M, and extraction constants Kex (crosses, left axis) for
the reaction Ln3+ + 4HA(org) X LnA3HA(org) + 3H+. (b) The formation constants, Kn,
for formation of LnA3n−n lanthanide acetylacetonate complexes (a break at 64Gd is indi-
cated); circles n = 1; crosses n = 2; triangles n = 3; squares n = 4. (c) The self-adduct
formation constants, Kad, for the reaction of LnA3(org) + HA(org) X LnA3HA(org) for
org = benzene. (A second adduct, LnA3(HA)2, also seems to form for the lightest Ln
ions.) (d) The distribution constant KDC for hydrated lanthanum triacetylacetonates, LnA3
(H2O)2−3, between benzene and 1M NaClO4. (From Ref. 28.)
Rydberg_5063-2_Ch05_R3_02-09-04 11:29:22
156 Rydberg et al.
Source: Ref. 4.
reaction probably occurs in the organic phase. Table 4.14 contains both very
inert (e.g., CCl4), polar (CHCl3), and pi-bonding solvents (C6H6).
RN reacts with the acid HL to form RNH + L−, but extracts with an excess
amount of acid HL (over the 1:1 HL:RN ratio) into the organic solvent, and
also with additional water. The practical concentration of amine in the organic
solvent is usually less than 20%; at higher concentrations the amine salt solu-
tions become rather viscous. The amine salts dissociate in highly polar solvents,
while in more inert diluents they easily polymerize to form micelles, and at
higher concentrations a third phase. For example, in xylene (TLA ⴢ HBr)n aggre-
gates with n = 2, 3, and 30 have been identified, and in other systems aggrega-
tion numbers above 100 have been reported. TLA ⴢ HNO3 is mainly trimeric in
m-xylene at concentrations 0.002–0.2 M, but larger aggregates are formed at
higher amine concentrations. These aggregates seem to behave like monofunc-
tional species, each extracting only one anionic metal complex. The aggregation
can be reduced, and the third phase formation avoided, by using aromatic diluents
and/or by adding a modifier, usually another strong Lewis base (e.g., octanol or
TBP). Such additions often lead to considerable reduction in the Kex value.
Four types of organic amines exist, as shown in Table 4.8: primary amines
RNH +3, secondary R2NH +2, tertiary R3NH +, and quaternary R4N +(Appendix D).
The hydrocarbon chain R is usually of length C8−C12, commonly a straight
aliphatic chain, but branched chains and aromatic parts also occur. In general
the amines extract metal complexes in the order tertiary > secondary > primary.
Only long-chain tertiary and—to a smaller extent—quarternary amines are used
in industrial extraction, because of their suitable physical properties; trioctylam-
ine (TOA, 8 carbons per chain) and trilauryl amine (TLA, 12 carbons per chain)
are the most frequently used. For simplicity we abbreviate all amines by RN,
and their salts by RNH + L−.
The tertiary and quaternary amine bases are viscous liquids at room tem-
perature and infinitely soluble in nonpolar solvents, but only slightly soluble in
water. The solubility of the ion-pair RNH + L− in organic solvents depends on
the chain length and on the counterion, L−: the solubility of TLA ⴢ HCl in wet
benzene is 0.7 M, in cyclohexane 0.08 M, in CHCl3 1.2 M, and in CCl4 0.7 M.
Nitrate and perchlorate salts are less soluble, as are lower molecular weight
amines.
The formation of the ion pair salt can be written
←
RN(org) + H + + L− → RNH + L− (org) K ex,am (4.61)
Table 4.15 gives the equilibrium constants Kex,am (for extraction, amine) for this
reaction with trioctylamine in various solvents. Although the ion pairs are only
slightly soluble in water, they can exchange the anion L− with other anions, X−,
in the aqueous phase. (Note that we use L− to indicate any anion, while X− is
used for (an alternative) inorganic anion.)
RNH + L− (org) + X − ← NH + X − (org) + L− K ex,ch (4.62)
→
Solvent Extraction Equilibria 159
F− Toluene 3.0
Cl− Toluene 5.9
Cl− Carbontetrachloride 4.0
Cl− Benzene 4.1
Br− Toluene 8.0
NO−3 Toluene 6.6
NO−3 Carbontetrachloride 5.0
SO24− a Carbontetrachloride 6.7
SO24− a Benzene 8.3
a
The equilibrium constant refers to the formation of (R3NH)2SO4.
Source: Ref. 5.
The equilibrium constant (Kex,ch for extraction, exchange) for this reaction in-
creases in the order ClO−4 < NO−3 < Cl − < HSO−4 < F −. From Chapter 3 it follows
that the formation constants for the metal complex MLnz−n usually increases in
the same order. Therefore, in order to extract a metal perchlorate complex, very
high ClO−4 concentrations are required; the perchlorate complex is easily re-
placed by anions higher in the sequence.
All negatively charged metal complexes can be extracted by liquid anion
exchangers, independent of the nature of the metal and the complexing ligand.
Liquid anion exchange has extensive industrial application, and many examples
are given in later chapters. Lists of extraction values are found in other works
[e.g., 5, 21].
A priori, it must be assumed that the aqueous phase contains all the stepwise
complexes MLnz−n. Thus the distribution ratio is
−p
[(RNH) +p
p ML n ]org βp [L− ]p [RNHL]org
p
DM = = K ex (4.64)
Σ [MLzn−n ] 1 + Σβp [L− ]p
The distribution of M depends on both the free amine salt in the organic phase
and the concentration of free L− in the aqueous phase until all metal in the
aqueous phase is bound in the ML−np complex. At constant amine concentration,
Eq. (4.64) indicates that a plot of DM vs. [L−] would have a linear slope p if the
denominator of Eq. (4.64) is Ⰶ1; i.e., the metal species in the aqueous phase
are dominated by the uncomplexed metal ion Mz+. At higher [L−] concentrations,
where the ML−np complex begins to dominate in the aqueous phase, the DM value
p
becomes equal to Kex [RNHL]org . Equations (4.64) and (4.4) show that S-shaped
curves result for metals with large Kex values. In a plot of DM vs. [RNHL]org a
straight line of slope p is obtained only at constant [L−]. From such measure-
ments both p, Kex and βp can be evaluated. The following example illustrates
this.
Example 8: Extraction of Trivalent Actinides by TLA.
In an investigation of the extraction of trivalent actinides, An(III) from
0.01 M nitric acid solutions of various LiNO3 concentrations into o-xylene
containing the tertiary amine salt trilaurylmethylammonium nitrate, TLMA
HNO3, Van Ooyen [29] found that the amine was monomeric only at very low
concentrations (≤0.1M in the organic phase) but at higher concentration formed
both dimers and trimers.
Using trace concentrations of Ce(III) and An(III) a log-log plot of DM
against the nitrate ion activity, mγ± = [LiNO3]1/2, had a slope of approximately
3, Fig. 4.16b. From Eq. (4.64) this slope corresponds to the p-value of 3 when
the aqueous phase is dominated by the free metal ion, which is not an unrea-
sonable assumption at low nitrate concentrations.
Solvent Extraction Equilibria 161
Fig. 4.16 Distribution ratio of M3+ ions between the trilaurylmethyl ammonium nitrate
(TLMA) in o-xylene and aqueous phases of varying LiNO3 concentrations. (a) As a
function of TLMANO3 concentration at 1–7 M, 2–5 M, 3–3 M LiNO3. (b) Extraction
of Eu(III) and tree actinide(III) ions at 0.1 M TLMANO3 in o-xylene and varying aque-
ous salt concentrations. (From Ref. 29.)
162 Rydberg et al.
When A− is added into the aqueous phase as a Na+ salt in large excess to K+,
the dissociation in the organic phase becomes negligible and Eq. (4.67) is re-
duced to
DK = K ex [A − ] (4.68)
Figure 4.17 shows the distribution ratio of K+ when a large excess of
Na+A− is added to the system. Although the extracted complex should be com-
Fig. 4.17 Distribution ratio of potassium(I) between nitrobenzene and water as a function
of initial aqueous tetraphenylborate concentration in 0.1M NaClO4. (From Refs. 30a,b.)
164 Rydberg et al.
Fig. 4.18 Decrease in distribution ratio of Be(II) as a function of oxalate ion (Ox2−)
concentration due to formation of aqueous BeOx2n−2n complexes. The extraction system is
0.03 M TTA in methylisobutylketone and 1.0 M Na(0.5 Ox2−, ClO−4 ). See Eq. (4.72) for
ordinate function. (From Ref. 31.)
166 Rydberg et al.
formation strongly dominates over hydrolysis, the latter can be neglected. In the
overview in section 4.3, it was shown in Fig. 4.3 that the general shape of the
extraction curves indicates the type of complexes formed. Thus, the extraction
of gallium(III) with acetylacetone, Fig. 4.19, indicates a behavior according to
Fig. 4.3c. in the plot of log D vs. free ligand concentration, [A−].
Example 11: Formation of hydrolyzed Ga(III) acetylacetonates.
Figure 4.19 shows the extraction of Ga(III) by acetylacetone into ben-
zene at various concentrations of total acetylacetone, [HA]t, and constant ionic
strength, using the AKUFVE technique (see section 4.15.3). By comparing
with Fig. 4.3, one may guess that hydroxy complexes are formed (diagram c).
If the general complex is designated GamAn(OH)p, a first investigation showed
Fig. 4.19 The distribution of Ga(III) between benzene and 1 M NaClO4 as a function
of pA = −log[A−] at 6 different acetylacetone (HA) concentrations (0.06–0.0006 M).The
different curves are due to different extent of hydrolysis in the aqueous phase. The con-
centration of HA decreases from the upper left corner, where [HA]aq is 0.06 M (▲),
toward the right lower corner, where it is 0.0006 M (䉫). (From Ref. 32.)
Solvent Extraction Equilibria 167
electrophilic groups which can increase acidity are HO−, NC−, NO2−, etc.
Conversely, nucleophilic substitution by aliphatic or aromatic groups usually
has little effect on the pKa, though it may affect the distribution constant KDR
(see section 4.13.2): e.g., the addition of a CH3− or C6H5− group in acetic acid
changes the pKa to 4.9 (propionic acid) and 4.3 (phenyl acetic acid). The further
the substitution is from the carboxylic group, the less the effect: e.g., while pKa
for benzoic acid is 4.19, a Cl in orthoposition yields pKa 2.92 and in the parapos-
ition, 3.98.
These general rules hold rather well for the acidic organic extractants and
can be used to extrapolate from related compounds to new ones as well as to
develop new extractants. The effect of various substituends on pKa is extensively
discussed in textbooks on organic chemistry [e.g., 14, 34].
4.13.1.2 The Complex Formation (or Stability)
Constant βn
Kex also increases with increasing formation constant of the uncharged metal
complex, βz. Thermodynamic factors and geometrical aspects that influence the
size of βn are discussed in Chapter 3. Some further observations follow.
To achieve the highest possible D value, the concentration of the extracted
uncharged complex (MAz, MAzBb or BbMLz+b) must be maximized. A large
value for the formation constant of the neutral complex favors this goal. In
Chapter 3 it was pointed out that for hard acids the complex formation constant
increases with the charge density of the metal ion, provided there is no steric
hindrance. This is seen for the metal fluoride complexes in Fig. 3.4 (no steric
hindrance). For the lanthanide acetylacetonates in Fig. 4.15b, the increase in βn
from 57La to 64Gd is due to a reduction in the ionic radius and, accordingly, to
the increased charge density. It should be noted that the diminishing lanthanide
size leads to a successive diminution of the coordination number (from 9 to 8
for H2O) for La → Gd; for Gd → Lu the coordination number is probably con-
stant, as is inferred from Fig. 4.15c; see also [35].
4.13.1.3 Correlations Between Ka and βn
Because the H+ ion acts as a hard metal ion, one should expect a close correla-
tion between the formation constants for HA and for MAz. Since Ka is defined
as the dissociation constant of HA, while βz is the formation constant of MAz,
a correlation is therefore expected between pKa (= −logKa) and logβz. The corre-
lations in Figs. 3.5 and 3.6 show that large logβn values are usually observed
for organic acids with large logKa. Since Kex is directly proportional to both βz
and K za [see Eq. (4.74)], and a large βz value is likely to be accompanied by a
small Ka value (large pKa); in Eq. (4.74) the two parameters counteract each
other. Thus in order to obtain high free ligand concentrations, which favors the
formation of the MAz complex, [H+] must be low, i.e., pH must be high. How-
Solvent Extraction Equilibria 169
ever, at high pH the metal may be hydrolyzed. A careful balance must be struck
between βz, Ka, and pH to achieve an optimum maximum concentration of the
uncharged complex.
KDC values for the lanthanide acetylacetonates are the reverse from that expected
for the size effect. The explanation is likely that the neutral complex with the
formula LnA3 is coordinatively unsaturated, which means that a hydrated com-
plex exists in the aqueous phase (possibly also in the organic phase). The more
coordinatively unsaturated lanthanide complexes (of the larger ions) can accom-
modate more water and thus are more hydrophilic. The result is a KDC several
orders of magnitude lower for the lightest, La, than for the heaviest, Lu.
The tetravalent metal acetylacetonates [36] have all coordination numbers
(CN) 8 or 9 in the neutral complexes. For Th (radius 109 pm at CN 9) there is
often one molecule of water in ThA4, whereas for the corresponding complexes
of U(IV) and Np(IV) (radius 100 and 98 pm, respectively, CN = 8), there seems
to be none. The measured distribution constants (log KDC) are for Th 2.55 and
for U and Np 3.52 and 3.45, respectively, in 1 M NaClO4/benzene. This agrees
with the greater hydrophilic nature of the ThA4 ⴢ H2O complex.
4.13.2.2 Correlations Between KDR and KDC
Qualitatively, the distribution constants of the complexing reagent and the corre-
sponding neutral metal complex are related due to their similar outer surface
toward the solvents. Thus it is reasonable that there should be good correlation
between KDR and KDC for homologous series of reagents or complexes. This is
indeed the case, as is seen when KD,R is plotted vs. KD,C, in Fig. 4.20 for a
complex with a variation of the solvent, and in Fig. 4.21 for complexes of
related reagents with the same solvent. This correlation can be related to the
solubility parameter concept discussed later.
Fig. 4.20 Distribution constants KDC for uncharged metal complexes MAz, vs. distribu-
tion constants KDR for the corresponding undissociated reagent, the acid ligand HA, for
various organic solvents and 1 M NaClO4: open circles Cu(II), solid circles Zn(II). Num-
bers refer to solvents listed in Table 4.10. (From Ref. 36.)
Pioneering work on the application of this theory for correlating and pre-
dicting distribution ratios was done in the 1960s [38a–40c]. Several reviews on
the use of this theory for two-phase distribution processes are also available
[41,42]. Recently this theory has been refined by the use of Hansen solubility
parameters [6,43,44], according to which
δ 2total = δdispersion
2
+ δ polar
2
+ δ hydrogen
2
(4.75c)
where interactions between the solute and the solvent are described by contribu-
tions from the various types of cohesive forces. In general, the dispersion (or
London) forces dominate. In Fig. 4.22 the measured distribution ratios of an
americium complex between an aqueous nitrate solution and various organic
solvent combinations are compared with distribution ratios calculated from tabu-
lated Hansen parameter values [45].
Few solubility parameters are available for the metal-organic complexes
discussed in this chapter. Another approach is then necessary. The distribution
constant for the reagent (extractant), R, can be expressed as:
172 Rydberg et al.
Fig. 4.21 Distribution constants, KDC, for uncharged metal complexes MAz vs. distribu-
tion constants KDR for the corresponding undissociated acid ligand HA; solid circles
Zn(II), solid triangles Co(III). Variation with ligand composition: HFA hexafluoroace-
tylacetone, TFA trifluoroacetylacetone, AA acetylacetone, FTA 2-furoyltrifluoroacetone,
TTA 2-thenoyltrifluoroacetone, PTA pivaloyltrifluoroacetone, BFA benzoyltrifluoroace-
tone, BZA benzoylacetone. (From Ref. 36.)
Solvent Extraction Equilibria 173
Fig. 4.22 Comparison of measured and calculated distribution ratios DAm of americi-
um(III)-terpyridine-decanoic acid complexes between 0.05 M HNO3 and various organic
solvent combinations. The calculated values are obtained with the Hansen partial solubil-
ity parameters. (From Ref. 45.)
Thus from solubility parameters, which are specific for the various solutes
and solvents, and molar volumes, values for KDR can be estimated, or deviations
from regularity can be assessed. These deviations can be estimated quantita-
tively and, in individual systems, can be ascribed to specific reactions in either
of the phases, e.g., hydration, solvation, adduct formation, etc.
From Eq. (4.76) the relation
log K DC = (VC / VR ) log K DR + const. (4.78)
can be derived; the subscript C refers to the neutral metal complex. The molar
volume ratio is close to or smaller than z, where z is the number of singly
charged anionic species attached to the central metal ion Mz+. From Fig. 4.23 a
ratio of ca. 1.5 (for z = 2) is obtained. This equation is useful for estimating KDR
values in extraction systems where the corresponding KDC value is known.
Rearrangement of Eq. (4.75) shows that log KDR/(δaq − δorg) vs. δ, where
δ = δorg RT (δaq − δorg ) −1 (Vorg
−1
− Vaq−1 ) (4.79)
should yield a straight line with slope VR / RT ln10. A plot of this relation in
Fig. 4.23 demonstrates a satisfactory correlation between distribution constants
and solvent parameters, indicating the usefulness of the solubility parameter
concept in predicting KDR as well as KDC values.
Fig. 4.23 Application of the regular solution theory for correlation of distribution con-
stants for ZnA2 and CuA2 with solvent properties (solubility parameters); the numbers
refer to the solvents listed in Table 4.10. (From Ref. 22.)
Solvent Extraction Equilibria 175
hexane 0.95, cyclohexane 1.03, ethylbenzene 3.31, and benzene 5.93 (aqueous
phase 1 M NaClO4), the decrease observed in Fig. 4.24 may simply be an effect
of the increasing aliphatic character of the solvents.
Fig. 4.25 The distribution constant KDC for zinc diacetylacetonate, ZnA2, between ben-
zene and various concentrations of NaClO4 in the aqueous phase; 25°C. (From Ref. 36.)
Solvent Extraction Equilibria 177
Fig. 4.26 Distribution constant KDC for metal diacetylacetonates, MA2, between or-
ganic solvents and 1 M NaClO4 as a function of temperature: Cu(II) (open circles) and
Zn(II) (solid circles) with various solvents (numbers refer to the solvents given in Table
4.10). (From Ref. 36.)
and (3) the number of chelate rings formed (mono- or polydentate). The follow-
ing diagram illustrates the general structure of UO22+ and Am3+ diamide chelate
complexes: two oxygen binding atoms; ring size is 5 + n, 2–3 chelate rings per
complex; the −O−M−O− part is commonly referred to as the “claw.”
| |
N Cn H 2n N
/ \ / \ / \
CH CH
O O
\ /
M
The extractants TBOA, TBMA, and TBSA are very similar, but their structural
differences (see formulas in Table 4.16) allow the formation of only one type
of metal chelate complex: 5-, 6-, and 7-membered rings, respectively. Similar-
ily, the reactants DMDOMA and DMDOSA form only 6- and 7-membered
chelates. Table 4.16 shows that extraction (i.e., largest Kex value) is favored by
6-membered rings. This is not unexpected as the Kex values in this case reflect
the stability constants βn acc. to Eq. (4.74).
The electron shells of the M3+ elements with unfilled or partially filled 3d,
4d, and 5f orbitals contract as these shells are being filled with electrons, in-
creasing the charge density (z/r) of the cation, leading to increasing stability
constants for the MA3 complexes with increasing atomic number. For bidentate
ligands the f-electron (lanthanide) MA3 complexes are coordinatively unsatu-
rated, i.e., only six of eight available coordination sites are filled. Therefore, the
two remaining sites are occupied by H2O or HA or some other donor molecules
B, leading to the self-adduct MA3HA−1 or adduct MA3B1–2. It has been shown
that the adduct formation constants for the reaction MA3 + B → MA3B decrease
in the same order [27a–28], see Fig. 4.15; i.e., the stronger the complex, the
weaker the adduct.
The increase, by adduct formation, in the coordination number of the cen-
tral atom is possible not only for weak but also for strong chelates, provided
their ligand bites are relatively short. To make room in the inner sphere of the
metal ion for another adduct-forming ligand, three chelating ligands of small
bite angle can easily be shifted away without significant energy-consuming dis-
tortion of the chelate rings [53a,b]. The electronic structure of the central atom
is also of key importance for synergism, as illustrated by the easy adduct forma-
tion of metal ions with unfilled or partially filled d or f orbitals, contrary to,
e.g., p-block elements. Inner- and outer-sphere complexation is further discussed
in the next two paragraphs and in Chapter 16.
4.13.7.2 Donor Ligand Effects
Table 4.17 shows extraction constants for some metal ions with three alkyl
phosphates substituted by 0, 1, and 2 sulfur atoms, but with almost identical
aliphatic branchings. Section 3.3 discusses hard and soft acids and bases (HSAB
theory). According to this theory, hard acids form strong complexes with hard
bases, while weak acids form strong complexes with weak bases. In Table 4.17,
the metals are ordered in increasing hardness, the sub II.b group being rather
soft (Table 3.2). Presumably the Kex values reflect this pattern, as they are pro-
portional to K az and βz. In Table 4.17, the acidity of the acids increases (i.e., Ka
increases) in order R ⴢ PSSH < R ⴢ POSH < R ⴢ POOH (consult Tables 4.8 and
4.9), as sulfur is less basic than oxygen. In general, the Kex increases for for the
dialkyl phosporic acid (hard donor ligand) with increasing metal charge density
within each group as predicted in Chapter 3. For the soft metals, Kex also in-
creases with increasing softness of the ligand, while the opposite effect is seen
for the hard Ln-metal ions. The divalent subgroup II.b metals prefer to bind to
sulfur rather than to oxygen because they have a rather soft acceptor character,
while the hard metals III.b prefer to bind to the hard O-atom of the ligand.
Table 4.17 Comparison of the Extraction Constants Kex for the IIb-Subgroup Divalent
Ions, and IIIb Lanthanide Ions, with Sulfur or Oxygen Dialkyl Phosphoric Acidsa
Metal ion Hg2+ Cd2+ Zn2+ 57 La3+ 63 Eu3+ 71 Lu3+
Fig. 4.27 Structure of the UO2(NO3)2 ⴢ 2(C2H5O)3PO. The uranyl oxygens are situated
perpendicular to the plane shown around the central atom. (From Ref. 12.)
182 Rydberg et al.
Fig. 4.28 Extraction of Eu(III) and Am(III) from K,HNO3 with 0.02 M oligopyridine
or triazine and 1 M 2-bromodecanoic acid in tert-butylbenzene. (From Ref. 59.)
Fig. 4.29 Structure of various oligopyridine and triazine adducts, and distribution ra-
tios (inserts) for Am(III) complexes with these oligopyridines (0.02 M) and 2-bromode-
canoic acid (0.00025 M) in tert-butylbenzene and 0.01 M HNO3. (From Ref. 59.)
Rydberg_5063-2_Ch05_R3_02-09-04 11:29:22
This method is useful when only one species is extracted, but it has little value
for the study of solvent extraction systems that contain several complex species.
4.14.1.2 Ligand Number Method
This method [65–67] is useful for identifying the average composition of the
metal species in the system. Consider Eq. (4.36a) for the extraction of MAz, and
assume—for the moment—that only one species, MAn, exists in the aqueous
phase. Taking the derivative of the logarithm of Eq. (4.36) yields
d log D / d log [A] = z − n (4.80)
In 1941, J. Bjerrum [65] developed the useful concept of average ligand num-
ber, n̄, defined as the mean number of ligands per central atom:
n = Σ n [MA n ]/[M]t (4.81)
It can be shown [66a–67] that n̄ equals n in Eq. (4.79), which can be rewritten
n = z − d log D / d log [A] (4.82)
For example, in Fig. 4.10 (Example 3), the slope of the plot of log DCu vs. log
[A−] can be used to conclude what species dominate the system at a given [A−]
value. The relation in Eq. (4.82) indicates an asymptote of slope 2, so the aque-
ous phase is dominated by uncomplexed Cu2+ (n = 0), while for slope 0 the
neutral complex CuA2 dominates the system (n = 2). Equation (4.82) shows that
any tangent slope of the curve (i.e., d log D / d log [A]) yields the difference
z − n̄ in these simple systems.
Example 12: Extraction of Th(IV) by acetylacetone.
Figure 4.30 shows a smoothed curve of measurements of the distribution
of Th(IV) from 0.1 M NaClO4 into chloroform containing the extractant acetyl-
acetone (HA) [66a,b]. Taking the derivative of this curve according to Eq.
(4.79) the average ligand number is derived as shown in the lower insert.
Th(IV) is successively complexed by A− forming ThA3+, ThA22+, ThA+3 and un-
charged ThA4, which is extracted. At pA > 8.5 the n̄-value is zero, i.e., Th is
uncomplexed, while at pA < 3.5 the average ligand number of 4 is reached,
i.e., Th is fully complexed as ThA4. See also Example 15.
The average ligand number can be used to obtain approximate equilibrium con-
stants, as described by [65], assuming that at half integer n̄-values the two adja-
cent complexes dominate: e.g., at n̄ = 0.5 the species Mz+ and MAz−1 dominate,
while at n̄ = 1.5 MAz−1 and MA2z−2 dominate, etc.. The following expression for
the stepwise formation constant is approximately valid at
n = n − 0.5, log K n ≈ −log [A]n (4.83)
This method of obtaining an estimate of the formation constants was done as a
first step in the Th(IV)-acetylacetone system in Fig 4.30, where in the lower
Solvent Extraction Equilibria 187
Fig. 4.30 Upper curve: the distribution of Th(IV) between benzene and 0.1 M NaClO4
as a function of aqueous acetylacetonate ion concentration; pA = −log[A−]; the asymp-
totes have slopes 0 and −4. Lower curve: The average number of ligands per central
atom, n̄, in same system, as obtained from a derivation of the log D(pA) curve. Using
the ligand number method, the following equilibrium constants were estimated (with
values from graphical slope analysis within parenthesis): logK1 8.0 (7.85), logK2 7.6
(7.7), logK3 6.4 (6.3), and logK4 5.1 (5.0). Log KD4 2.50 is obtained from the horizontal
asymptote. (From Refs. 66a,b.)
figure, n̄ is plotted against −log [A−], yielding the preliminary log Kn values
given in Fig. 4.3. In a similar manner, the adduct formation constants can be
determined:
Example 13: The extraction of Zn(II) by β-diketones and phosphoryl adduct
formers (cont. of Example 5).
Consider the formation of adducts of the type MAzBb, as described in
Example 5 (Fig. 4.12). The derivative of Eq. (4.48b) with respect to [B] at
constant [A] yields
d log D / d log [B] = b (4.84)
where b is the average number of adduct forming molecules in the molecule
at given [B] value. In Fig. 4.12 the asymptote has a slope of 2, indicating that
a maximum of two molecules of TBP (or TOPO) bind to the neutral metal
complex. From the lower slopes of the curve, the average number, b, can be
estimated according to Eq. (4.83).
188 Rydberg et al.
A plot of F1 vs. x−1 yields a1 at the intercept and a2 as the slope. In third step,
F2 = (F1 − a1) x = a2 + a3 x−1 + a4 x−2 is calculated, yielding a2 and a3, etc. Using
this technique all four an values are obtained, from which one can deduce the
βn values [actually the Kn values in Eq. (3.5)] and the KDC value.
Dyrssen and Sillén [68] pointed out that distribution ratios obtained by
conventional batchwise techniques are often too scattered to allow the determi-
nation of as many parameters as used in Examples 15 and 16. They suggested
a simplified graphic treatment of the data, based on the assumption that there is
a constant ratio between successive stability constants, i.e., Kn/Kn+1 = 102b, and
that all distribution curves can be normalized so that N −1 log βN = a, where N is
the number of ligands A− in the extracted complex. Thus, the distribution curve
log DM vs. log[A−] is described by the two parameters a and b, and the distribu-
tion constant of the complex, KDC. The principle can be useful for estimations
when there is insufficient reliable experimental data.
4.14.2.2 Nonlinear Plots
It is not possible to obtain simple linear relations between D and the variables
when both the aqueous or organic phase contain several metal species. Instead
a double polynomial such as
b0 + b1 + b2 y 2 + b3 y 3 + L Σ bi y i
D =Y = = (4.89)
a0 + a1 x + a2 x 2 + a3 x 3 + L Σ an x n
is obtained for which there is no simple solution.
Example 16: Extraction of U(VI) by acetylacetone.
Fig. 4.31 shows the extraction of U(VI) by acetylacetone (HA) from
nitric acid solutions, DU, as a function of pA = −log [A−] at six different total
concentrations of HA. A comparison with Fig. 4.3 indicates that we can expect
complexes of the type MAn(OH)p(HA)r, which formally is equivalent to
Hr+pMAn+r. The extraction reaction is of the type Eq. (4.86), as one expects a
series of self-adduct complexes MAn(HA)r in the organic phase, and one or
two series of stepwise complexes MAn and M(OH)p in the aqueous phase. Thus
the extraction reaction is of the type Eq. (4.89) with a ratio of several polyno-
mials. Even in this case, a graphic extrapolation technique was useful by deter-
mining a set of intermediary constants for each constant [HA]org from logDU
vs. pA. Plotting these intermediate constants vs. [HA]org, a new set of constants
were obtained, from which both the stepwise formation constants βn,p and the
adduct formation constants Kad, 1 and Kad,2 for the reaction UO2A2(H2O)2–3 +
nHA → ← UO2A2(HA)1–2 in the organic phase. A more detailed analysis also al-
lowed the determination of formation constants for UO2(OH)p (p = 1 and 2).
There may be other explanations to the extraction results. However, with the
model used and the constants calculated, log DU could be correctly predicted
over the whole system range [U]tot 0.001–0.3 M, pH 2–7, and [HA]org 0.01–1.0
M. See also Example 17.
190 Rydberg et al.
Fig. 4.31 Extraction of U(VI) by acetylacetone (HA) from 0.1 M NaClO4 into chloro-
form as a function of pH at different total concentrations of HA. (From Ref. 69.)
Rydberg_5063-2_Ch05_R3_02-09-04 11:29:22
ratio D. The least square method requires that S (the weighted squared residu-
als) in expression Eq. (4.91) is minimized
2
L
N
S = ∑ wi ∑ (an xin ) − yi (4.91)
i =1 n =0
where N + 1 is the number of parameters, and L the number of experimentals
points; N + L − 1 is referred to as the number of degrees of freedom of the
system; each point has a value xi / yi. Because experiments are carried out over
a large range of D, [A−] and/or [B] values, the points carry different algebraic
weight (e.g., the value 1000 obscures a value of 0.001). Therefore, in order to
use the LSQ technique properly, each point must be correctly weighted, wi. This
can be done in several ways, the most common being to weight it by yi, or by
σ−i 2, or by a percentage value of yi ; σi is the standard deviation in the measure-
ment of yi. The difference an xni − yi (the residual) is not zero, because the differ-
ence is to be taken between a measured value, yi (meas.), and the corresponding
calculated value, yi (calc.), by the an x ni function [i.e., yi(calc.) − yi(meas.)], using
the actual an values at the time of the operation.
The principle of the LSQ technique is to compute the set of positive an
values that give the smallest sum of the residuals; Eq. (4.91) reaches the Smin
value. If the residuals equal zero (which rarely occurs in practice), Smin would
be zero, and there would be a perfect fit between the experimental points and
the calculated curve.
There are several mathematically different ways to conduct the minimiza-
tion of S [see Refs. 70–75]. Many programs yield errors of internal consistency
(i.e., the standard deviations in the calculated parameters are due to the devia-
tions of the measured points from the calculated function), and do not consider
external errors (i.e., the uncertainty of the measured points). The latter can be
accommodated by weighting the points by this uncertainty. The overall reliabil-
ity of the operation can be checked by the χ2 (chi square) test [71], i.e., Smin/(L
+ N − 1) should be in the range 0.5–1.5 for a reasonable consistency between
the measured points and the calculated parameters.
Example 17: Extraction of Pm(III) by acetylacetone.
In Example 7, it was concluded that a number of self-adducts PmA3
(HA)b were formed in the organic phase (0 < b < 2) in addition to the PmAn
complexes in the aqueous phase (0 < n < 4); some extraction curves are given
in Fig. 4.14. Equation (4.55) is of the same form as Eq. (4.89); rewriting Eq.
(4.55b)
Y = DPm Σβn [A]n / β3 [A]3 = K DC (1 + K ad1[HA]org + L) (4.92)
and
Y[HA]org = C2 DPm Σβn [A]n / β3 [A]3 (4.93b)
Y[HA]org contains the measured DPm, the [A] values, and the βn values. It can thus
in principle be treated as an Example 16 to yield the βn values and the C2 values
(one for each [HA]org), remembering that a0 = 1. Because Y[A] /C1 = Y[HA]org/C2
= Y′, a second plot of Y′ against [HA]org yields the KD3 and Kadb values. The
analysis of the system yielded KDC = 0.008, Kad,1 = 7 and Kad,2 = 3. In these cal-
culations, a SIMPLEX program was used [75].
In comparisons of equilibrium constants collected from the literature (e.g.,
Fig. 4.22 or [47]), or correlations of data for a large number of systems (e.g.,
Figs. 4.20–4.23), it is desirable to present both the statistical uncertainty of each
“point,” which is often given by the standard deviation (one or several σ’s) of
the point, and the general reliability (statistical significance) of the whole corre-
lation [76], for which the chi-square test offers a deeper insight into the reliabil-
ity of the experimental results [77]. More advanced statistical tests for systems
of our kind have been described by Ekberg [78].
Fig. 4.32 A thermostated double jacket (1) cell for solvent extraction studies (heavier
phase 8, lighter phase 9) under nonoxidizing conditions, using a hydrogen gas inlet (6)
to a Pd-black catalyst (11), pH glass electrode (3), magnetic stirrer (10), connections for
additions (5). Alternative constructions contain rotating paddles and fixed pipings con-
nected to the two phases for frequent sample withdrawals.
in order not to destroy the interface. The stirring rate is optimized to the time
for reaching equilibrium and complete phase separation. The experiments are
either carried out with intermittent violent stirring, in which case samples are
withdrawn after each complete phase separation, or with mild stirring during
which it is possible to continously withraw samples. Equal volumes are sampled
each time, commonly <1 mL. The simple stirred cell has been improved by
introducing phase discriminating membranes in the sampling outlet [79]. This
is particulary advantageous for kinetic experiments and is further described in
Chapter 5. The sampling of the stirred cell can be automated, so that at regular
intervals pH and temperature are recorded, and samples withdrawn for automatic
analysis of concentration of interesting species in more or less standard fashion.
It is also possible to use ion-sensitive probes in one of the phases instead of
sampling.
illustrated in Fig. 4.33. Efficient mixing of the two phases and their additions is
achieved in the mixing vessel and at the inflow into the phase separator, which
consists of a continuous flow centrifuge, which in a special separation chamber
and at very high rotational velocity (5,000–50,000 rotations per minute) sepa-
rates the mixture into two very pure phases, containing <0.01% of entrainment
of one phase in the other phase. The H-centrifuge may be made of Pd-stabilized
Fig. 4.33 The AKUFVE solvent extraction apparatus: Efficient mixing is achieved in
the separate mixing vessel, from which the mixture flows down into the continuous
liquid flow centrifugal separator (the H-centrifuge, hold-up time <1 s). (From Refs.
83a,b.) The outflow from the centrifuge consists of two pure phases, which pass on-line
detectors, AMXs, for on-line detectors or continuous sampling. (From Refs. 80a–80d,
81.)
196 Rydberg et al.
Ti, or PEEK (polyether ether ketone) to allow measurements under very corro-
sive conditions. The separated phases pass AMX gadgets for on-line detection
(radiometric, spectrophotometric, etc.) or phase sampling for external measure-
ments (atomic absorption, spectrometric, etc.), depending on the system studied.
The aqueous phase is also provided with cells for pH measurement, redox con-
trol (e.g., by reduction cells using platinum black and hydrogen, metal ion deter-
mination, etc.) and temperature control (thermocouples).
The AKUFVE technique allows a large number of points (50–100) to be
determined in a one-day experiment over a D-range of better than 103 to 10−3,
not counting time of preparation. In a special version of this technique (LISOL
for LIquid Scintillation On Line) [82], the D range 10−5 to 104 has been accu-
rately covered, as for the Pm-acetylacetone system, Example 7 and Fig. 4.14.
The centrifugal separator of the AKUFVE system is also used for phase
separation in the SISAK technique [84]. SISAK is a multistage solvent extrac-
tion system that is used for studies of properties of short-lived radionuclides,
e.g., the chemical properties of the heaviest elements, and solvent extraction
behavior of compounds with exotic chemical states. In a typical SISAK experi-
ment, Fig. 4.34, radionuclides are continuously transported from a production
Fig. 4.34 A typical SISAK setup used for studies of α-decaying nuclides, e.g., trans-
actinides. (From Ref. 84.)
Solvent Extraction Equilibria 197
REFERENCES
1. Grahame, D. C.; Seaborg, G. T. J. Am. Chem. Soc. 1938, 60:2524.
2. Irving, H. p. 91, ISEC’66, Solvent Extraction Chemistry Dyrssen, D.; Liljenzin,
J. O.; Rydberg, J. Eds.; North Holland, Amsterdam (1967).
3a. Rydberg, J. Arkiv Kemi 1954, 8:101.
3b. Rydberg, J. Rec. Trav. Chim. 1956, 75:737.
4. Sekine, T.; Hasegawa, Y. Solvent Extraction Chemistry, Marcel Dekker, New
York, 1977.
5. Marcus, Y.; Kertes, S. Ion Exchange and Solvent Extraction of Metals, Wiley-
Interscience, Chichester, 1969.
6. Barton, A. F. M. CRC Handbook of Solubility Parameters and Other Cohesion
Parameters, CRC Press, 1991.
7. Kojima, I.; Yoshida, M.; Tanaka, M. J. Inorg. Nucl. Chem. 1970, 32:987.
8a. Dyrssen, D. Svensk Kem. Tidskr. 1956, 68:212.
8b. Dyrssen, D. Svensk Kem. Tidskr. 1952, 64:213.
9. Sekine, T. et al. Bull. Chem. Soc. Jap. 1983, 56:700.
10. Rydberg, J. Svensk Kem. Tidskr. 1950, 62:179.
11a. Sekine, T.; Isayama, M.; Yamaguchi, S.; Moriya, H. Bull. Chem. Soc. Jap. 1967,
40:27.
11b. Dyrssen, D. Private communication. See also Acta Chem. Scand. 1960, 14:1091.
12. Ahrland, S.; Liljenzin, J. O.; Rydberg, J. The Chemistry of the Actinides. In Com-
prehensive Inorganic Chemistry, Pergamon Press, London, 1973.
13. Jassim, T. M.; Fridemo, L.; Liljenzin, J. O. in T. M. Jassim, Co-Extraction of
Pertechnetate with some Metal Nitrates in TBP-Nitric Acid Systems. Diss. Chal-
mers Techn. Univ., Gothenburg (1986).
14. Martell, A. E.; Calvin, M. Chemistry of the Metal Chelate Compounds, Prentice-
Hall, Englewood Cliffs, N.J., 1952.
15. Stary, J. The Solvent Extraction of Metal Chelates, Pergamon Press, New York,
1964.
16. Zolotov, Yu. Extraction of Chelate Compounds, Humprey Sci. Publ., Ann Arbor,
London, 1970.
17a. Narbutt, J. J. Inorg. Nucl. Chem. 1981, 43:3343.
17b. Narbutt, J. J. Phys. Chem. 1991, 95:3432.
17c. Narbutt, J.; Fuks, L. Radiochimca Acta 1997, 78:27.
18. Liljenzin, J. O.; Stary, J.; Rydberg, J. ISEC’68, Solvent Extraction Research,
Kertes, A. S.; Marcus, Y., Eds., J. Wiley and Sons, New York, 1969.
19. Sillén, L. G.; Martell, A.; Högfeldt, E., Eds., Stability Constants of Metal-Ion Com-
plexes. The Chemical Society Spec. Publ. No. 25, Burlington House, London, 1971.
20. Martell, A. E.; Smith, R. M. Critical Stability Constants, Vol. 1–5, Plenum Press,
New York 1974–1982.
21. Marcus, Y.; Kertes, A. S.; Yanir, E. Equilibrium Constants of Liquid-liquid Distri-
bution Reactions, Part 1, Butterworths, London, 1974.
22. Sekine, T.; Dyrssen, D. J. Inorg. Nucl. Chem. 1964, 26:1727.
23a. Allard, B.; Johnsson, S.; Rydberg, J. ISEC’74, Proc. Int. Solvent Extraction Conf.,
Lyon, 1974.
Solvent Extraction Equilibria 199
23b. Allard, B.; Johnsson, S.; Narbutt, J; Lundquist, R. ISEC ‘77, CIM Special Volume
21, 1977.
24. Sekine, T.; Ihara, N. Bull. Chem. Soc. Jap. 1971, 44:2942.
25. Cecconie, T.; Freiser, H. Anal. Chem. 1990, 62:622.
26. Moryia, H.; Sekine, T. Bull. Chem. Soc. Jap. 1971, 44:3347.
27a. Rydberg, J.; Albinsson, Y. J. Solv. Extr. Ion Exch. 1989, 7:577.
27b. Albinsson, Y.; Rydberg, J. Radiochim. Acta 1989, 48:49.
27c. Albinsson, Y. Acta Chem. Scand. 1989, 43:919.
28. Albinsson, Y. Development of the AKUFVE-LISOL Technique. Solvent Extraction
Studies of Lanthanide Acetylacetonates. Diss. Chalmers Univ. Techn., Gothenburg
(1988).
29. Van Ooyen, J.; Dyrssen, D.; Liljenzin, J. O.; Rydberg, J. (Eds.) Solvent Extraction
Chemistry (ISEC 1966), North-Holland Publ. Co., Amsterdam, 1967.
30a. Sekine. T.; Dyrssen, D. Anal. Chim. Acta 1969, 45:433.
30b. Sekine, T.; Fukushima, T.; Hasegawa, Y. Bull. Chem. Soc. Japan 1970, 43:2638.
31. Sekine, T.; Sakairi, M. Bull. Chem. Soc. Jap. 1967, 40:261.
32. Liljenzin, J. O.; Vadasdi, K.; Rydberg, J. Coordination Chemistry in Solution Hög-
feldt, E. Swedish Nat. Res. Council; Trans.Roy. Inst.Techn., 280, Stockholm, 1972.
33a. Rydberg, J. Rev. Inorg. Chem. 1999, 19:245.
33b. Rydberg, J. Min. Pro. Ext. Met. Rev. 2000, 21:167.
34. Carey, F. A.; Sundberg, R. J. Advanced Organic Chemistry, Third Edition, Plenum
Press, New York, 1990.
35. Fuks, L.; Majdan, M. Min. Pro. Ext. Met. Rev. 2000, 21:25.
36. Allard, B. The Coordination of Tetravalent Actnide Chelate Complexes with β-
Diketones, Diss. Chalmers Univ. Techn. Gothenburg, 1975.
37. Hildebrand, J. H.; Scott, R. C. Solubilities of Nonelectrolytes, Dover Publishers,
1964.
38a. Siekierski, S. J. Inorg. Nucl. Chem. 1962, 24:205.
38b. Siekierski, S.; Olszer, R. J. Inorg. Nucl. Chem. 1963, 24:1351.
39. Buchowski, H. Nature 1962, 194:674.
40a. Wakahayashi, T.; Oki, S.; Omori, T.; Suzuki, N. J. Inorg. Nucl. Chem. 1964, 26:
2255.
40b. Omori, T.; Wakahayashi, T.; Oki, S.; Suzuki, N. J. Inorg. Nucl. Chem. 1964, 26:
2265.
40c. Oki, S.; Omori, T.; Wakahayashi, T.; Suzuki, N. J. Inorg. Nucl. Chem. 1965, 27:
1141.
41. Irving, H. M. N. H. Ion Exchange and Solvent Extraction, Vol. 6, (Marinsky, J.;
Marcus, Y. Eds.) Marcel Dekker, New York, 1973.
42. Siekierski, S. J Radioanal. Chem. 1976, 73:335.
43. Hansen, C.; Skaaup, K. Dansk Kemi 1967, 48:81.
44. Hansen, C. Hansen Solubility Parameters: A Users Handbook, CRC Press, 2000.
45. Nilsson, M. Influence of Organic Phase Composition on the Extraction of Trivalent
Actinides, Dipl. work, Dept. Nucl. Chem., Chalmers Univ. of Techn., Gothenburg,
2000.
46. Johansson, H.; Rydberg, J. Acta Chem. Scand. 1969, 23:2797.
47. Hancock, R. D.; Martell, A. Chem. Rev., 1989, 89:1875–1913.
200 Rydberg et al.
76. Meinrath, G.; Ekberg, C.; Landgren, A.; Liljenzin, J. O. Talanta, 2000, 51:231.
77. Rydberg, J.; Sullivan, J. Acta Chem. Scand. 1959, 13:2057.
78. Ekberg, Ch. Uncertainties in Actinide Solubility Calculations Illustrated Using the
Th−OH−PO4 System, Diss. Chalmers tekniska högskola, Göteborg, 1999.
79. Watari, H.; Cunningham, L.; Freiser, H. Anal. Chem. 1982, 54:2390.
80a. Reinhardt, H.; Rydberg, J. Solvent Extraction Chem. (ISEC) 1966, North-Holland
Publ. Co. Amsterdam, 612 pp.
80b. Reinhardt, H.; Rydberg, J. Chem. Ind. 1970, 488 pp.
80c. Rydberg, J., et al. Acta Chem. Scand. 1969, 23:647.
80d. Rydberg, J., et al., Acta Chem., 1969, 2773, 2781, 2797 pp.
81. MEAB Metal Extraction Co Ltd, V. Frölunda, Sweden.
82. Albinsson, Y.; Ohlsson, L. E.; Persson, H.; Rydberg, J. Appl. Radiat. Isot. 1988,
39:113.
83a. Rydberg, J.; Reinhardt, H. Centrifuge for complete phase separation of two liquids,
U.S.Patent 3,442,445, 1966.
83b. Reinhardt, H. En kontinuerlig separator för två-fas vätskeblandningar, Diss. Chal-
mers Univ. Techn., Gothenburg, 1973.
84. Persson, H.; Skarnemark, G.; Skålberg, M.; Alstad, J.; Liljenzin, J. O.; Bauer, G.;
Heberberger, F.; Kaffrell, N.; Rogowski, J.; Trautmann, N. Radiochim. Acta 1989,
48:177.
85. Skarnemark, G., et al., J. Nucl. Radiochem. Sci. 2002, 3:1.
86. Skarnemark, G. J. Radioanl. Nucl. Chem. 2000, 243(1):219.
5
Solvent Extraction Kinetics
PIER ROBERTO DANESI International Atomic Energy Agency,
Vienna, Austria
203
204 Danesi
Fig. 5.1 Interfacial diffusion films. δo and δw are the thickness of the organic and
aqueous films, respectively. The presence of an adsorbed layer of extractant molecules
at the interface is also shown.
These two thin liquid films, which are also called diffusion films, diffusion
layers, or Nernst films, have thicknesses that range between 10−2 and 10−4 cm
(in this chapter centimeter-gram-second (CGS) units are used, since most pub-
lished data on diffusion and extraction kinetics are reported in these units; com-
parison with literature values is, therefore, straightforward).
The description of the diffusion films as completely stagnant layers, having
definite and well-identified thicknesses, represents only a practical approxima-
tion useful for a simple mathematical description of interfacial diffusion. A
206 Danesi
more realistic physical description should consider that, starting at a given dis-
tance from the liquid interface, the renewal of the organic and the aqueous
fluids becomes progressively less as the interface is approached.
5.2 DIFFUSION
Diffusion is that irreversible process by which matter spontaneously moves from
a region of higher concentration to one of lower concentration, leading to equal-
ization of concentrations within a single phase.
The laws of diffusion correlate the rate of flow of the diffusing substance
with the concentration gradient responsible for the flow. In general, in a multi-
component system, the process is described by as many diffusion equations as
the number of chemical species in the system. Moreover, the diffusion equations
are intercorrelated. Nevertheless, when dealing with solutions containing ex-
tractable species and extracting reagents at concentrations much lower than
those of the molecules of the solvents (dilute solutions), diffusion is sufficiently
well described by considering only those species for which the concentrations
appreciably change during the extraction reaction.
The diffusion flow, J, of a chemical species is defined as the amount of
matter of this species passing perpendicularly through a reference surface of
Solvent Extraction Kinetics 207
unit area during a unit time. The dimensions of J are those of mass per square
centimeter per second (mass cm−2 s−1). When the concentrations of the diffusing
species (c) are expressed in moles per cubic centimeter (mol cm−3), the correla-
tion between the flux across a unit reference area located perpendicularly to the
linear coordinate x (along which diffusion occurs) and the concentration gradi-
ent, ∂c /∂x, will be given by the Fick’s first law of diffusion, that is,
J = − D∂c / ∂x (5.1)
D is the diffusion coefficient of the species under consideration. Its dimen-
sions are given in cm2 s−1. Equation (5.1) implies that D is independent of con-
centration; it is a constant, characteristic of each diffusing species, in a given
medium, at constant temperature. This is only approximately true, but holds
sufficiently well for most diffusing species of interest in solvent extraction sys-
tems. The values of D for the majority of the extractable species fall in the
range 10−5 –10−6 cm2 s−1 in both water and organic solutions. When dealing with
viscous phases or bulky organic extractants, or with complexes that may un-
dergo polymerization reactions, the D values can drop to 10−7 or 10−8 cm2 s−1.
Equation (5.1) is extremely useful to evaluate the flux whenever the con-
centration gradient can be considered constant with time (i.e., when we can
assume a steady state). When a steady state cannot be assumed, the concentra-
tion change with time must also be considered. The non–steady-state diffusion
is expressed by Fick’s second law of diffusion:
∂c / ∂t = D∂ 2c / ∂x 2 (5.2)
Diffusion is a complex phenomenon. A complete physical description in-
volves conceptual and mathematical difficulties associated with the need to in-
volve theories of molecular interactions and to solve complicated differential
equations [3–6]. Here and in sections 5.8 and 5.9, we present only a simplified
picture of the diffusional processes, which is valid for limiting conditions. The
objective is to make the reader aware of the importance of this phenomenon in
connection with solvent extraction kinetics.
For steady-state diffusion occurring across flat and thin diffusion films,
only one dimension can be considered and Eq. (5.1) is greatly simplified. More-
over, by replacing differentials with finite increments and assuming a linear
concentration profile within the film of thickness δ, Eq. (5.1) becomes
J = − D dc / dx = − D(c2 − c1 ) / δ = (c2 − c1 ) / ∆ (5.3)
where
∆ = δ/ D (5.4)
is a diffusional parameter dependent on the thickness of the diffusion film and the
value of the diffusion constant of the diffusing species. The units of ∆ are cm−1 s.
208 Danesi
Fig. 5.2 Concentration profile of a solute diffusing across an unstirred liquid film of
thickness δ in contact with two well-stirred reservoirs. The same liquid phase is assumed
throughout this imaginary model system.
Solvent Extraction Kinetics 209
the concentration of a reactant X, and the rate of the reverse reaction, rate
(reverse), is always equal to zero. Simple distribution processes between immis-
cible liquid phases of noncharged components characterized by very large val-
ues of the partition constant can be formally treated as first-order irreversible
reactions.
Reversible first-order reactions are given by Eq. (5.15) of Table 5.1.
Equation (5.15c) can be rewritten in a more elegant form by considering that at
equilibrium (t = ∞ ) the net rate is equal to zero, that is
k1[X]eq = k2 [Y]eq (5.17)
where subscript eq indicates equilibrium concentrations. The equilibrium con-
stant of the reaction is
K = k1 / k2 = [Y]eq /[X]eq (5.18)
Equation (5.15c) can then be reformulated as
ln{([X] − [X]eq )([X]0 − [X eq ]) −1} = −(k1 + k2 )t (5.19)
Equation (5.19) shows that the rate law is still first-order, provided the
quantity ([X] − [X]eq) is used instead of [X]. A plot of ln([X] − [X]eq) vs. t will
then be a straight line of slope -(k1 + k2). The individual rate constants of the
reaction can still be evaluated from the slope of such a plot, providing the
212 Danesi
5.4.2 Mechanisms
Solvent exchange and complex formation are special cases of nucleophilic sub-
stitution reactions.
Generally, when a coordination bond is broken in a metal ion complex,
the electron pair responsible for the bond accompanies the leaving group (com-
monly, the organic ligand). It follows that reactions involving solvent exchange
and complex formation have as reactive centers electron-deficient groups. There-
fore, they can all be considered as nucleophilic substitutions (in symbols SN).
The basic classification of nucleophilic substitutions is founded on the
consideration that when a new metal complex is formed through the breaking
of a coordination bond with the first ligand (or water) and the formation of a
new coordination bond with the second ligand, the rupture and formation of the
two bonds can occur to a greater or lesser extent in a synchronous manner.
When the rupture and the formation of the bonds occur in a synchronous way,
the mechanism is called substitution nucleophilic bimolecular (in symbols SN2).
On the other extreme, when the rupture of the first bond precedes the formation
of the new one, the mechanism is called substitution nucleophilic unimolecular
(in symbols SN1). Mechanisms SN2 and SN1 are only limiting cases, and an
entire range of intermediate situations exists.
Inorganic chemists prefer a slightly more detailed classification and subdi-
vide the range of possible mechanisms into the following four groups.
1. D mechanism. This is the limiting dissociative mechanism, and a transient
inter-mediate of reduced coordination number is formed. The intermediate
persists long enough to discriminate between potential nucleophiles in its
vicinity. Here we are dealing with an SN1 limiting process, since the dissoci-
ation of the metal-ligand bond fully anticipates the formation of the new
bond. The vacancy in the coordination shell that occurs as a result of the
dissociation is then taken by the new ligand.
2. Id mechanism. This is the dissociative interchange mechanism and is similar
to the previous one in the sense that dissociation is still the major rate-
controlling factor. Therefore, we are still dealing with an SN1 process. Nev-
ertheless, differently from the D mechanism, no experimental proof exists
that an intermediate of lower coordination is formed. The mechanism in-
volves a fast outer sphere association between the initial complex and the
214 Danesi
incoming nucleophile (the new ligand), and when the complex metal ion
and the incoming ligand have opposite charges, this outer sphere associate
is an ion pair. As a result of the formation of this new outer sphere associ-
ate, the new group is now suitably placed to enter the primary coordination
sphere of the metal ion as soon as the outgoing group has left.
3. Ia mechanism. This is the associative interchange mechanism and is similar
to the Id mechanism, in the sense that here, also, no proof exists that an
intermediate of different coordination is formed. However, differently from
Id, we have significant interaction between the incoming group and the
metal ion in the transition state. In this instance, the process is partially of
the SN2 type.
4. A mechanism. This is the pure associative mechanism, and here we are
dealing with a process entirely of the SN2 type. The rate-determining step
of the mechanism is the association between the complex metal ion and the
entering ligand, leading to the formation of an intermediate of increased
coordination number that can be experimentally identified. This mechanism
is often operative in ligand-displacement reactions occurring with planar
tetracoordinated complexes.
In practice, there is a continuous gradation of mechanisms from D,
through Id and Ia to A, depending on the extent of interaction between the metal
cation and the incoming group in the transition intermediate. Moreover, the
diagnosis of the mechanism is not a straightforward and unambiguous process.
A variety of methods have to be used, depending on circumstances, since an
approach based only on rate-law consideration can easily lead to false conclu-
sions.
5.3 on a logarithmic scale. They have been obtained by nuclear magnetic reso-
nance (NMR) and sound absorption, for the rapidly exchanging ions, and by
isotopic labeling for the more inert ones.
Generally, in solvent extraction, for the simplest case of an organic com-
plexing reagent (ligand) reacting with an aquometal ion, when the water ex-
change rate constant is <102 s−1, the reaction can be considered slow enough that
the complex formation rate may compete with the rate of diffusion through the
interfacial films in controlling the overall extraction kinetics. In exceptional
cases, dealing with very efficiently stirred phases (in which the thickness of the
diffusion films may even be reduced to about 10−4 cm), complexation reactions
with rate constants as high as 106 s−1 can be rate limiting in solvent extraction.
In all other cases, the ions can be considered as reacting instantaneously relative
to the rate of film diffusion. The rate of diffusion also is the predominating rate-
controlling process when the aquocations are extracted into the organic phase
with their entire coordination sphere of water molecules. These considerations
do not hold when other slow chemical processes, such as hydrolysis reactions
of the metal cation, polymerizations in the organic phase of the extractant or
the metal complex, adsorption-desorption processes at the interphase, keto-enol
Fig. 5.3 Logarithm of water exchange constants (s−1) for aquometal ions at 25°C.
(From Ref. 26.)
Solvent Extraction Kinetics 217
by assuming the presence on either side of the interface of two stagnant films
(the diffusional films). Although these films are thin when compared with the
bulk phases, their thickness is still macroscopic (about 10−3 or 10−4 cm) in com-
parison with the molecular dimensions and the range of action of the molecular
and ionic forces. Thus, these films have a finite volume, and the concentrations
of chemical species within them must always be considered as volume concen-
trations. Moreover, the physicochemical properties of the liquids within these
films, such as density, viscosity, dielectric constant, charge distribution, are the
same as those of the bulk phases. However, the situation changes when the
distance from the interface approaches the order of magnitude of the molecular
dimensions and the range of action of the molecular and ionic forces. In this
region, generally defined as the interface, the physicochemical properties differ
from those of the bulk phases and, therefore, any chemical change occurring
herein is likely to be affected by the different environment. The difference will
be particularly enhanced by the presence on the interface of adsorbed layers of
polar or ionizable extractant molecules. The extractant, because of its simultane-
ous hydrophobic-hydrophilic (low water affinity–high water affinity) nature,
necessary to maintain a high solubility in low dielectric constant diluents, and
selective complexing power relative to water-soluble species, tends to orient
itself with the polar (or ionizable) groups facing the aqueous side of the inter-
face. The rest of the molecule, having a prevalent hydrophobic character, will
be directed instead toward the organic phase.
Unfortunately, little direct information is available on the physicochemical
properties of the interface, since real interfacial properties (dielectric constant,
viscosity, density, charge distribution) are difficult to measure, and the interpre-
tation of the limited results so far available on systems relevant to solvent ex-
traction are open to discussion. Interfacial tension measurements are, in this
respect, an exception and can be easily performed by several standard physico-
chemical techniques. Specialized treatises on surface chemistry provide an ex-
haustive description of the interfacial phenomena [10,11]. The interfacial ten-
sion, γ, is defined as that force per unit length that is required to increase the
contact surface of two immiscible liquids by 1 cm2. Its units, in the CGS system,
are dyne per centimeter (dyne cm−1). Adsorption of extractant molecules at the
interface lowers the interfacial tension and makes it easier to disperse one phase
into the other.
Interfacial tension studies are particularly important because they can pro-
vide useful information on the interfacial concentration of the extractant. The
simultaneous hydrophobic-hydrophilic nature of extracting reagents has the re-
sulting effect of maximizing the reagent affinity for the interfacial zone, at
which both the hydrophobic and hydrophilic parts of the molecules can mini-
mize their free energy of solution. Moreover, as previously mentioned, a prefer-
ential orientation of the extractant groups takes place at the interface. Conse-
Solvent Extraction Kinetics 219
quently, most solvent extraction reagents are interfacially adsorbed and produce
a lowering of the organic-aqueous interfacial tension. The extent of this adsorp-
tion is a function of the chemical nature and structure of the hydrophilic and
hydrophobic groups, of the extractant bulk concentration, and of the physico-
chemical properties of the diluent. The extractant interfacial concentration can
be extremely high, depending on the way molecules can pack themselves at the
interface, and even moderately strong surfactants can form an interface fully
covered with extractant molecules when their bulk concentration is as low as
10−3 M or less. It follows that, when an extractant is a strong surfactant and
exhibits low solubility in the water phase, the interfacial zone is a region where
a high probability exists that the reaction between an aqueous soluble species
and an organic soluble extracting reagent can take place. This is why a strong
surface activity of an extractant can sometimes lend support to extraction mech-
anisms involving interfacial chemical reactions.
Interfacial concentrations can be evaluated from interfacial tension mea-
surements by utilizing the Gibbs equation
d ∏ / d ln c = nikT (5.26)
In Eq. (5.26), Π is the interfacial pressure of the aqueous-organic system,
equal to (γ0 − γ) [i.e., to the difference between the interfacial tensions without
the extractant (γ0) and the extractant at concentration c (γ)], c is the bulk organic
concentration of the extractant, and ni is the number of adsorbed molecules of
the extractant at the interface. The shape of a typical Π vs. ln c curve is shown
in Fig. 5.4; ni can be evaluated from the value of the slopes of the curve at each
c. However, great care must be exercised when evaluating interfacial concentra-
tions from the slopes of the curves because Eq. (5.26) is only an ideal law, and
many systems do not conform to this ideal behavior, even when the solutions
are very dilute. Here, the proportionality constant between dΠ/d ln c and ni is
different from kT. Nevertheless, Eq. (5.26) can still be used to derive infor-
mation on the bulk organic concentration necessary to achieve an interface
completely saturated with extractant molecules (i.e., a constant interfacial con-
centration). According to Eq. (5.26), the occurrence of a constant interfacial
concentration is indicated by a constant slope in a Π vs. ln c plot. Therefore,
the value of c at which the plot Π vs. ln c becomes rectilinear can be taken as
the bulk concentration of the extractant required to fully saturate the interface.
Many extractants reach a constant interfacial concentration at bulk organic
concentrations far below the practical concentrations that are generally used to
perform extraction kinetic studies. This means that when writing a rate law for
an extraction mechanism that is based on interfacial chemical reactions, the
interfacial concentrations can often be incorporated into the apparent rate con-
stants. This leads to simplifications in the rate laws and to ambiguities in their
interpretation, which are discussed in later sections.
220 Danesi
Fig. 5.4 Typical interfacial pressure (Π) vs. logarithm of bulk organic concentration
(log c) plotted for an extractant exhibiting surface-active properties.
Fig. 5.5 Simplified picture of the interfacial structuring effect caused by an ionizable
extractant adsorbed at the organic-water interface.
curve of rate of extraction vs. stirring rate is shown in Fig. 5.6. Such curves
are generally obtained both when constant interfacial-area-stirred cells
(Lewis cells) and vigorously mixed flasks are used (see section 5.10 for a
description of the various techniques). In general, a process occurring under
the influence of diffusional contributions is characterized by an increase of
the rate of extraction, as long as the stirring rate of the two phases is
increased (see Fig. 5.6, zone A). When, on the other hand, the rate of ex-
traction is independent of the stirring rate (see Fig. 5.6, zone B), it is some-
times possible to assume that the extraction process occurs in a kinetic
regime.
The rationale behind this criterion is that an increase in stirring rate produces
a decrease in thickness of the diffusion films. Moreover, since at low stirring
rates the relationship between stirring rate and the inverse thickness of the
diffusion films (at least in a constant interfacial-area-stirred cell) is approxi-
mately linear, the first part of such a plot will usually approximate a straight
line. In any case, the rate of extraction will increase with the rate of stirring,
as long as a process is totally or partially diffusion controlled. When, eventu-
ally, the thickness of the diffusion films is reduced to zero, only chemical
reactions can be rate controlling, and the rate of extraction becomes indepen-
dent of the stirring rate. Unfortunately, this kind of reasoning can lead to erro-
Fig. 5.6 Typical extraction rate vs. stirring rate for constant interfacial–area-stirred cell.
226 Danesi
on the nature of the reagents involved in the extraction reaction. Since, in many
metal extractions, ligand substitution reactions take place, rate laws similar to
those for complexation reactions in solution may be expected. During most ex-
traction processes, coordinated water molecules or ligands are substituted in part
or wholly by molecules of a more organophilic ligand (the extractant) or of the
organic diluent. When the extractant shows little solubility in the aqueous phase
and is a strong surfactant, the ligand substitution reaction may take place at the
interface. The highest concentration of extractant molecules is at the interface,
so the probability of reaction is higher than in the bulk phases.
In this section, we describe three simple cases of rates and mechanisms
that have been found suitable for the interpretation of extraction kinetic pro-
cesses in kinetic regimes. These simple cases deal with the extraction reaction
of a monovalent metal cation M+ (solvation water molecules are omitted in the
notation) with a weakly acidic solvent extraction reagent, BH. The overall ex-
traction reaction is
M + + BH(org) → MB(org) + H + (5.27)
with equilibrium extraction constant, Kex equal to
K ex = [MB]eq [H + ]eq /[M + ]eq [BH]eq (5.28)
Ideal behavior of all solute species will be assumed.
Case 1: The rate-determining step of the extraction reaction is the aque-
ous phase complex formation between the metal ion and the anion of the extract-
ing reagent. Even if at very low concentration, BH will always be present in
the aqueous phase because of its solubility in water. The rate-determining step
of the extraction reaction is as follows:
k1
M + + B− ↔ MB (slow) (5.29)
k−1
−1
rate = k1K a / K DB [M + ][BH][H + ]−1 − k−1K DM
−1
[MB] (5.34)
In the right-hand side of Eq. (5.34), the first term represents the forward
rate of extraction and the second term the reverse rate of extraction.
The values of the apparent rate constants of the extraction reaction,
−1 −1
k1K a K DB and k−1K DM (5.35)
permit the evaluation of the rate constants of the aqueous complex formation
reaction only if KDB, KDM, and Ka are known. Here, comparisons can be made
with literature values of the rate constants for reaction in aqueous solutions
between the same metal ion and aqueous soluble ligands containing the same
complexing groups as the extracting reagent. Agreement between the values
supports, but does not prove, the validity of the proposed mechanism.
The apparent values of the rate constants of the solvent extraction reaction
are usually evaluated by measuring the rate of extraction of M+ as function of
[BH] (at [H+] constant), of [H+] (at [BH] constant), and of [MB] (at [H+] and
[BH] constant). The experimental conditions are usually chosen in such a way
that the reaction can be assumed pseudo-first-order for [M+]. The apparent rate
constants are evaluated from the slope of the straight lines obtained by plotting
ln ([M + ] − [M + ]eq )/([M + ]0 − [M + ]eq ) vs. t (5.36)
Subscript 0 indicates initial concentrations in the aqueous phase.
It must be observed that when this mechanism holds, the rate of the extrac-
tion reaction is independent of the interfacial area, Q, and the volume of the
phases, V. The expected logarithmic dependency of the forward rate of extrac-
tion on the specific interfacial area (S = Q/V), the organic concentration of the
extracting reagent and the aqueous acidity, is shown in Fig. 5.7, case 1.
Case 2: The rate-determining step of the extraction reaction is the inter-
facial formation of the complex between the metal ion and the interfacially
adsorbed extracting reagent. Here, the rate-determining step of the extraction
reaction can be written as
k2
M + + B− (ad) ↔ MB(ad) (slow) (5.37)
k−2
where (ad) indicates species adsorbed at the liquid–liquid interface. The rate of
the reaction that follows is
rate = −d [M + ]/ dt = k2 S w [M + ][B− ]ad − k2 So [MB]ad (5.38)
where Sw and So represent the specific interfacial areas [i.e., the ratios between
the interfacial area Q (cm2) and the volumes of the aqueous (Vw) and organic
(Vo) phase, respectively]. Throughout the following treatments we will always
assume, for simplicity, that Vw = Vo = V, and
Solvent Extraction Kinetics 229
Fig. 5.7 Logarithm of the forward extraction rate (normalized to the extractable metal
concentration) vs. log interfacial area (a), log [BH], and log [H+] for some exemplary
kinetic regimes.
230 Danesi
Q / V = S = SW = So (5.39)
To derive a rate of extraction in terms of easily measurable quantities, we
express the concentration of the reagent adsorbed at the liquid–liquid interface
as function of its bulk organic concentration. This can be done by utilizing the
Langmuir’s adsorption law: that is,
[BH]ad = (α 2 [BH]/ γ 2 ) /(1 + [BH]/ γ 2 ) (5.40)
In Eq. (5.40), α2 and γ2 are Langmuir adsorption constants. Their values
are characteristic of each specific system. We can then distinguish two regions
of adsorption. One defined as ideal, where 1 Ⰷ [BH]/ γ2, yielding
[BH]ad = (α 2 [BH]/ γ 2 ) = α[BH] (5.41)
k4
MB(ad) + BH(org) ↔ BH(ad) + MB(org) (slow) (5.50)
k−4
The rate equations holding for the two slow steps are:
rate 1 = S k3 [M + ][B− ]ad − S k−3 [MB]ad (5.51)
232 Danesi
includes the diffusional contributions on the water side of the interface, assumed
to occur entirely through the stagnant film of thickness, δw. Ro includes the
contributions on the organic side, where the thickness of the stagnant film is δo.
The resistance associated with the crossing of the interface is indicated as Ri.
As a consequence, the total transfer resistance, R, is
R = Rw + Ri + Ro (5.64)
This equation holds at the steady state. In a diffusional regime and in ab-
sence of rigid interfacial films, Ri is generally negligible relative to Rw and Ro.
Consider two simple cases of extraction processes in which kinetics are
controlled by interfacial film diffusion (the solutions are always considered
stirred). The two cases are treated with the simplifying assumptions introduced
in section 2 (i.e., steady-state and linear concentration gradients throughout the
diffusional films).
Case 4: The interfacial partition between the two phases of unchanged
species is fast. The rate is controlled by the diffusion to and away from the
interface of the partitioning species. In the absence of an interfacial resistance,
the partition equilibrium of A between the aqueous and organic phase, occurring
at the interface, can be always considered as an instantaneous process. Here, A
is any species, neutral or charged, organic or inorganic. This instantaneous parti-
tion process (interfacial equilibrium) is characterized by a value of the partition
coefficient equal to that measured when the two phases are at equilibrium.
The following equations describe the extraction process:
A ↔ A(org) (5.65)
with an extraction equilibrium constant KDA, reached at t = ∞ ;
A i ↔ A i (org) (fast) (5.66)
interfacial equilibrium, holding at any time (subscript i indicates species in con-
tact with the interface); thus
K DA = [A]eq /[A]eq = [A]i /[A]i (5.67)
The symbol [ ]i indicates concentrations at the extreme limit of the diffu-
sional film (i.e., volume concentrations in the region in direct contact or very
close to the liquid-liquid interface). Because of the fast nature of the distribution
reaction, local equilibrium always holds at the interface.
The diffusional process can be treated by applying the first Fick’s diffu-
sion law to the two diffusional processes occurring on both sides of inter face,
under the simplifying assumptions of steady-state and linear concentration gra-
dients shown in Fig. 5.8. The diffusional fluxes through the aqueous, Jw and
organic, Jo, diffusion films will then be
J w = − DA ∂ [A]/ ∂x = − DA ([A]i − [A]) / δ w (5.68)
236 Danesi
Fig. 5.8 Linear concentration profiles for the partition of A between an aqueous and
an organic phase.
The diffusion-controlled process [see Eqs. (5.66), (5.68), and (5.69)] can
be experimentally differentiated from the process occurring in the kinetic regime
[see Eq. (5.75)] only by measuring the variations of kl and k−1 with δw and δo .
Otherwise, the identical rate laws will not permit one to distinguish between the
two mechanisms.
Equation (5.73) can be integrated to obtain
ln ([A]o − [A]eq ) /([A] − [A]eq ) = S ( K DA + 1)t /
( K DA ∆ w + ∆ o ) = S (k1 + k−1 )t (5.76)
Equations (5.74) and (5.76) show that when KDA is very high (i.e., the
partition of A is all in favor of the organic phase):
k1 ~ DA / δ w and k−1 ~ 0
and the extraction rate is controlled only by the aqueous-phase diffusional resis-
tance. On the other hand, when KDA is very small, it is
k1 ~ 0 and k−1 ~ DA / δo
and the rate is only controlled by the organic-phase diffusional resistance.
Case 5: A fast reaction between the metal cation and the undissociated
extracting reagent occurs at the interface (or in proximity of the interface). The
rate is controlled by the diffusion to and away from the interface of the species
taking part in the reaction.
The overall extraction reaction (extraction stoichiometry) is represented
by the equation:
M + + BH(org) = MB(org) + H + (5.77)
with equilibrium extraction constant Kex chemical reaction occurring at or near
the interface is considered to be always at equilibrium, although bulk phases
reach this condition only at the end of the extraction process. The condition of
interfacial (local) equilibrium is expressed by
M i+ + BH i (org) ↔ MBi (org) + H i+ (fast) (5.78)
with interface equilibrium constant Ki, which alternatively can be written as the
sum of the two equilibria.
M i+ + Bi− (org) ↔ MBi (org), K1 = [MB]i /[M + ]i [B− ]i (5.79)
238 Danesi
Fig. 5.9 Linear concentration profiles for the extraction of M+ by BH(org) in presence
of an interfacial chemical reaction.
Solvent Extraction Kinetics 239
In the limiting condition of very low metal ion concentrations, the terms con-
taining [M+] and [MB] can be neglected in the denominator of Eq. (5.88). This
means that [BH] and [H+], in practice, are constant with time, and their initial
and equilibrium concentrations are essentially the same. The rate equation then
becomes
rate = {S K ex [M + ][BH]eq }{∆ w K ex [BH]eq + ∆ o [H + ]eq }−1 −
− {S [MB][H + ]eq }{∆ w K ex [BH]eq + ∆ o [H + ]eq }−1 (5.89)
By dividing the numerators and denominators in Eq. (5.89) by [H+]eq and
setting
K ex* = K ex [BH]eq /[H + ]eq (5.90)
where K*e x is the conditional equilibrium constant of equilibrium [see Eq. (5.77)]
at [H+] and [BH] constant, Eq. (5.89) becomes identical with Eq. (5.73) when
KDA is replaced by Kex. As expected, the rate of extraction of species present at
high dilution in a system of fixed composition can be formally treated as the
simple distribution of an unchanged species between two immiscible phases.
The comparison of Eq. (5.89) with Eq. (5.63) also indicates that the rate laws
obtained for a system in a kinetic regime controlled by slow two-step interfa-
cial chemical reactions [see Eqs. (5.58) and (5.59)] and for a system in diffu-
sional regime controlled by slow film diffusion processes coupled to an in-
stantaneous reaction occurring at the interface (or in the proximity of the
interface) [see Eqs. (5.82a–d) and (5.78)], have the same functional depen-
dence on the concentration variables when low fluxes and low metal concen-
trations are involved. Alternatively, this result can be formulated by stating
that, because of diffusional contributions to the extraction rate, a fast extrac-
tion reaction occurring with a simple stoichiometry can mimic a two-step
interfacial chemical reaction occurring in absence of diffusional contributions
[14,15]. This type of ambiguity, in the past, has led solvent extraction chem-
ists (including this author) to sometimes erroneously derive extraction mech-
anisms invoking a series of two slow interfacial chemical reactions as rate
determining steps in systems for which the two phases were insufficiently
stirred. However, a distinction between the two cases can occasionally be
made if it is possible to conduct kinetic experiments in which the values of
δw and δo are varied. When the rate is independent of δw and δo, Eq. (5.63)
rather than Eq. (5.89) of this section describes the rate of extraction. Never-
theless, since the film thicknesses never seem to go to zero, the ambiguity
may not be resolved.
Thus, the chemical kinetics at the interface must be known from separate
experiments carried out in a pure kinetic regime. In fact, although the problem
has some similarities with that of a diffusional regime associated with a fast
interfacial chemical reaction, for which the equilibrium law was set up as an
interfacial boundary condition to the diffusion of the species, here, the further
complication exists that the interfacial chemical rate laws cannot be a priori
known and have to be themselves derived from kinetic experiments. This diffi-
culty can sometimes be avoided by assuming that the interfacial rate laws are
simply related to the reaction stoichiometry or by trying to derive information
on the interfacial reactions through fitting, to experimental data obtained in a
mixed regime, analytical solutions of differential equations that take into ac-
count both diffusion and chemical reactions. Unfortunately, both of these proce-
dures can lead to erroneous interpretations, as in only a few cases can the rate
laws be correctly derived from stoichiometric considerations, and when too
many variables (e.g., rate constants, rate laws, and diffusional parameters) are
simultaneously adjusted to fit experimental data, many alternative models can
usually satisfy the same equations. Therefore, it is important that the boundary
condition of the diffusion equations (i.e., the interfacial rate laws) be derived
by separate, suitable experiments.
We now treat, in some detail, under the usual assumptions of steady-state
diffusion and linear concentration gradients through the diffusion films, only
one simple case of mixed regime. The case deals with a situation that is an
extension of case 4 described in section 5.8. Here the additional complication
of a slow interfacial reaction is introduced.
Case 6: The interfacial partition reaction between the two phases of un-
charged species is slow. The rate is controlled both by the slow partition reaction
and by diffusion to and away from the interface of the partitioning species.
Although the extraction process is still described by Eq. (5.65), the slow
interfacial chemical reaction and the corresponding interfacial flux, Ji, are
k1
A i ↔ A i (org) (5.91)
k−1
on the extraction regime. Typical drop diameters are 200 µm, that is, about
10–20 times larger than the claimed minimum thickness of the diffusion films
usually reached in well-stirred biphasic systems.
phases are stirred. In presence of these baffles and grids, the range of variation
of the stirring speed of the phases can be quite wide. The grids (or screens)
transform most of the translation kinetic energy of the liquids into turbulent
energy, preventing the rippling of the interface. The cells allow an independent
change of the revolution number of the stirrers in the two phases. In this way,
the different influence of the density and viscosity of the two phases can be
taken into account, and the organic and aqueous phase can be stirred in such a
way to have similar fluid dynamic conditions.
Fig. 5.12 Cross section of the rotating diffusion cell: B, bearing; BA, internal baffle;
FM, filter mount; I, inner compartment; L, lid with holes; M, membrane; MA, mounting
rod; O, outer compartment; P, pulley block; S, hollow rod; T, thermostated beaker. (From
Ref. 20.)
Solvent Extraction Kinetics 247
sides. This allows description of the diffusional processes of the reactants and
products of the extraction reactions on the aqueous and the organic side of the
interface.
The basic assumption is that the rotating filter creates a laminar flow field that
can be completely described mathematically. The thickness of the diffusion
boundary layer (δ) is calculated as a function of the rotational speed (ω), vis-
cosity, density, and diffusion coefficient (D). The thickness is expressed by the
Levich equation, originally derived for electrochemical reactions occurring at
a rotating disk electrode:
δ = 0.643 ω−1 / 2 D1 / 3ν1 / 6 (5.96)
ν is the kinematic viscosity. The organic phase inside the membrane is con-
sidered at rest, and the membrane thickness and porosity control the diffusion
rate. The major drawbacks of this technique are difficulties in well characteriz-
ing the thickness and porosity of the membrane filter, in utilizing really thin
filters (less than 10−3 cm), and that, although the hydrodynamics inside the
rotating cylinder can be assimilated to those of a rotating disk, no studies have
conclusively shown that the flows inside and outside the cylinder are the same.
A more recent version of this cell [21] makes use of a rotating capillary,
containing a gelified aqueous phase. This eliminates the microporous membrane
filter.
extractant adsorbs on the drop surface, the internal circulation can be absent
or strongly reduced, and it is difficult to establish whether the degree of
turbulence of the two phases in the moving drop method can ever be enough
to ensure the measurement of a transfer rate in a kinetic regime or, on the
opposite side, is so low that the extraction rate is always occurring in a diffu-
sional regime.
SYMBOLS
REFERENCES
1. Whitman, W. G. Chem. Metal Eng., 1923, 29(4):146.
2. Whitman, W. G.; Davies, D. S. Ind. Eng. Chem., 1924, 16:1233.
3. Bird, R. B.; Stewart, W. E.; Lightfoot, E. N. Transport Phenomena, John Wiley
and Sons, New York, 1960.
4. Cussler, E. L. Diffusion, Mass Transfer in Fluid Systems, Cambridge University
Press, Cambridge, 1984.
5. Astarita, G. Mass Transfer with Chemical Reaction, Elsevier, Amsterdam, 1967.
6. Frank-Kamenetskii, D. A. Diffusion and Heat Transfer in Chemical Kinetics, Ple-
num Press, New York, 1969.
7. Frost, A. A.; Pearson, R. G. Kinetics and Mechnanism, Wiley-Interscience, New
York, 1956.
8. Moelwin-Hughes, E. A. The Kinetics of Reactions in Solution, Oxford University
Press, London, 1950.
9. Eigen, M.; Wilkins, R. W. Mechanisms of Inorganic Reactions, Vol. 49, Adv.
Chem. Ser., American Chemical Society, Washington, DC, 1965.
10. Adamson, A. W. Physical Chemistry of Surfaces, John Wiley and Sons, New York,
1982.
11. Davies, J. T.; Rideal, E. K. Interfacial Phenomena, Academic Press, New York,
1961.
12. Danesi, P. R.; Chiarizia, R.; Vandergrift, G. F. J. Phys. Chem., 1980, 84:3455.
13. Danesi, P. R.; Chiarizia, R. The Kinetics of Metal Solvent Extraction, Crit. Rev.
Anal. Chem., 1980, 10:1.
14. Danesi, P. R. Solv. Extr. Ion Exch. 1984, 2:29.
15. Danesi, P. R.; Vandegrift, G. E.; Horwitz, P.; Chiarizia, R. J. Phys. Chem., 1980,
84:3582.
16. Watarai, H.; Cunningham, L.; Freiser, H. Anal. Chem., 1982, 54:2390.
17. Lewis, J. B. Chem. Eng. Sci., 1954, 3:218.
18. Nitsch, W.; Kahn, G. Ger. Chem. Eng. Engl. Trans., 1980, 3:96.
19. Danesi, P. R.; Cianetti, C.; Horwitz, E. P.; Diamond, H. Sep. Sci. Technol. 1982,
17.961.
20. Albery, W. J.; Burke, J. E; Lefier, E. B.; Hadgraft, J. J. Chem. Soc. Faraday Trans.,
1976, 72:1618.
21. Simonin, J. P.; Musikas, C.; Soualhia, E.; Turq, P. J. Chim. Phys., 1987, 84:525.
22. Tarasov, V. V.; Yagodin, G. A.; Kizim, N. F. Dokl. Akad. Nauk. SSR, 1971, 45:
2517.
23. Tarasov, V. V.; Yagodin, G. A. VINITI, Khim. Teknol. Neorganich. Khim., 1974, 4.
24. Yagodin, G. A.; Tarasov, V. V. Proc. Int. Solv. Extr. Conf., ISEC 71, Vol. 1,
Solvent Extraction Kinetics 251
Society of Chemical Industry, (Gregory, J. G.; Evans, B; Weston, P.C. Eds.), Lon-
don, p. 888, 1971.
25. Tarasov, V. V.; Yagodin, G. A. Interfacial Phenomena in Solvent E. Ytraction. In
Ion Exchange Solvent Extraction, Vol. 10, Ch. 4 (Marinsky, J. A. and Marcus, Y.
Eds.), Marcel Dekker, New York, pp. 141–237, 1988.
26. Eigen, M. Pure Appl. Chem., 1963, 6:97.
6
Ionic Strength Corrections
INGMAR GRENTHE* Royal Institute of Technology,
Stockholm, Sweden
HANS WANNER Swiss Federal Nuclear Safety Inspectorate,
Villingen-HSK, Switzerland
*Retired.
By kind permission from the OECD NEA-TDB project, reprinted from Report TDB-2,
Guidelines for the extrapolation to zero ionic strength, OECD Nuclear Energy Agency, Data
Bank, F-91191 Gif-Sur Yvette, France. For latest revised version, see Chemical Thermo-
dynamics of Americium Silva, R. J., Bidoglio, G., Rand, M. H., Robouch, P. B., Wanner, H.,
and Puigdomenech, I., Eds., Elsevier/North-Holland 1995.
253
254 Grenthe and Wanner
The method preferred in the NEA Thermochemical Data Base review is the
specific ion interaction model in the form of the Brønsted-Guggenheim-
Scatchard approach.
One may sometimes have access to the parameters required for the
Pitzer approaches, e.g., for some hydrolysis equilibria and for some solu-
bility product data, cf. Baes and Mesmer [3] and Pitzer [4]. In this case,
the reviewer should perform a calculation using both the B-G-S and the
P-B equations and the full virial coefficient methods and compare the
results.
by Scatchard [7] and Guggenheim [8]. The two basic assumptions in the
specific ion interaction theory are:
Assumption 1: The activity coefficient gj of an ion j of charge zj in the
solution of ionic strength Im may be described by Eq. (6.1).
X
log 10 j ¼ z2j D þ "ð j;k;Im Þ mk (6.1)
k
t(8C) A B(108)
0 0.4913 0.3247
5 0.4943 0.3254
10 0.4976 0.3261
15 0.5012 0.3268
20 0.5050 0.3276
25 0.5091 0.3283
30 0.5135 0.3291
35 0.5182 0.3299
40 0.5231 0.3307
45 0.5282 0.3316
50 0.5336 0.3325
55 0.5392 0.3334
60 0.5450 0.3343
65 0.5511 0.3352
70 0.5573 0.3362
75 0.5639 0.3371
Baj, ranging from Baj = 1.0 [11] to Baj = 1.6 [12]. However, the parameter
Baj is empirical and as such correlated to the value of "ð j;k;Im Þ . Hence, this
variety of values for Baj does not represent an uncertainty range, but rather
indicates that several different sets of Baj and "ð j;k;Im Þ may describe equally
well the experimental mean activity coefficients of a given electrolyte. The
ion interaction coefficients listed later in the chapter in Tables 6.3–6.5 have
to be used with Baj = 1.5.
The summation in Eq. (6.1) extends over all ions k present in solution.
Their molality is denoted mk. The concentrations of the ions of the ionic
medium is often very much larger than those of the reacting species. Hence,
the ionic medium ions will make the main contribution to the value of
log10 gj for the reacting
P ions. This fact often makes it possible to simplify
the summation k "ð j;k;Im Þ mk so that only ion interaction coefficients
between the participating ionic species and the ionic medium ions are
included, as shown in Eqs. (6.5)–(6.9).
Assumption 2: The ion interaction coefficients "ð j;k;Im Þ are zero for ions of the
same charge sign and for uncharged species. The rationale behind this is that
e, which describes specific short-range interactions, must be small for ions of
the same charge, since they are usually far from one another due to elec-
trostatic repulsion. This holds to a lesser extent also for uncharged species.
Equation (6.1) will allow fairly accurate estimates of the activity
coefficients in mixtures of electrolytes if the ion interaction coefficients are
known. Ion interaction coefficients for simple ions can be obtained from
tabulated data of mean activity coefficients of strong electrolytes or from the
corresponding osmotic coefficients. Ion interaction coefficients for com-
plexes can either be estimated from the charge and size of the ion or
determined experimentally from the variation of the equilibrium constant
with the ionic strength. Ion interaction coefficients are not strictly constant
but vary slightly with the ionic strength. The extent of this variation depends
on the charge type and is small for 1:1, 1:2, and 2:1, electrolytes for mol-
alities less than 3.5 m. The concentration dependence of the ion interaction
coefficients can thus often be neglected. This point was emphasized by
Guggenheim [8], who has presented a considerable amount of experimental
material supporting this approach. The concentration dependence is larger
for electrolytes of higher charge. In order to accurately reproduce their
activity coefficient data, concentration-dependent ion interaction coef-
ficients have to be used: (see Pitzer and Brewer [14]; Baes and Mesmer [3];
or Ciavatta [10]). By using a more elaborate virial expansion, Pitzer
and coworkers [4, 15–21] have managed to describe measured activity
coefficients of a large number of electrolytes with high precision over a large
Ionic Strength Corrections 257
k mk
log 10 aH2 O ¼ (6.10)
lnð10Þ 55:51
where F is the osmotic coefficient of the mixture and the summation extends
over all ions k with molality mk present in the solution. In the presence of
an ionic medium NX in dominant concentration, Eq. (6.10) can be simpli-
fied by neglecting the contributions of all minor species, i.e., the reacting
ions. Hence, for a 1:1 electrolyte of ionic strength Im & mNX, Eq. (6.10)
becomes
2mNX
log 10 aH2 O ¼ (6.11)
lnð10Þ 55:51
1¼
A lnð10Þjzþ z j
mNX (1:5) 3 [ pffiffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffi
1 þ 1:5 mNX 2 log 10 (1 þ 1:5 mNX )
1
1 þ 1:5 mNX ]
pffiffiffiffiffiffiffiffiffiffi þ
1
"
ln (10Þ (N;X)
mNX (6.12)
The activity of water is obtained by inserting Eq. (6.12) into Eq. (6.11). It
should be mentioned that in mixed electrolytes with several components
at high concentrations, it is necessary to use Pitzer’s equation to calculate
the activity of water. On the other hand, aH2 O is near constant (and = 1)
in most experimental studies of equilibria in dilute aqueous solutions,
where an ionic medium is used in large excess with respect to the reactants.
The ionic medium electrolyte thus determines the osmotic coefficient of the
solvent.
In natural waters the situation is similar; the ionic strength of most
surface waters is so low that the activity of H2O(l) can be set equal to unity.
A correction may be necessary in the case of seawater, where a sufficiently
good approximation for the osmotic coefficient may be obtained by con-
sidering NaCl as the dominant electrolyte.
Ionic Strength Corrections 259
In more complex solutions of high ionic strengths with more than one
electrolyte at significant concentrations, e.g., (Na +, Mg2+, Ca2+) (Cl,
SO42), Pitzer’s equation may be used to estimate the osmotic coefficient; the
necessary interaction coefficients are known for most systems of geochem-
ical interest.
Note that in all ion interaction approaches, the equation for mean
activity coefficients can be split up to give equations for conventional single
ion activity coefficients in mixtures, e.g., Eq. (6.1). The latter are strictly
valid only when used in combinations that yield electroneutrality. Thus,
while estimating medium effects on standard potentials, a combination of
redox equilibria with H + + e) * 12 H2(g) is necessary (see Example 3).
The following formula is deducted from Eq. (6.6) for the extrapolation to
I = 0:
log 10 1 þ 4D ¼ log 10 1o "Im (6.14)
The linear regression is done as described in the NEA Guidelines for
the Assignment of Uncertainties [24]. The following results are obtained:
The experimental data are depicted in Fig. 6.1, where the area between the
dashed lines represents the uncertainty range that is obtained by using the
results in log10b1o and De and correcting back to I = 0.
Example 3: When using the specific ion interaction theory, the relationship
between the normal potential of the redox couple UO22+/U4+ in a medium
of ionic strength Im and the corresponding quantity at I = 0 should be
calculated in the following way. The reaction in the galvanic cell
is
þ
2 þ H2 (g) þ 2H Ð U
UO2þ þ 2H2 O
4þ
(6.16)
Ionic Strength Corrections 261
Fig. 6.1 Plot of log10b1 + 4D vs. Im for Eq. (6.13). The straight line shows the result of
the weighted linear regression, and the area between the dashed lines represents the
uncertainty range of log10b1o and De.
262 Grenthe and Wanner
Hence,
log 10 K o ¼ log 10 K 10D þ ð"ðU4þ ;ClO4 Þ "ðUO2þ ;ClO4 Þ
2
Table 6.3 Ion Interaction Coefficients ej,k for Cations j with k = Cl, ClO4, and NO3a
#
j k!
Cl ClO4 NO3
H+ 0.12 ± 0.01 0.14 ± 0.02 0.07 ± 0.01
NHþ 4 0.01 ± 0.01 0.08 ± 0.04b 0.06 ± 0.03b
ZnHCOþ 3 0.2c - -
CdCl + - 0.25 ± 0.02 -
CdI + - 0.27 ± 0.02 -
CdSCN + - 0.31 ± 0.02 -
HgCl + - 0.19 ± 0.02 -
Cu + - 0.11 ± 0.01 -
Ag + - 0.00 ± 0.01 0.12 ± 0.05b
UOþ 2 - 0.26 ± 0.03d -
UO2OH + - 0.06 ± 3.7d 0.51 ± 1.4d
ðUO2 Þ3 ðOHÞþ5 0.81 ± 0.17d 0.45 ± 0.15d 0.41 ± 0.22d
UFþ3 0.1 ± 0.1e 0.1 ± 0.1e -
UO2F + 0.04 ± 0.07f 0.29 ± 0.05d -
UO2Cl + - 0.33 ± 0.04d -
UO2 ClOþ 3 - 0.33 ± 0.04e -
UO2Br + - 0.24 ± 0.04e -
UO2 NOþ 3 - 0.33 ± 0.04e -
UO2SCN + - 0.22 ± 0.04e -
NpOþ 2 - 0.25 ± 0.05f -
PuOþ 2 - 0.17 ± 0.05f -
AlOH2 + 0.09g 0.31g -
Pb2 + - 0.15 ± 0.02 0.20 ± 0.12b
Zn2 + - 0.33 ± 0.03 0.16 ± 0.02
ZnCO2þ 3 0.35 ± 0.05c - -
Cd2 + - - 0.09 ± 0.02
Hg2 + - 0.34 ± 0.03 0.1 ± 0.1b
Hg2þ
2 - 0.09 ± 0.02 0.2 ± 0.1b
Cu2 + 0.08 ± 0.01 0.32 ± 0.02 0.11 ± 0.01
Ni2 + 0.17 ± 0.02 - -
Co2 + 0.16 ± 0.02 0.34 ± 0.03 0.14 ± 0.01
FeOH2 + - 0.38h -
FeSCN2 + - 0.45h -
Mn2 + 0.13 ± 0.01 - -
UO2þ2 0.21 ± 0.02i 0.46 ± 0.03 0.24 ± 0.03i
UF2þ
2 - 0.3 ± 0.1e -
USO2þ 4 - 0.3 ± 0.1e -
UðNO3 Þ2þ 2 - 0.49 ± 0.14j -
Mg2 + 0.19 ± 0.02 0.33 ± 0.03 0.17 ± 0.01
Ca2 + 0.14 ± 0.01 0.27 ± 0.03 0.02 ± 0.01
Ba2 + 0.07 ± 0.01 0.15 ± 0.02 0.28 ± 0.03
(continued )
264 Grenthe and Wanner
#
j k!
Cl ClO4 NO3
Al3 + 0.33 ± 0.02 - -
Fe3 + 0.56 ± 0.03 0.42 ± 0.08
Cr3 + 0.30 ± 0.03 - 0.27 ± 0.02
La3 + 0.22 ± 0.02 0.47 ± 0.03 -
La3 + ? Lu3 + - 0.47 ? 0.52h -
UOH3 + - 0.48 ± 0.08j -
UF3 + - 0.48 ± 0.08e -
UCl3 + - 0.59 ± 0.10j -
UBr3 + - 0.52 ± 0.10e -
UNO3þ 3 - 0.62 ± 0.08j -
Be2OH3 + - 0.50 ± 0.05k -
Pu4 + - 1.03 ± 0.05f -
Np4 + - 0.82 ± 0.05 -
U4 + - 0.76 ± 0.06e -
Th4 + 0.25 ± 0.03 - 0.11 ± 0.02
a
Source: Refs. [10,24] unless specified. Uncertainties are 95% confidence level.
b
Ion interaction coefficients can be described more accurately with an ionic strength-dependent
function, listed in Table 6.5.
c
Ferri et al. [26].
d
Evaluated in NEA-TDB review on uranium thermodynamics [37].
e
Estimated in NEA-TDB review on uranium thermodynamics [37].
f
Riglet et al. [36], where the following assumptions were made: "ðNp3þ ;ClO4 Þ "ðPu3þ ;ClO4 Þ ¼ 0:49 as
for other ðM 3þ ; ClO 4 Þ interactions, and "ðNpO2þ " " ¼ 0:46.
2 ;ClO4 Þ ðPuO2þ
2 ;ClO4 Þ ðUO2þ
2 ;ClO4 Þ
g
Hedlund [35].
h
Taken from Spahiu [25].
i
The coefficients "ðMnþ ;Cl Þ and "ðMnþ ;NO3 Þ reported by Ciavatta [10] were evaluated without taking
chloride and nitrate complexation into account. See section 6.3.
j
Evaluated in NEA-TDB review on uranium thermodynamics [37] using "ðU4þ ;ClO4 Þ ¼ ð0:76 0:06Þ.
k
Taken from Bruno [38], where the following assumptions were made: "ðBe3þ ;ClO4 Þ ¼ 0:30 as for other
"ðM2þ ;ClO4 Þ ; "ðBe2þ ;Cl Þ ¼ 0:17 as for other "ðM2þ ;Cl Þ , and "ðBe2þ ;NO3 Þ ¼ 0:17 as for other "ðM2þ ;NO3 Þ .
Table 6.4 Ion Interaction Coefficients ej,k for Anions j with k = Li +, Na+, and K+ a
j k!
# Li + Na+ K+
OH 0.02 ± 0.03b 0.04 ± 0.01 0.09 ± 0.01
F - 0.02 ± 0.02c 0.03 ± 0.02
HF2 - 0.11 ± 0.06c -
Cl 0.10 ± 0.01 0.03 ± 0.01 0.00 ± 0.01
ClO 4 0.15 ± 0.01 0.01 ± 0.01 -
Br 0.13 ± 0.02 0.05 ± 0.01 0.01 ± 0.02
I 0.16 ± 0.01 0.08 ± 0.02 0.02 ± 0.01
HSO 4 - -
0.01 ± 0.02
NO 3 0.08 ± 0.01 0.04 ± 0.03b 0.11 ± 0.04b
H2 PO 4 - 0.08 ± 0.04b 0.14 ± 0.04b
HCO 3 - 0.03 ± 0.02 -
SCN - 0.05 ± 0.01 0.01 ± 0.01
HCOO - 0.03 ± 0.01 -
CH3COO 0.05 ± 0.01 0.08 ± 0.01 0.09 ± 0.01
SiOðOHÞ 3 - 0.08 ± 0.03c -
BðOHÞ 4 - 0.07 ± 0.05b -
UO2 ðOHÞ 3 - 0.09 ± 0.05c -
UO2 F 3 - 0.00 ± 0.05c -
ðUO2 Þ2 CO3 ðOHÞ 3 - 0.00 ± 0.05c -
SO2
4 0.03 ± 0.04b 0.12 ± 0.06b 0.06 ± 0.02
HPO2 4 0.15 ± 0.06b 0.10 ± 0.06b
CO23 0.05 ± 0.03 0.02 ± 0.01
CrO2 4 0.06 ± 0.04b 0.08 ± 0.04b
UO2 ðSO4 Þ2 2 0.12 ± 0.06c -
UO2 ðCO3 Þ2 2 0.04 ± 0.09c -
PO3
4 0.25 ± 0.03b 0.09 ± 0.02
P2 O47 0.26 ± 0.05 0.15 ± 0.05
FeðCNÞ4 6 - 0.17 ± 0.03
UðCO3 Þ4 4 0.09 ± 0.10c -
UO2 ðCO3 Þ4 3 0.08 ± 0.11c -
UO2 ðCO3 Þ5 3 0.66 ± 0.14c -
6
UðCO3 Þ5 0.27 ± 0.15c -
a
Source: Refs. [10,34] unless specified. Uncertainties are 95% confidence level.
b
Ion interaction coefficient can be described more accurately with an ionic strength-dependent
function, listed in Table 6.5.
c
Evaluated in NEA-TDB review on uranium thermodynamics [37].
Table 6.5 Ion Interaction Coefficients e(1,j,k) and e(2,j,k) for Cations j with k = Cl, ClO
4 , and NO3 (First Part), and for Anions j with
+ + + a
266
Cl ClO
4 NO
3
j k!
# e1 e2 e1 e2 e1 e2
NHþ 4 0.088 ± 0.002 0.095 ± 0.012 0.075 ± 0.001 0.057 ± 0.004
Ag + 0.1432 ± 0.0002 0.0971 ± 0.0009
Tl + 0.18 ± 0.02 0.09 ± 0.02
Hg2þ
2 0.2300 ± 0.0004 0.194 ± 0.002
Hg2 + 0.145 ± 0.001 0.194 ± 0.002
Pb2 + 0.329 ± 0.007 0.288 ± 0.018
Li + Na + K+
j k!
# e1 e2 e1 e2 e1 e2
OH 0.039 ± 0.002 0.072 ± 0.006
NO 2 0.02 ± 0.01 0.11 ± 0.01
NO 3 0.049 ± 0.001 0.044 ± 0.002 0.131 ± 0.002 0.082 ± 0.006
BðOHÞ 4 0.092 ± 0.002 0.103 ± 0.005
H2 PO 4 0.109 ± 0.001 0.095 ± 0.003 0.1473 ± 0.0008 0.121 ± 0.004
SO2
3 0.125 ± 0.008 0.106 ± 0.009
SO2
4 0.068 ± 0.003 0.093 ± 0.007 0.184 ± 0.002 0.139 ± 0.006
S2 O2
3 0.125 ± 0.008 0.106 ± 0.009
HPO2 4 0.19 ± 0.01 0.11 ± 0.03 0.152 ± 0.007 0.123 ± 0.016
CrO2 4 0.090 ± 0.005 0.07 ± 0.01 0.123 ± 0.003 0.106 ± 0.007
PO24 0.29 ± 0.02 0.10 ± 0.01
a
Source: Refs. [10,34]. Uncertainties are 95% confidence level.
Grenthe and Wanner
Ionic Strength Corrections 267
where the summation over j covers all anions for the case that i is a cation
and vice versa. Tables of B(I, j) are given by Pitzer and Brewer [14] and by
Baes and Mesmer [3]. The Debye-Hückel term [see Eq. (6.2)] is different
from that in the B-G-S equation. Apart from a slightly different value for A,
the factor Baj has been chosen equal to 1.0 in the P-B equation compared to
1.5 in the B-G-S equation. The B-G-S equation is preferred to the P-B
equation in the critical evaluations of the NEA-TDB Project for the reasons
given in section 6.1.1.
where nM and nX are the numbers of M and X ions in the formula unit
and zM and zX their charges. m is the molality of the solution and n =
nM + nX. In aqueous solutions at 258C and 105 Pa, the following relations
are given [16]:
pffiffiffiffiffi
Im pffiffiffiffiffi
f ¼ 0:392 pffiffiffiffiffi þ 1:667 lnð1 þ 1:2 Im Þ (6.30)
1 þ 1:2 Im
ð1Þ
MX pffiffiffiffiffi pffiffiffiffi
ð0Þ
BMX ¼ 2MX þ (1 (1 þ 2 Im 2Im )e2 Im ) (6.31)
2Im
CMX ¼ 32 CMX (6.32)
[4].
P Here Pa and a0 P cover all anions, c and c0 cover all cations, and
( mz) = cmczc = ama|za|.
X
2M X X
lnMX ¼ jzM zX j f þ ma BMa þ mz CMa þ Xa
a M
X
2X X M
þ mc BcX þ mz CcX þ Mc
c X
XX
0 1
þ mc ma jzM zX jBca þ ð2M zM Cca þ M Mca þ X caX Þ
c a
1 X X X
0
þ mc mc0 cc0 X þ jzM zX jcc0
2 c c0
1XX M
0
þ ma ma0 Maa 0 þ jz z j
M X aa 0 (6.33)
2 a a0
In Eq. (6.33) f g is the same as in Eq. (6.29); CMX, BMX, and its derivative
with ionic strength, B0 MX, have the forms
where b(0), b(1), Cf are the same as for pure electrolytes. In Eq. (6.33) the y
terms summarize the interactions between ions of the same charge sign that
are independent of the common ion in a ternary mixture, and the c terms
account for the modifying influence of the common ion on these interac-
tions. Pitzer [4] points out that higher order electrostatic terms (beyond
the Debye-Hückel approximation) become important in cases of unsym-
metrical mixing, especially if one of the ions has a charge of three or higher.
Higher order electrostatic effects, on the other hand, were found to be
unimportant for cases of symmetrical mixing and for pure unsymmetrical
electrolytes.
Based on the cluster integral method [27], Pitzer [19] divided the dif-
ference terms yMN into two parts
be approximated as
!1
1 2 X 6
JðxÞ ¼ x ðln xÞe10x þ Ck xk
2
(6.40)
6 k¼1
The first term is in fact important only at very low ionic strengths; thus for
cases used in equilibrium analysis one has
4:118 7:247 4:408 1:837 0:251 0:0164 1
JðxÞ ¼ þ 2 þ 3 þ 4 þ 5 þ (6.41)
x x x x x x6
The syMN parameters can be evaluated from data on mixtures of electrolytes
by calculating the differences between the experimental value of gexp and the
value calculated with the appropriate values for all pure electrolyte terms
and Ey terms but zero values for sy and c terms. For the activity coefficient
of MX in a MX-NX mixture one has Eq. (6.42) and equivalent expressions
for other cases.
1 zM
ln ¼ MN þ
s
mX þ mM MNX (6.42)
2 m
M N 2 zX
This is the equation of a straight line with intercept syMN and slope cMNX,
when the left side of Eq. (6.42) is plotted against a function of the com-
position 12 ðmX þ j zzMX jmM Þ.
The Pitzer equation for single electrolytes with the value of parameters
collected in several publication may be used as a compact source of activity
o
coefficient data. From the values g jk thus obtained, one may calculate, for
example, "ðj;k;Im Þ values using Eq. (6.1).
For the estimation of medium effects on solubility equilibria in mix-
tures of electrolytes involving ions of charge less than three, one may neglect
y and c terms in Eq. (6.33).
In equilibrium analysis studies carried out in the presence of an inert
salt (medium salt NX) and small (trace) concentrations of the reactants,
only the terms involving mNX have to be considered in Eq. (6.33), while
those involving mitrace can be neglected. Nevertheless, as the main difficulty
270 Grenthe and Wanner
there still remains the accurate estimation of the parameters of single elec-
trolyte b(0), b(1), and Cf for species such as metal ion complexes.
Equations for single ion activity coefficients [4], osmotic coefficients
[17], and other thermodynamic quantities [28], as well as applications in
different cases (e.g., H2SO4 and H3PO4 solutions) have been given by Pitzer
and coworkers [4,20].
From Tables 6.3 and 6.4 it seems that the size and charge correlations
can be extended to complex ions. This observation is very important because
it indicates a possibility to estimate the ion interaction coefficients for
complexes by using such correlations. It is, of course, always preferable to
use experimental ion interaction coefficient data. However, the efforts nee-
ded to obtain these data for complexes will be so great that it is unlikely that
they will be available for more than a few complex species. It is even less
likely that one will have data for the Pitzer parameters for these species.
Hence, the specific ion interaction approach may have a practical advantage
over the inherently more precise Pitzer approach.
BMX ¼ B1 1
MX þ (BMX BMX )F(Im )
o
(6.44)
pffiffiffiffiffi pffiffiffiffi
1 ð1 þ 2 Im 2Im Þe2 Im
FðIm Þ ¼ ; Fð0Þ ¼ 1; Fð1Þ ¼ 0:
4Im
(6.45)
The Pitzer function linearizes the dependence of the ion interaction coeffi-
cient on the ionic strength quite well, even in the cases of 4:1 and 5:1
electrolytes, where constant e(M,X) values [see Eq. (6.1)] are not obtained at
high ionic strengths. The parameters BoMX and B? MX can be determined from
a single electrolyte activity coefficient datum by calculating first BMX. (Note
that BMX=e( j,k), since the Debye-Hückel term in Eq. (6.43) does not have
the factor 1.5 in the denominator.) By plotting BMX(Im) values against F(Im),
B? o
MX is obtained as the intercept while B MX is obtained from the slope of
the straight line [see Eq. (6.44)]. The equation for a mixture is similar to
Ionic Strength Corrections 271
Eq. (6.43) and BMX = 0 if M and X are of the same charge sign. In the case
of equilibrium constant measurements, DB values are expressed by equa-
tions similar to Eq. (6.6).
The corresponding DBo and DB? can be obtained together with the bo
values from the system of equations
pffiffiffiffiffiffiffiffi
2 0:511 Im;n
log 10 ðIm;n Þ zi pffiffiffiffiffiffiffiffi q log 10 ðaH2 O Þn
1 þ Im;n
¼ log 10 o þ B1 ½1 FðIm;n ÞIm;n þ Bo FðIm;n ÞIm;n (6:46Þ
where b(Im,n) and ðaH2 O Þn refer to the values of b and aH2 O at ionic strength
Im,n. From the values obtained for DBo and DB? and equations similar to
Eq. (6.9), one may estimate the unknown BoMX and B? MX values.
ionic charge and the ionic strength, but it accounts for the medium-
specific properties by introducing ionic pairing between the medium ions
and the species involved in the equilibrium reactions. Earlier, this
approach has been used extensively in marine chemistry (See Refs. [30–
32, 39]).
It can be shown that the virial type of activity coefficient equations and the
ionic pairing model are equivalent, provided that the ionic pairing is weak.
In these cases, it is in general difficult to distinguish between complex for-
mation and activity coefficient variations unless independent experimental
evidence for complex formation is available, e.g., from spectroscopic data,
as is the case for the weak uranium(VI) chloride complexes. It should be
noted that the ion interaction coefficients evaluated and tabulated by Cia-
vatta [10] were obtained from experimental mean activity coefficient data
without taking into account complex formation. However, it is known that
many of the metal ions listed by Ciavatta form weak complexes with
chloride and nitrate ions. This fact is reflected by ion interaction coefficients
that are smaller than those for the noncomplexing perchlorate ion (see
Table 6.3). This review takes chloride and nitrate complex formation into
account when these ions are part of the ionic medium and uses the value of
the ion interaction coefficient "ðMnþ ;ClO4 Þ for "ðMnþ ;Cl Þ and "ðMnþ ;NO3 Þ . In
this way, the medium dependence of the activity coefficients is described
with a combination of a specific ion interaction model and an ion pairing
model. It is evident that the use of NEA-recommended data with ionic
strength correction models that differ from those used in the evaluation
procedure can lead to inconsistencies in the results of the speciation calcu-
lations.
It should be mentioned that complex formation may also occur
between negatively charged complexes and the cation of the ionic medium.
An example is the stabilization of the complex ion UO2(CO3)35 at high ionic
strength.
6.5 CONCLUSION
The specific ion interaction approach is simple to use and gives a fairly good
estimate of activity factors. By using size/charge correlations, it seems pos-
sible to estimate unknown ion interaction coefficients. The specific ion
interaction model has therefore been adopted as a standard procedure in the
NEA Thermochemical Data Base review for the extrapolation and correc-
tion of equilibrium data to the infinite dilution standard state. For more
details on methods for calculating activity coefficients and the ionic medium/
ionic strength dependence of equilibrium constants, the reader is referred to
Ref. 40, Chapter IX.
REFERENCES
1. Lafitte, M. J. Chem. Thermodyn., 1982, 14, 805.
2. Wanner, H. Standards and conventions for TDB publications, TDB-5.1, Rev. 1,
OECD Nuclear Energy Agency, Data Bank, Gif-sur-Yvette, France, 1989.
3. Baes, C. F., Jr.; Mesmer, R. F. The Hydrolysis of Cations; John Wiley and Sons,
New York, 1976.
4. Pitzer, K. S. Theory: Ion interaction approach. In Activity Coefficients in
Electrolyte Solutions, Pytkowicz, R. M., Ed., Vol. I, CRC Press, Boca Raton,
Florida, 1979; 157–208.
5. Brønsted, J. M. Studies of solubility: IV. The principle of specific interaction of
ions. J. Am. Chem. Soc., 1922, 44, 877–898.
6. Brønsted, J. M. Calculation of the osmotic activity functions of univalent salts.
J. Am. Chem. Soc., 1922, 44, 938–948.
7. Scatchard, G. Concentrated solutions of strong electrolytes, Chem. Rev., 1936,
19, 309–327.
8. Guggenheim, E. A. Applications of Statistical Mechanics; Oxford, Clarendon
Press, 1966.
274 Grenthe and Wanner
7.1 INTRODUCTION
The development of a solvent extraction process from a given aqueous feed
solution may be confined by several restraints. For example, the temperature,
flow rate, acidity, and so forth may be (essentially) fixed, and the solvent
extraction process developed must then be capable of accommodating these
restrictions.
The initial bench-scale experimental investigations into solvent extrac-
tion processes are conducted with small apparatus, such as separating fun-
nels. Following the successful completion of these tests, when the best reagent
and other conditions for the system have been established, small-scale con-
tinuous operations are run, such as in a small mixer-settler unit. The data so
obtained are used to determine scale-up factors for pilot plant or plant design
and operation (see Chapters 7 and 8).
The need for pilot plant operations is considered by one school of
thought to be an unnecessary expense in the development of a solvent pro-
cess. The rationale here is that monies and time spent in piloting the process
can be used to modify, if necessary, the full-scale plant. Thus, if few (or no)
modifications are required, considerable savings can result. This approach
is probably satisfactory if the process is very similar to an existing process
and sufficient data are available on which to base the plant design and
operation. The practical experience of the plant design engineers and oper-
ators is also of considerable importance in such cases. On the other hand,
problems encountered in a plant operation that has been designed without
piloting the process could result in the loss of considerable time and expense.
277
278 Ritcey
Of course, almost any process or plant can be made to work, even though the
basic data and design are poor. The costs, however, may be astronomical! It
should be pointed out, however, that the size of a pilot plant could vary from
a few gallons per minute of total throughput to several hundred gallons per
minute. Normally, one would expect that the size of a pilot plant would be in
direct proportion to the size of the final plant.
This chapter comprises some selected excerpts taken from a com-
prehensive two-volume text on the subject coauthored by the author [1]. It
specifically addresses the extraction of metals, but many of the (general)
conclusions are equally valid for the extraction of organic compounds, with
appropriate modifications.
Having formed the amine salt, the solvent can be used for the extraction of
uranium. This is shown as an ion exchange process in Eq. (7.3):
2[(R3 NH)2 SO4 ]ðorgÞ +UO2 ðSO4 Þ4
3
in which the anionic sulfate moiety of the amine salt (organic phase) is
exchanged for the anionic uranyl sulfate species (aqueous phase), to form a
uranyl amine sulfate complex as the extracted species (anion exchange).
Stripping the uranium from the solvent can be accomplished by using
either acid or alkaline solutions. If an alkali carbonate solution is used, the
stripped solvent then requires equilibrating with sulfuric acid before recy-
cling to the extraction stage. Sulfuric acid stripping obviates the need for
such equilibration.
Acidic extractants that undergo a cationic exchange reaction with
metal ions do not generally require equilibration before extraction, because
the exchange involves protons that are already present in the extractant:
Mnþ +nHXðorgÞ $ MXnðorgÞ +nHþ (7.4)
and the solvent phase is then equilibrated with ammonia [Eq. (7.5)] for
recycling to the extraction stage.
Commercially produced extractants are not pure and always contain
some residual starting materials, by-products formed during manufacture, or
degradation products. For example, DEHPA contains some 2-ethylhexanol
(starting material) and some of the monoester acid. Tertiary amines invari-
ably contain primary or secondary amines, and so on. If the impurities in
the extractant used in the solvent are more soluble than the extractant it-
self, which is usually the case, they will be lost from the system to the aqueous
phase after some recycling of the solvent. The concentration and solubility of
such impurities will depend on the time required to remove them from the
system. The problems associated with such impurities become particularly
evident if their effects are not considered during bench-scale investigations.
Thus, their presence in a solvent system can produce cruds, good or poor
phase separation, and enhanced or poor loading characteristics. Such effects
could result in abandoning a particular solvent because of poor chemical and
physical characteristics; conversely, problems could arise when the process is
tried on a continuous scale if the enhancing effects of the impurities result in
poor extraction characteristics as a result of their loss to the raffinate. It is
suggested, therefore, that all solvents be conditioned before use in bench-
scale studies that will ultimately be developed into a continuous system.
Although there is no specific method of conditioning, a rule of thumb
is to treat the freshly prepared solvent with a solution similar to that used in
the test work. This may require several contacts with the aqueous solution,
with stripping between contacts if a metal is extracted.
Inclusion of this practice in all solvent extraction studies will ensure
that the solvent is not discarded as being unsuitable, and that problems in
pilot plant or continuous operations are not the result of impurities that were
present in the bench-scale test work.
It should also be pointed out that extractant quality might vary from
sample to sample. This can be particularly frustrating in both development
282 Ritcey
work and pilot plant operations. It is good practice to test every batch of
extractant before use, to ensure that its characteristics are similar to the
material previously used. Such testing will also provide the investigator with
the knowledge that the reagent is as ordered, and is not another reagent
entirely. It is not uncommon for the wrong material to be shipped.
Similar comments can be applied to diluents and modifiers. The inves-
tigator should be aware of the problems that can arise from impurities in the
components that go to make up a solvent.
Fig. 7.1 Distribution isotherms for nickel at pH 6 for three different concentrations
(vol%) of DEHPA(Na), in the kerosene-aqueous sulfate systems.
phases are again contacted until equilibrium is obtained, and the same
procedure performed as just described. This process is repeated until sat-
uration of the solvent with the metal is obtained. Here again, care must be
taken to maintain the same pH throughout the procedure. One drawback
to this method is that a large volume of the organic phase must be used
initially to allow for sampling, especially if numerous contacts are required.
Generally, the first method employing variation in the phase ratio is
recommended.
The extraction isotherms constructed from these data can be used in
the construction of a McCabe-Thiele diagram (see Chapter 8).
The loading capacity of the solvent can be obtained from either of
these two methods for the particular concentration of extractant used. It
cannot be assumed that the loading capacity is linear with increasing
extractant concentration (see Fig. 7.1). If high concentrations (>10 vol%)
are to be used, it is advisable to determine the loading capacities at these
concentrations. This is readily done either by contacting a solvent several
times with fresh aqueous solution, or by using a concentrated solution of
284 Ritcey
7.4.2 Stripping
Data required for the construction of stripping isotherms are obtained in a
manner similar to those for extraction. A general procedure is as follows:
the loaded solvent is contacted with a suitable strip solution (e.g., acid,
base, etc.) at an appropriate phase ratio until equilibrium is attained. The
aqueous phase is then removed, fresh strip solution is added to the organic
phase, and the procedure repeated. This process is continued until all (or as
much as possible) of the metal has been stripped from the organic phase.
Analysis of the aqueous strip liquors, if no volume changes occur, allows
the calculation of the concentration of metal in the organic phase after
each strip. However, if volume changes do occur, then the organic phases
must also be analyzed. Then the stripping isotherm can be drawn and a
McCabe-Thiele diagram constructed to determine the number of theore-
tical strip stages required at the phase ratio and strip solution concentra-
tion used.
7.4.4 Scrubbing
The object of scrubbing a loaded solvent is to remove as much as possible of
any unwanted coextracted metal. There are many possibilities of accom-
plishing this by varying the scrub solution. Normally, however, scrubbing is
achieved by water, dilute acid, or base, or by an aqueous solution of a salt of
the metal of primary interest in the solvent phase. Scrubbing tests are carried
out in a manner similar to those described earlier with the loaded solvent
being contacted with the scrub solution at the appropriate concentration, pH,
phase ratio, contact time, and temperature. After separating the phases, the
organic phase is usually analyzed to determine whether the unwanted metals,
and how much of the metal of interest, have been removed. The best scrub
solution and conditions are determined by varying the individual conditions.
The scrub raffinate may contain substantial amounts of valuable
metals; thus, consideration should be given to where in the overall process
this raffinate should be recycled.
Data obtained in scrubbing tests can be presented graphically, such as
by plotting the concentration of metals in the organic phase against A/O
ratio, salt concentration in the scrub solution, temperature, etc. Typical plots
for the removal of nickel from cobalt in a DEHPA-containing solvent are
shown in Figs. 7.3 and 7.4 [3].
Fig. 7.3 Effect of scrub solution composition on the removal of nickel from cobalt in a
DEHPA-kerosene solvent.
Fig. 7.5 The effect of temperature on the loading and separation of cobalt and nickel.
Solvent: 15 vol% DEHPA; 5 vol% TBP; 80 vol% kerosene.
both these metals. Analysis of the solvent after each contact stage, and
plotting these concentrations against the number of stages, will give a curve
for each metal similar to those shown in Fig. 7.6. Without completing the
curves, a good approximation of the number of stages required to provide
a given Co/Ni ratio can be obtained by extrapolation.
In the example shown in Fig. 7.6, using an alkylphosphoric acid, about
60 stages are required to give a Co/Ni ratio of about 100. So many stages
would be too many for mixer-settler operation, and other types of con-
tactors would have to be considered. In this particular example, a sieve plate
pulsed column has been shown to be very effective [3]. However, with the
development of the alkylphosphonic and alkylphosphinic acids, the sep-
aration of cobalt and nickel can be achieved in very few stages, owing to the
high rejection of nickel (see Chapter 11).
It is a good plan to begin solvent extraction studies involving several
metals with solutions containing only single metals to obtain the basic
extraction data. It also provides time for any necessary analytical develop-
ment to be carried out. However, synthetic solutions should not be worked
on longer than necessary, as the actual plant liquor will almost certainly be
much different. Another point to be made here is that in leaching, filtering,
precipitation, and flotation, chemicals are added to aid these processes.
Industrial Solvent Extraction Processes 291
These chemicals can occasionally cause adverse effects, such as poor phase
disengagement, during solvent extraction (see section 7.13).
more like rigid spheres, and so the rate of mass transfer decreases. Thus, the
kinetics, along with the proper dispersion and coalescence, will influence the
choice of contacting equipment. In addition to drop size, other factors
affecting coalescence include the temperature, viscosity, and density of the
phases, and the presence of surface-active agents, or solids. The addition of
wetting materials within the contactor or phase separation equipment will
promote the coalescence rate.
Coalescence can be divided into primary and secondary break time.
Primary break time is the time required for the two phases to meet at
a sharply defined interface and provides a measure of the settler or phase
separation requirement of the process. The secondary break concerns the
fine haze of droplets and is the major cause of entrainment losses. The speed
and completeness of phase disengagement have a marked effect on the
capital and operating costs of the plant and are directly linked to the design
and method of operating of the mixer. In general, but only to a point, higher
extraction efficiencies are achieved as the speed of the mixer is increased and
the resulting droplet size of the dispersion is reduced. However, the dis-
persion band thickness increases as the droplet size of the dispersion is
reduced, and a larger settler area (with consequently larger inventories of
solvent) are then required for a given plant throughput. Figure 7.7 shows the
variation of dispersion band depth with the total flow/unit settler area under
varying conditions of mixing, expressed by the term nR3d 2 where nR is the
speed of the impeller (rev/s) and d its diameter (ft) [6].
Increased capital costs will be incurred if the mixer is designed to
produce a high proportion of extremely small droplets, and this cost has to
be balanced against the increased revenue obtained from the improved
extraction efficiency. At upper levels of N3D 2, the problem of organic carry-
over with the aqueous phase may become an important consideration.
In addition to the directly calculable increase in settler costs when operat-
ing with very highly dispersed phases, the possibilities of emulsion forma-
tion and flooding caused by the accumulation of ‘‘crud’’ in the settler must
be considered. Both of these become more likely as the settler duty be-
comes more severe. Many authors have cited design ideas on mixer settlers
[7, 8].
In all practical mixing systems, a range of droplet sizes is produced by
the mixer, and there will always be a proportion of droplets of the dispersed
phase that will not settle out readily and, therefore, will remain entrained in
the continuous phase leaving the settler. Impellers that consist of enshrou-
ded radial vanes create very high shear rates within the mixer, particularly
when they are run close to the bottom of the mixing tank to provide a
pumping as well as a mixing action. These high-shear rates are the prime
cause of the haze of fine droplets produced in many plants. Thus, the design
294 Ritcey
Fig. 7.7 Effect of variation of mixer N3D2 on settler dispersion band depth.
Nonagitated Agitated
Mixer-settlers differential differential Centrifugal
Advantages Good contacting of Low initial cost Good dispersion Handle low gravity
phases Low operating cost Reasonable cost difference
Handle wide range flow Simplest construction Many stages possible Low holdup volume
ratio (with recycle) Relatively easy scale-up Short holdup time
Low headroom Small space
High efficiency requirement
Many stages Small inventory of
Reliable scale-up solvent
Industrial Solvent Extraction Processes
Low cost
Low maintenance
Disadvantages Large holdup Limited throughout Limited throughout High initial cost
High power costs with small gravity with small gravity High operating cost
High solvent inventory difference difference High maintenance
Large floor space Cannot handle wide Cannot handle emulsifying Limited number of
Interstage pumping may flow ratio systems stages in single unit,
be necessary High headroom Cannot handle high although some units
Sometimes low flow ratio have 20 stages
efficiency Will not always handle
Difficult scale-up emulsifying systems,
except perhaps pulse
column
297
298 Ritcey
having up to about 500 cm3 capacity for each stage has been used. Much of
the early work employed homemade equipment, and even now much of the
equipment is constructed in the laboratory.
Several designs of small-scale mixer-settler units can be obtained com-
mercially and can be assembled to provide several extraction, scrub, and strip
stages. Regular turbine blades are used for pumping and mixing; thus the
mixer-settlers are horizontal and the cascading principle is not necessary. The
major problem with small-scale equipment made of Plexiglas is that it tends
to craze quite rapidly and discolor. Consequently, it becomes difficult to
observe accurately the interface level or the dispersion band in the settler.
However, the ease with which mixer-settler units can be made out of Plexiglas
means that they can be discarded and replaced by new ones when such
conditions occur.
Mixer-settlers are the smallest continuous countercurrent equipment
available. Other types of contactors can be purchased or constructed, but
require larger volumes of solutions for operation. Hence, a Karr column is
available having a 1 inch diameter column. Similarly, a sieve plate pulse
column of similar dimensions can be readily constructed. However, scale-
up from small diameter columns such as these is not too satisfactory
[G. M. Ritcey, unpublished data]. The major problem appears to be that wall
effects in small-diameter columns significantly influence both the physical
and chemical characteristics of a solvent extraction system. Thus, a 2 inch
diameter sieve plate column appears to be the minimum size for which wall
effects have a minimum effect on the system. Other columns, with greater
axial and backmixing, require a larger diameter column for scale-up.
Whether a 2 inch diameter column can be considered small-scale or pilot
plant size is debatable as the total flow capacity of such a column can be up to
0.5 US gal min1 (1.8 dm3 min1). The size of a pilot plant, as we have noted
before, covers a wide range of throughput.
Other small-scale contacting equipment is available, e.g., cen-
trifuges and the ‘‘raining bucket’’ or RTL contactor. The minimum
flow rate of available centrifuges seems to be about 1 gal (US) min1
(3.6 dm3min1). A small-scale raining bucket contactor can be obtained that
is suitable for bench-scale operations. The 4 inch (10 cm) diameter, 30 inch
(0.76 m) long unit would have a maximum throughput of about 1 US gal hr1
(3.6 dm3 hr1).
In-line mixers manufactured by, for example, Kenics, Lightning, and
Sulzer are also applicable for continuous small-scale testing of a solvent
extraction process, and 1 inch diameter models are available. This mixer
system can be used either horizontally or vertically. However, few data are
available for this type of contactor, although they would appear to offer
many possibilities, not only for liquid–liquid systems, but also for use in
Industrial Solvent Extraction Processes 299
solvent-in-pulp processes (in the case of perhaps the Kenics and Lightning
mixers) [G. M. Ritcey, unpublished data].
Because of the diversity of contacting equipment available, it is unlikely
that all these contactors will be available in any one laboratory or pilot plant.
Consequently, unless test work is carried out on similar contactors, the
system may not be optimized. Since mixer-settlers are the easiest to construct,
are simple to operate, and require little room and low-flow rates, these con-
tactors are, in many cases, the only ones used to investigate a continuous
solvent extraction process. This is by no means ideal and may result in
abandonment of a process that, using another type of contactor, could be
found to be entirely satisfactory.
The data obtained from small-scale continuous operations will be
required to determine whether the process should be investigated on a larger
scale, such as a pilot plant. These data should be sufficient to draw a con-
ceptual flow sheet, which will include a number of stages for extraction,
scrubbing [10], and stripping; the flow rates, size and type of equipment, and
the various parameters considered earlier in this chapter. Another important
aspect should also be considered in the continuous test work, and that is
chemical analysis of both the aqueous and organic phases for their various
components.
companies who have experience in this field, or to have the pilot work per-
formed by the manufacturers of the contactors chosen for the operation.
It will be instructive to consider briefly the methodology employed
in the development of solvent extraction processes that have become
operational. The development of the Bluebird Mine operation for the
extraction of copper from dump leaching liquors by solvent extraction and
the subsequent recovery of copper as either copper sulfate or cathode copper
is used as an example. The initial investigations [11] included the following:
1. Bench-scale batch tests to determine the characteristics of the extraction
of copper from sulfuric acid liquors by LIX 64
2. Continuous bench-scale operations to study the extraction of copper
and other metals present in the leach liquor: effects of variables such as
temperature, O/A ratio, solvent composition, extraction kinetics, set-
tling characteristics, etc.
3. Determination of solvent losses from solubility, entrainment, and chem-
ical degradation
4. Stripping characteristics and the production of copper sulfate crystal,
saturated copper sulfate solution, powdered copper, or electrolytic
copper from the strip liquors
Following the successful completion of this test work, pilot plant operations
were initiated, having the following objectives:
1. To establish the economic feasibility of producing wire bar-grade cop-
per by a combined SX–EW (EW=electrowinning) process
2. To develop the engineering data to enable estimates of operating and
capital costs of a process capable of producing approximately 30,000 lb
(13,600 kg) of copper per day
3. To develop the engineering data required for the design and operation
of the commercial plant
4. To produce electrolytic copper for evaluation of product quality
Other objectives of the pilot plant operation include those given for bench-
scale and small-scale continuous studies, together with the following:
Fig. 7.9 Solubility of C7 – C9 aliphatic carboxylic acids in water. (From Ref. 14.)
Fig. 7.10 Solubility of Versatic 911 (0.50 mol dm3) in ammonium sulfate solution.
(From Ref. 15.)
Fig. 7.11 Solubility of DEHPA in various aqueous solutions. (From Ref. 17.)
Fig. 7.12 Solubility of the sodium salt of DEHPA in NaOH (wt%) solution at 208C.
308 Ritcey
Fig. 7.13 Solubility of the sodium salt of DEHPA in water as a function of temperature.
Fig. 7.14 Solubility of amines in acidic sulfate solution at (a) pH 1.4 and (b) pH 1.8.
310 Ritcey
Blake et al. [17], Table 7.2. Aqueous solubility decreases with increase in
the number of carbon atoms and with a decrease in concentration in the
organic phase. This observation is general and has been found to apply in
other cases, for example, for C8–C9 carboxylic acids, for which the solubility
is almost a linear function of the carboxylic acid concentration in the
organic phase [25]. (see also Chapter 2).
7.13.6 Volatilization
Volatilization of solvent components can become a problem when the sys-
tem is operated at elevated temperatures or in hot climates. The human
toxicity of solvent components is a generally unknown factor and could be a
problem in a system enclosed in a building.
Generally, one would expect that the most volatile component would
evaporate first, and this would probably be the diluent. In several cases of
operating simulated solvent extraction processes at temperatures up to
708C, it has been noted that the diluent is rapidly volatilized [G. M. Ritcey
and B. H. Lucas, unpublished data]. Problems of volatilization appear not
to have occurred to any great extent in the past (perhaps the losses were not
measured), but any trend to the use of elevated temperatures would require
that this form of solvent loss be thoroughly investigated.
7.13.7 Entrainment
Entrainment of solvent in the aqueous raffinate, in the strip liquor, or in
cruds can be the most serious cause of solvent loss. Here, it is expected that
the ratio of solvent components in the entrained material will be the same as
in the solvent. Entrainment losses can occur because of insufficient settling
area or time allowed for phase disengagement; poorly designed or operated
mixers; too much energy input into the mixing stage; lack of additives
to suppress emulsion formation; poor diluent choice; high extractant con-
centration in the solvent; solids in the aqueous feeds causing crud formation,
etc. Entrainment losses can be minimized by the incorporation of various
devices in the circuit, such as skimmers, centrifuges, coalescers, after-
settlers, flotation, foam fractionation, activated carbon, or others. Although
Industrial Solvent Extraction Processes 313
flotation units are effective for the removal of entrained solvent and are not
affected by the presence of solids, they are expensive to operate as regards
reagents and power, whereas packed coalescence beds, although relatively
inexpensive and effective, can be easily blocked by solids. Baskets contain-
ing polypipe pieces tend to be coated with a thin film in a reasonably short
time, perhaps less than a month, and thus require expensive periodic main-
tenance. Centrifuges, although capable of high throughput, are expensive,
have high power consumption, and may require frequent cleaning.
7.13.8 Crud
Common to all or most solvent extraction operations in the mining industry
is the problem of stable formation of cruds. The crud can constitute a major
solvent loss to a circuit and thereby adversely affect the operating costs.
Because there can be many causes of crud formation, each plant may have
a crud problem unique to that operation. Factors such as ore type, solution
composition, solvent composition, presence of other organic constituents,
design and type of agitation all can adversely affect the chemical and phys-
ical operation of the solvent extraction circuit and result in crud formation
[32–34].
Crud is defined as the material resulting from the agitation of an
organic phase, an aqueous phase, and fine solid particles that form a stable
mixture. Crud usually collects at the interface between the organic and aque-
ous phases. Other names that have been used for the phenomenon are grun-
gies, mung, gunk, sludge, and others.
The following section covers the general aspects of solvent losses by
crud; its formation and characteristics and its treatment and prevention.
Losses attributed to emulsion and crud formation, can in part be related
to (1) nature of feed; (2) reagent choice; (3) equipment selection; and
(4) method of operation, such as the droplet size, continuous phase, ex-
cessive turbulence, etc.
7.13.8.1 Possible Causes of Crud Formation
7.13.8.1.a Nature of Feed
The nature of the feed composition can be a major determining factor as to
whether crud will form in the subsequent extractive operations [33,34]. The
presence and concentration of certain cations, such as Fe, Si, Ca, Mg, or Al,
with sufficient shear in the mixing process can produce stable cruds [32,34].
Solids must be absent from most solvent extraction circuits, and clarification
is usually aimed at achieving about 10 ppm of solids. One of the major causes
of crud is the lack of good clarification of the feed solution, with the result
that solids get through to the solvent extraction circuit. The presence of
314 Ritcey
colloids, such as silica [35], can produce stable emulsions and crud during
the mixing of the phases to achieve mass transfer. Aged feeds can constitute
a greater potential crud problem than fresh leach solutions [36]. In plants
where bacteria have been prevalent because of favorable environmen-
tal conditions, crud has resulted, and expensive circuit modifications
were subsequently required. The elimination of air to such circuits is often
necessary to minimize bacterial and fungal growth [37]. Certain systems may
have hydrolyzed compounds precipitating out of solution, and thus a crud
results. In certain extraction systems, the anionic strength of the aqueous feed
solution may be insufficient, so that stable emulsions occur when the two
phases are mixed. If sufficient agitation is applied over a period, then crud
can result. One other important cause of cruds in solvent extraction plants is
dust from the air if drawn into the agitation in a mixer-settler circuit. Thus,
vessels should be covered to prevent dust accumulation. Organic matter in
the feed, such as lignin or humic acids, may also promote crud formation.
7.13.8.1.b Nature of Solvent
The choice of the extractant and solvent composition is an important aspect
of the successful solvent extraction operation, but the possibility of crud
due to the solvent composition must not be overlooked. Many systems
require a modifier to improve phase disengagement, to assist in solubilizing
the metal-organic species, and to reduce third-phase and emulsion tendency.
If a solvent has the tendency to produce emulsions on mixing with the
aqueous feed solution, which could cause cruds if colloids or suspended
solids are present in the aqueous feed, then the cause may be due to several
factors. Perhaps the system requires the addition of a modifier, a change to a
different modifier, or a higher modifier concentration. Also, possibly the
diluent type and composition may not be compatible with the system. An
aromatic diluent or an aliphatic diluent with some aromatic content may be
more desirable than a completely aliphatic diluent for that particular pro-
cess. Frequently, in the solvent makeup, there are unreacted chemicals from
the manufacturing process or, possibly, impurities from the containers used
to transport the solvent components. The problems associated with such
impurities become particularly evident if their effects are not considered
during bench-scale investigations [1]. Their presence in a solvent system can
produce cruds, good or poor phase separation, and enhanced- or poor-
loading characteristics. Such effects could result in abandoning a particular
solvent because of its poor chemical and physical characteristics.
Degradation of the solvent due to the presence of certain metals in the
feed solution, use of oxidizing agents during stripping, high-temperature
processing, and biodegradability all may result in decomposition products
forming stable emulsions and cruds. Several uranium plants have reported
Industrial Solvent Extraction Processes 315
operation [49]. New plants are attempting to achieve this objective by the use
of Enviroclear thickeners, leaf clarifiers, and sand filters. Although sand
filters are used successfully in some plants, other plants have not been able to
achieve their objective. Insufficient frequency of backwashing the sand filters
is a probable cause of poor clarification.
Because dust can cause crud if permitted in a mixer-settler circuit,
particularly if the settlers are located in the open, adequate covers over the
settlers should be provided.
Colloidal silica in some circuits such as TBP–HNO3 in uranium repro-
cessing can be reduced by the addition of gelatin and heating to 808C to
coagulate, followed by centrifuge separation [33]. The addition of certain
surfactants to lower the surface tension may also reduce cruds caused by
colloidal silica [47], as well as break emulsions. Addition of sequestering
agents was successful in eliminating the deposition of calcium–rare earth
precipitate, in a rare earth–DEHPA circuit [50]. However, the addition of
such reagents could also cause solvent degradation with continual cycling,
so the approach must be carefully investigated before adoption. Surfactants
and other compounds used to enhance the liquid–solid separation after
leaching could also enhance emulsion and crud tendency.
Certain organic constituents, such as lignins and humic acids, may be
solubilized in the aqueous feed and thus cause problems in the solvent
extraction circuit [33]. Very little is yet known about this particular area of
crud formation, but it has been found in phosphoric acid circuits that
coagulation of humates with surfactants followed by filtration has been
useful [51]. Lignin can be removed by passing the leach solution through a
bed of activated carbon, and pressure treatment under oxygen at 200–2508C
will destroy the lignin components.
Crud caused by bacterial and fungal growth may be minimized by
elimination of as much air as possible to the system [52]. These growths, in
320 Ritcey
association with any solids present in the feed liquor, lead to the formation of
crud. Fungal growth in some circuits, arising from the isodecanol modifier in
an amine-uranium system, was eliminated by using 35% Solvesso 150 aro-
matic diluent [53,54]. The aromatic diluent acts as a bactericide and fungi-
cide. Commercial bactericides have also been used successfully.
Proper selection of the extractant as well as the other solvent compo-
nents can minimize emulsions and crud formation. Freshly prepared solvents
can often contain impurities that could subsequently cause operational
problems. Therefore, all solvents should be conditioned before use. Having
selected a suitable solvent system, cyclic tests should be performed to deter-
mine whether degradation of any of the solvent components is taking place,
as any degradation products could cause crud formation in the circuit.
There are several items for consideration in the operation of the circuit
to minimize crud formation. In certain systems, solvent saturation can result
in the formation of gelatinous solids, as in the rare earth–DEHPA system
[50,55] and in the zirconium–TBP circuit [56]. The phenomenon is partially
due to the increase in viscosity as loading is reached. An increase in the O/A
ratio thus results in a decrease in crud formation. As the viscosity increases,
excessive agitation can produce stable emulsions.
Flow patterns can be altered by change of the continuous phase and,
accordingly, the tendency for the formation of emulsions or cruds is altered.
At one uranium refinery, the solvent is maintained in the continuous phase
to produce flow patterns to reduce emulsion tendency [57].
Occasionally, the water used for solution makeup to the scrub or strip
circuits, because of impurities, can cause subsequent emulsions and cruds.
At one uranium plant in South Africa, deionized water was used to prepare
the ammoniacal strip solution, rather than normal plant water, which tended
to cause crud formation [53,54].
Equipment selection is important, as is also the proper operation of
the contacting devices. It is generally recognized that high shear is the pri-
mary cause of droplet haze and subsequent emulsion and crud formation.
Thus, the type and amount of agitation (shear) must be optimized for mass
transfer while minimizing emulsion and crud formation. Table 7.5 indicates
items of information that may be required in analysis of crud formation
problems in a plant [32].
Table 7.5 Possible Information Required for Consideration in Solving Plant Crud
Problems
Ore: mineralogy and analysis
Leach: oxidant and quantity added:
Possible effects on degradation, emulsion production,
or crud stabilization
Liquid–solid separation:
Type of separation (e.g., CCD)
Type and quantity of surfactant added
Use of sand filters or other types of clarifiers (e.g., anthracite) and
frequency of regeneration
Feed solution to solvent extraction:
Suspended solids
Dissolved solids
Solution composition (especially silica, aluminum,
molybdenum, zirconium)
Presence of humic acids, lignin
Solvent extraction:
Extraction circuit:
Extractant
Diluent
Modifier and concentration
Mixer design (e.g., baffling)
Agitation (rpm and design, energy)
Vortex in mixer
Continuous phase
Degradation of diluent, modifier and extractant
(and surfactant from L/S separation)
Settler design (entry to dispersion band,
flow rate design, baffling)
Velocity across settler
Viscosity, surface tension of interfacial tension
Presence of fungus or bacteria
Presence of precipitates
Stripping circuit:
Mixer and settler operations as in ‘‘extraction’’ above
Stripping agent
Viscosity, surface tension, or interfacial tension
Continuous phase
Degradation of diluent, modifiers, and extractant
Presence of fungus or bacteria
Presence of precipitates
322 Ritcey
Guppies Minnows
Aliquat 336 0.18
Primene JMT 0.70
Kelex 100 1.10
LIX 63 4.0 1.6
Alamine 336 10.0 4.0
Isodecanol 12.0 8.4
LIX 64N 15.0 2.7
TBP 18.0 9.6
LIX 70 32.0 15.0
Versatic 911 102.5
DEHPA 173.0
Acute toxicity tests are reported for the Cyanamid reagent Cyanex
272, showing TLm values of 4.9 g kg1 and >2.0 g kg1, respectively, for rats
(oral) and rabbits (dermal) [73]. The TLm 96 hr tests for bluegill sunfish and
rainbow trout gave values of 45 and 22 ppm, respectively.
Tests by Mobil Oil Corporation [74] on dibutyl butylphosphonate
showed that its toxic effects on rats were moderate at 3 g kg1 (oral), and
for rabbits it was nontoxic at >5 g kg1 (dermal) exposure. With DEHPA,
acute toxicity was found for rats at concentrations of 1.4 g kg1 (oral).
In most countries, the manufacturer now has to provide such toxicity
data for approval before marketing.
7.14 ECONOMICS
Any successful solvent extraction process depends upon the selection of
inexpensive extractants that can operate at the natural condition of the
solution, with minimum loss of the organic phase to the aqueous solution.
Also, an inexpensive means of recovery of the metal from the organic
phase is necessary. In some cases, the cost of neutralization of the feed
solution before solvent extraction processing or the cost of maintaining a
buffered pH during extraction may prove excessive when coupled with
solvent losses.
The economics of the solvent extraction process are very dependent
upon ‘‘upstream’’ and ‘‘downstream’’ portions of the plant. Integration of
the total processing step is essential to obtain maximum return. Variables,
such as tonnage rates and changes in solution composition, can have a most
significant result on the economics of solvent extraction. Generally, economic
considerations may be divided into two major areas: (1) capital investments
and (2) operating costs.
house the equipment and, therefore, the capital costs of the process can be
decreased.
Although the capital cost depends on such factors as flow rates,
solvent inventory, equipment, and building, it should not be assumed that
for a given chemical process involving solvent extraction, the capital costs
will be the same in one area of the world as in another. In one area,
something like three plants were installed for an identical process, the
capital cost varying as much as 300% from the lowest to the highest [76].
For any given process, there are many possible designs for the plant,
incorporating different types of equipment and instrumentation that result
in variations in capital cost. Therefore, it must be emphasized that (1)
careful selection of equipment in relation to adherent costs of solvent and
building, and (2) critical evaluation of the engineering contractor’s design
are necessary to optimize the lowest possible capital cost of the solvent
extraction operation.
In summary, capital costs can be influenced by the following:
1. The number of stages and, thus, the number of mixer-settlers or col-
umns required
2. Kinetics and, therefore, the size of the mixing device
3. Coalescence and, therefore, the settling requirements, which dictate
equipment and solvent inventory costs
4. Entrainment and, therefore, the cost of solvent recovery equipment
5. Flow rates and flow ratios
6. Building requirements
7. Ancillary equipment, such as pumps, piping, instrumentation, etc.
8. Equipment to meet environmental regulations
9. Engineering design
where
A is the feed rate of the aqueous phase
Cc is the cost of copper
Xc is the concentration of copper in the raffinate
Sl is the solvent loss per litre
Xl is the volume concentration of LIX 64N
Cxl is the cost of LIX 64N
Csol is the cost of diluent
The capital costs depend on the (1) number of stages of extraction and
stripping, (2) concentration of LIX 64N, (3) size of stages, and (4) ratio of
flow rates.
The size of settlers is determined by the rate of phase disengagement,
which, here, is a function of the solvent concentration of LIX 64N and the
aqueous phase feed rate. The capital cost of extraction equipment is a
function of the number of tanks, the size of the tanks, and the solvent
inventory [9]:
A A
CostðequipÞ =NKl b+ Xl Cxl +(1 Xl )Csol (7.9)
S ex NV
where
N is the number of tanks
Sex is the design settling rate
328 Ritcey
stirring speed, stirring time, and phase ratio. From this data, the stage
efficiency in a flow system may be estimated. Optimization of these para-
meters can result in providing sufficient data for the most economic design
of large-scale mixer-settlers.
The design criteria necessary for optimization of the engineering and
design of the solvent extraction process can probably be summarized as
follows:
1. Clarification of feed liquors
2. pH adjustment and possible filtration
3. Heating requirements
4. Nominal feed flow rate
5. Solution grade expected
6. Production capacity demanded
7. Concentration of solvent
8. Selection of diluent and possible modifier
Industrial Solvent Extraction Processes 331
REFERENCES
1. Ritcey, G. M.; Ashbrook, A. W. Solvent Extraction, Principles and Application
to Process Metallurgy, Parts 1 and 2; Elsevier Science Publishers, Amsterdam,
(1979, 1984).
2. Slater, M. J.; Ritcey, G. M.; Pilgrim, R. F. Proc. ISEC’74, Lyon, Society of
Chemical Industry, London, 1974, 1, 107–140.
3. Ritcey, G. M.; Ashbrook, A. W.; Lucas, B. H. Development of a solvent
extraction process for the separation of cobalt from nickel, presented at the
Annual AIME Meeting, San Francisco, 1972. CIM Bull., January 1975.
4. Murray, K. J.; Bouboulis, G. J. S/X carrier solvents, Proceedings of AICHE
Symposium on Solvent Ion Exchange, Tucson, Arizona, May, 1973.
5. Lo, T. C.; Baird, M. H. I.; Hanson, C. Eds., Handbook of Solvent Extraction;
John Wiley and Sons, New York, 1982.
334 Ritcey
6. Warwick, G. C. I.; Scuffham, J. B.; Lott, J. B. Proc. ISEC ’71, The Hague,
Society of Chemical Industry, London, 1971, 1373p.
7. Warwick, G. C. I.; Scuffham, J. B. Proceedings of International Symposium
on Solvent Extraction on Metallurgical Processes, Antwerp, 1972, pp. 40–
47.
8. Agers, D. W.; Dement, E. R. Proceedings of International Symposium on
Solvent Extraction on Metallurgical Processes, Antwerp, 1972, pp. 31–39.
9. Ritcey, G. M. Proceedings of AIChE Symposium on Solvent Ion Exchange,
Tucson, Arizona, May 1973.
10. Nelson, R. R.; Brown, R. L. The Design of Metal Producing Processes, Kibby,
R. M., Ed., American Institute of Mining Metallurgy, New York, 1967, p. 324.
11. Miller, A. The Design of Metal Producing Processes, Kibby, R. M., Ed.,
American Institute of Mining and Metallurgy, New York, 1967, p. 337.
12. Kunda, W.; Warner, J. P.; Mackiw, V. N. Trans. CIM, 1962, 65, 21.
13. McKay, H. A. C.; Healy, T. V.; Jenkins, I. L.; Naylor, A. Eds., Solvent Ex-
traction Chemistry of Metals, MacMillan, London, 1966.
14. Gindin, L. M.; Bobikov, P. I.; Rozen, A. M. Russ. J. Inorg. Chem., 1960, 5,
906, 1146.
15. Ashbrook, A. W. J. Inorg. Nud. Chem., 1972, 34, 1721.
16. Ritcey, G. M.; Lucas, B. H. Proc. ISEC’71, The Hague, Society of Chemical
Industry, London, 1971, p. 463.
17. Blake, C. A.; Brown, K. B.; Coleman, C. F. USAEC Rep. ORNL-1903, 1955.
18. Slyusarev, D. F. Zh. Anal. Khim., 1972, 27, 753.
19. Pich, H. C.; Santos, C. S.; Elias, J. T.; Alves, M. F.; Conceicao, E. H.; Viera, H.
X. Proceedings Symposium on Recovery of Uranium, IAEA, Vienna, paper
IAEA-SM-135/17, 1971.
20. Ashbrook, A. W. Anal. Chim. Acta, 1972, 58, 115.
21. Ritcey, G. M. Proceedings Second Annual Meeting, Canadian Hydro-
metallurgists of Metal. Soc., CIMM, Canadian Institute Minerals and Met-
allurgy, October 1972, p. 11.
22. Ritcey, G. M.; Lucas, B. H. CIM Bull., February 1974.
23. Shulz, W. W.; Navratil, J. D.; Kertes, A. S. Science and Technology of Tributyl
Phosphate, CRC Press, Vols. 1–4, 1991.
24. Higgins, C. E.; Baldwin, W. H. Anal. Chem., 1960, 32, 233.
25. Rydberg, J.; Reinhardt, H. Chem. lnd., 1970, p. 488.
26. Ashbrook, A. W. Commercial chelating solvent extraction reagents. 1. o-Hy-
droxyoximes; Purification and isomer separation/extraction. Metallurgy Divi-
sion, Mines Branch, Dept. Energy, Mines and Resources, Canada, Report
EMA 73-10, 1973.
27. Takahashi, M.; Ogata, T.; Okino, H.; Abe, Y. Paper abstracts ISEC’83,
Denver, USA, pp. 359–360, 1983.
28. Ashbrook, A. W. Commercial chelating solvent extraction reagents, II. n-Al-
kenyl-8-hydroxyquinoline; Purification and properties, Metallurgy Division,
Mines Branch, Dept. Energy, Mines and Resources, Canada Report EMA 73–
34 1973.
Industrial Solvent Extraction Processes 335
29. Rickelton, W. A.; Mihaylov, I.; Love, B.; Louie, P. K.; Krause, E., U.S. Patent
5,759,512, June 2, 1998.
30. Rickelton, W. R.; Robertson, A. J.; Hillhouse, J. H., Solv. Extr. Ion Exch.,
1991, 9(1):73–84.
31. Marston, A. L.; West, D. L.; Wilhite, R. N. Solvent Extraction Chemistry of
Metals, McKay, H. A. C.; Healy, T. V.; Jenkins, I. L.; Naylor, A., Eds.,
MacMillan, London, 1966, p. 213.
32. Ritcey, G. M. Hydrometallurgy, 1980, 5, 97.
33. Ritcey, G. M. Proceedings of 14th International Mineral Processing Congress,
CIM, Toronto; Vol. 1, 1982.
34. Ritcey, G. M. International Solvent Extraction Conference, ISEC ’83, Denver,
AIChE, 1983, pp. 88–89.
35. Cao, S.; Sworschak, H.; Hall, A. Proc. ISEC ’74, Lyon. Society of Chemical
Industry, London, 1974, pp. 1453–1480.
36. Ritcey, G. M.; Slater, M. J.; Lucas, B. H. Proceedings of International Hy-
drometallurgy Symposium, AIME, Chicago, February 1973, AIME, New
York, 1973, pp. 419–474.
37. Orth, D. A.; McKibben, J. M.; Scotten, W. C. Proc. ISEC’74, Lyon, Society of
Chemical Industry, London, 1974, pp. 514–433.
38. Fletcher, A. W.; Hester, K. W. A new approach to copper-nickel ore proces-
sing. Paper presented at the Annual AIME Meeting, New York, February
1964.
39. Burkin, A. R. Proceedings of First Hydrometallurgy Meeting, CIM, Ottawa,
October 1971.
40. Ritcey, G. M.; Joe, E. G.; Ashbrook, A. W. Trans. AIME, 1967, 238, 330–334.
41. Huppert, K. L.; Issel, W.; Knoch, W. Proc. ISEC’74, Lyons, 1974, The Hague,
Society of Chemistry Industry, London, 1971, pp. 2063–2074.
42. Young, W. Crud in Gulf Minerals Rabbit Lake solvent extraction circuit, Paper
at AIME Annual Meeting, New Orleans, February 1979.
43. Kowalik, T.; Cantwell, T. Crud problems in solvent extraction and strip circuits
of Conquista uranium, Paper at AIME Annual Meeting, New Orleans, Feb-
ruary 1979.
44. Bakshani, N.; Maurer, E. E.; Alien, M. P.; Degenhart, A. L. Crud at the Sohio
uranium solvent extraction circuit. Paper at AIME Annual Meeting, New
Orleans, February 1979.
45. Moyer, B.; McDowell, W. J. Proceedings of AIME International Hydro-
metallurgy Symposium, Hydrometallurgy-Research, Development and Plant
Practice, Osseo-Asare, K.; Miller, J. D. Eds., Metallurgical Society of AIME,
1983, pp. 503–516.
46. Ritcey, G. M.; Wong, E. W. Proc.ISEC’88, Moscow, 1988, 2:116.
47. Lucas, B. H.; Ritcey, G. M. CIM Bull. June 1975; Canadian Patent No.101759,
September 1977; U.S. Patent 3,969,476, 1976.
48. Rossiter, G. Anamax Twin Buttes oxide plant operating experience-first year.
Presented at the Arizona Section, AIME, Hydrometallurgical Division, Spring,
1976.
336 Ritcey
8.1 INTRODUCTION
The theory of solvent extraction was considered in Chapter 1, and Chapter 7
covered the application of liquid–liquid extraction in industry. The princi-
ples underlying the design of industrial applications are addressed in this
chapter.
At the very simplest level, an aqueous solution contains a valuable
component to be recovered, and a number of other components from which
the desired component should be separated. It is assumed that, as a result of
laboratory studies such as those outlined in previous chapters:
1. A suitable extraction system has been identified that will extract the
desired component selectively from the less desired components.
2. Suitable physical and chemical conditions for carrying out the extrac-
tion have been found.
3. The rates of extraction of the desired and possibly also some of the
undesired components have been determined at least qualitatively.
To design a process to recover the valuable component, a number of
questions which must be answered:
What fraction of the desired component can be recovered?
How much of the extractant must be added to achieve the desired
recovery?
How much of the undesired components are extracted with the desired
component?
339
340 Lloyd
8.2 EXTRACTION
8.2.1 Single-Stage Extraction
As described in Chapter 4, if a solution containing a desired component X is
contacted with an immiscible solvent phase, then X distributes itself between
the feed solution and the solvent according to:
DX ¼ [Xe ]=[Xa ] (8.1)
where the subscript e represents the extract (solvent) phase and subscript a
the aqueous phase.
If the phase volume ratio Y = Ve/Va, then the fraction extracted, EX,
is given by:
EX ¼ Dx =(Dx þ 1) (8.2)
that is, the fraction extracted in a single stage is a function of both the dis-
tribution ratio and the phase ratio. This is an important finding. Much of the
work in previous chapters has concentrated on the equilibrium distribution
of species between the extract and raffinate phase. However, as soon as an
answer to the question ‘‘What fraction of the desired component can be
recovered?’’ is sought, the volumetric ratio between the two phases becomes
almost as important as the equilibrium distribution. Consider now another
simple case with a desired component A for which DA = 10 and a con-
taminant X for which DX = 0.1. Table 8.1 shows the effect of varying the
phase ratio on (1) the fractions of A and X which are extracted, and (2) the
ratio of the fractions extracted, which is a measure of the product purity.
This not only shows how the extent of extraction varies with phase
ratio at a constant distribution ratio, but also how varying the phase ratio
affects the relative purity of the product. Increasing the phase ratio by a
factor of 100 nearly doubles the recovery, but drops the relative purity by a
factor of over 25. Note that, in a single stage, it is not possible to achieve
both high recovery and a high degree of separation simultaneously. Also,
342 Lloyd
although the distribution ratios of the two species differ by a factor of 100,
the relative purity is always less than 100.
A second phenomenon of great industrial importance is the effect of
saturation of the solvent on the product purity. Implicit in the derivation of
Eq. (8.2) is the assumption that DX is constant. As discussed in Chapter 2,
DX is only a constant under ideal and constant conditions (usually at trace
concentrations in both phases). It changes markedly as the concentrations
vary in the two phases.
At higher concentrations, so much of an extractable component may
be extracted that an appreciable fraction of the extractant is bonded to the
extracted component, so that in turn the concentration of the free ligand in
the extract phase is significantly reduced.
Industrial practice naturally requires the maximum use of the rela-
tively expensive extractant, so that saturation of the extractant phase with
reduction of the free ligand concentration to a minimum is the general rule.
A model is thus needed to quantify the effect of a reduction in DX as the
concentration of the extracted species in the organic phase increases.
In many extraction systems, DX is proportional to [L]n, where: [L] is
the free ligand concentration, and the exponent n is determined by the
number of ligand molecules per molecule of extracted complex. In these
systems, therefore, we may write:
DX =[Xe ]=[Xa ]=D0 f[Le ]tot m [Xe ]gn (8.3)
in which:
D0 is the distribution ratio of species X at trace concentrations, readily
determined in the laboratory.
[Le]tot is the total ligand concentration in the extract phase.
m approximates the ratio of ligand molecules to extracted molecules close to
saturation.
n approximates the ratio of ligand molecules to extracted molecules close to
infinite dilution.
The term in curved brackets on the right-hand side of the equation is the free
ligand concentration.
It should be noted that m and n in Eq. (8.3) are sometimes equal, but
often differ in value, which underlines the fact that this equation has no
theoretical basis. It is merely a convenient way of representing much ex-
perimental data using three parameters that can readily be determined
experimentally.
The equation can be fitted to a wide range of isotherms with quite
small residual errors over the entire range of concentrations of interest. It is
a liquid–liquid analogue of the Langmuir adsorption isotherm applicable to
Principles of Industrial Solvent Extraction 343
Fig. 8.2 Isotherms for the extraction of two species at a phase ratio of 1 and equal
concentrations in the feed.
stage to stage, but there is no particular advantage in doing so. Four stages
are shown with a feed at a relative concentration of 1.
Table 8.2 gives the stage-by-stage performance of the four stages of
repeated extraction, the cumulative extraction, ETOT, and the average con-
centration of AE, [AE]AVG, in the organic phases mixed together after each
repeated extraction. In addition, Table 8.2 also shows the performance of a
single stage in which the volume of extractant is the same as the total used in
the four stages, i.e., a single extraction at a phase ratio of 4.
This illustrates that even with a moderate distribution ratio (10 in
this case), useful recoveries (>99% as shown by ETOT for Stage 4) can be
achieved in comparatively few stages by repeated extraction. However, as
the number of stages increases, the concentration of the extracted species in
the combined extract, [AE]AVG, drops significantly.
A single-stage extraction using the same total volume of solvent
achieves only 92% extraction, and the extract concentration is only 0.23, vs.
nearly 0.25 for the cross-flow extraction. The use of four cross-flow extrac-
tion stages is clearly preferable to a single extraction. Equally, of course, the
use of more than four extraction stages, each with a proportionately smaller
volume, would improve the performance. In the limit, one would seek a
differential contacting process similar to the Soxhlet extractor employed for
extraction from solid phases, but such a contactor has not found use in
solvent extraction.
The mathematical treatment of a cross-flow cascade is straightfor-
ward. The mass, WE, which is extracted in n stages is given by:
X
WE ¼ VEn [CE ]n (8.7)
The cumulative recovery is given by:
ETOT ¼ WE =([CF ] V) (8.8)
while the average concentration in the combined extracts is given by:
Principles of Industrial Solvent Extraction 347
([AE ]1 [AE ]3 )
AF ¼ ¼ 1= (8.12)
([AF ] [AR ]2 )
which is identical to Eq. (8.11) except it represents the mass balance over
two stages.
Thus to achieve an extraction equivalent to that shown in Fig. 8.3
requires an overall mass balance for which, at the first stage, [AE]1 = 0.377
and [AF] = 1.0; while at the last stage [AE]3 = 0 and [AR]3 = 0.008. Then by
Eq. (8.11) over the whole cascade:
(1:00 0:008)
¼ ¼ 2:63
(0:377 0)
or the slope of the mass balance line = 1/2.63 = 0.38.
Mass balance lines such as AF are important and are known as
operating lines in countercurrent cascades. Clearly, an operating line can-
not cross an equilibrium line. Wherever an operating line approaches an
equilibrium line, a pinch point is the result and the number of stages needed
to achieve the desired degree of extraction approaches infinity.
In Table 8.4, the calculations behind Fig. 8.5 are repeated for a value
of Y = 2.63. Comparison with the results of Table 8.2 indicates clearly that
the countercurrent cascade offers a similar overall extraction in the same
number of stages, together with an extract of significantly higher con-
centration.
The data of Table 8.4 are also shown in Fig. 8.6, which is shown as an
equilibrium isotherm, an operating line, and a series of steps between the
operating line and the isotherm. These steps are entirely equivalent to the
lines establishing the mass balances for each stage in Figs. 8.3 and 8.5.
For instance, the horizontal line AB represents [AF]1[AR]1, while the ver-
tical line BC represents [AE]1[AE]2.
The graphical construction of an extraction isotherm, an operating
line, and the stepwise evaluation of the number of stages in this manner is
known as a McCabe-Thiele diagram. Historically, it found great application
in a variety of mass transfer operations, from gas adsorption through dis-
tillation to solvent extraction. However, the advent of modern computa-
tional techniques has made it largely redundant, as it is often easier and
certainly more accurate to calculate the cascade directly.
In part also this is because better equilibrium data are now available
through the use of equipment such as AKUFVE, which has brought the
realization that it is not sufficient to view the isotherm as fixed. Quite small
changes in aqueous composition or in phase ratio can change the isotherms
and cause dramatic effects on the performance of a countercurrent cascade.
Even relatively crude models for the isotherms, such as Eq. 8.3, can demon-
strate these effects in cascades.
Consider now the extraction of two species simultaneously in a coun-
tercurrent cascade. To obtain the equilibria requires solving two equations
similar to Eq. (8.4) simultaneously. Determining the concentrations within
the cascade means taking into account a mass balance such as Eq. (8.11) at
each stage. Table 8.5 presents such calculations for the same two equilibria as
used before, with A having D = 10 and B having D = 1.0. It should be noted
that the calculation is not simple, and requires estimation of the equilibrium
at each stage before the next stage is computed.
The behavior of the system is very counterintuitive. The more extract-
able component is relatively well behaved and a recovery of 99.4% is
achieved in four stages. In contrast, the less extractable component behaves
unusually. Its concentration increases in the aqueous phase between the first
and the third stages. In the extract phase, it has a peak concentration in the
extract from the fourth stage. Thus, the less extractable component circulates
between the first and last stages, being ‘‘squeezed out’’ of the organic phase
by the more extractable component, then being extracted back from the
aqueous phase once the concentration of the more extractable component in
the aqueous phase has fallen.
This is caused by the coupling of the two equilibria. The presence of
high concentrations of the less extractable component so disturbs the equi-
libria that the more extractable component has a distribution ratio of only
about 2.5 in the fourth stage of the cascade. If the less extractable component
352 Lloyd
were not present, the distribution ratio would approach the infinite dilution
value of 10.
Because of the high concentrations of the less extractable component
within the cascade, the final separation between the two components is not
particularly good. The final extract is only 85% pure. This could be improved
by reducing the phase ratio, which would have the effect of loading the more
extractable component sooner in the cascade, and thus squeezing out the
less extractable component better. In practice, however, one cannot reduce
the phase ratio too far without running the risk of the cascade becoming
unstable, or requiring, for instance, temperature control. It is preferable to
scrub the extract, as described in the next section.
It should be noted that this was a fairly severe test. Having a feed
containing equimolar quantities of components whose extractability differs
by a factor of only 10 is rare. Usually the differences in extractability are
greater, or the less extractable component is present in low concentrations
relative to the more extractable. In both these cases, good separations can be
achieved in a countercurrent cascade.
The cascade can be drawn graphically in the McCabe-Thiele form, but
the equilibria are so distorted that the exercise is not very valuable. This is a
general finding in industrial practice. Graphical methods are adequate to
give an indication of the number of stages likely to be needed for a partic-
ular duty. They fail when separations between similar species are sought.
Highly selective extractants are desirable, but often show poorer phys-
ical or chemical properties than less selective extractants. As the example
above illustrates, a difference of a factor of only 10 in the distribution ratios
Principles of Industrial Solvent Extraction 353
8.3 STRIPPING
The general principles established for extraction apply to stripping, although
of course distribution ratios are sought that are significantly less than unity in
order to accomplish the strip as efficiently as possible.
Stripping can equally be done in single-stage, cross-flow, or counter-
current systems. To illustrate how the overall concepts remain valid, the per-
formance of a countercurrent cascade accepting as feed the scrubbed extract
Principles of Industrial Solvent Extraction 355
of Table 8.6 is calculated assuming that the equilibria involved are described
by:
DA ¼ 0:1(1 2[Ae þ Be ]) (8.13a)
another. As long as they provide adequate interfacial area for the extrac-
tion to take place, without creating such small droplets that they will not
settle efficiently, and provided there is sufficient residence time for the desired
degree of extraction to take place, then one mixer will work as well as
another.
There is, however, one physicochemical criterion that is important in
industrial mixing, and that is ensuring that the correct phase is dispersed in
the other. There are several reasons for this:
It is sometimes found that mass transfer is more rapid if one phase is the
dispersed phase rather than the other.
Alternatively, the dispersed phase is chosen because, by definition, it will
not contain droplets of the continuous phase. In this way the dispersed
phase, after settling, will not entrain the continuous phase and entrain-
ment losses from the settler will be reduced.
Whatever the reason for choosing the dispersed phase, it is important to
ensure that the mixer will keep that phase dispersed during operation, as
changes in the dispersed phase, i.e., phase inversion, can cause considerable
operating problems.
Usually the continuous phase is the phase present in greater volume. It
is possible to run for long periods with the greater volume phase dispersed,
but phase inversion is always a risk in such circumstances. To overcome this
risk, where it is desired that the lesser volume phase is continuous, then a
portion of that phase may be recycled from the settler back to the same
mixer to ensure that within each stage it is the greater volume, even if it is
the lesser volume phase overall.
In large-scale operation, the volumetric flow of the phase to be dis-
persed is so large that it becomes necessary to disperse that phase into the
mixed phases. Otherwise ‘‘blobs’’ of the dispersed phase will act locally as
the continuous phase, and the intended continuous phase will be dispersed
in the blobs before the shear forces in the mixer break them up. This can
lead to excessive entrainment losses.
In some mixer-settler designs, the impeller is arranged both to mix the
two phases and to provide the necessary energy to transfer the phases from
one stage to the next, in which case it is known as a pump-mix mixer. The
head required to move a phase from one stage to the next is small, so the
impeller need not be efficient as a pump. Nevertheless, the design of impellers
for the dual purpose of both mixing and pumping is more of an art than a
science. Moreover, in full-scale operation it has been found difficult to start
up cascades of pump-mix mixers and achieve equilibrium rapidly. Accord-
ingly, this design has primarily found use in small-scale applications in the
nuclear and pharmaceutical industries.
Principles of Industrial Solvent Extraction 359
oscillations in the level of the interface that light phase often passed over the
heavy phase weir.
The introduction of the mixed phases into the settler has been found to
be important if clean separation is to be ensured. ‘‘Picket fences,’’ vertical
plates set at an angle to the flow from the mixer into the settler, have been
used to calm the flow and ensure the spreading of the mixed stream across the
width of the settler. Various packings have been employed to aid settling, but
under industrial conditions they are liable to clog with adventitious material
or ‘‘third phases.’’
Baffles placed across the settler, of progressively lower height from
entrance to overflow, have been employed to hold back the mixed phases to
permit them to separate. The mixed phases will spread rapidly right across a
settler unless there is a baffle to hold them back. There have been many
reports of ‘‘wedges’’ of mixed phase in small-scale settlers. These are never
seen in industrial practice because considerations of pressure at a given point,
as were used above to determine the height of the interface, show that a
wedge is inherently unstable.
In one mixer-settler design, the mixed phases flow down a shallow
trough placed over the settler, which gives them an opportunity to coalesce
and separate before entering the settler. In this way, the capacity of the settler
is markedly increased, with a concomitant reduction in the inventory of
solvent required for a given duty.
The manner in which individual mixer-settler stages can be linked
together to form countercurrent cascades is illustrated in Fig. 8.7. If each
stage is on the same level, then some form of pump must be provided to
move each phase from one stage to the next. As indicated earlier, it is
sometimes convenient to use the mixer for this duty. Another arrange-
ment has the individual stages set at different elevations, so that one phase
(usually the phase with the greatest flow rate) can gravitate from one stage to
the next.
where [A]f and [A]r are the feed and raffinate concentrations of A respec-
tively. This is illustrated in Fig. 8.9. The driving force is [A] [A]e and the
inverse of the driving force is to be integrated between the feed and the
raffinate concentrations (not shown in Fig. 8.9).
This integral clearly depends on the slope of the operating line and, as
in the case of stagewise operations, if the operating line approaches the
equilibrium curve too closely, then the driving force approaches zero and
the inverse becomes very large. That is, when the operating line is close to
Fig. 8.9 Illustration of the determination of the driving force from equilibrium and
operating lines.
362 Lloyd
the equilibrium curve and there is a pinch point, the number of transfer units
becomes large.
Equation 8.15 refers to a single phase, and [A]e is related to [A] at each
point via the phase ratio (or operating line, which is the same thing). The
NTU could be calculated for the extract phase instead of the aqueous phase,
that is, either from the difference between [AR] and [AR]e or from that
between [AE]e and [AE].
Where there is an analytical expression for the equilibrium such as
Eq. (8.3), then Eq. (8.15) may be integrated directly. Otherwise it is neces-
sary to perform the integration numerically or graphically.
Once the number of transfer units has been found, the height of the
tower is determined from the product of the number and the height of each
transfer unit (HTU). The HTU is determined by physical parameters such as
the droplet size, the flow patterns in the tower, and the effect of any packing.
These all affect the rate of mass transfer, which is addressed in Chapter 9.
Very often the rate of mass transfer cannot be estimated from first princi-
ples, and it is necessary to estimate the height by determining the number of
transfer units achieved and then dividing the actual height of the column
employed by the number of transfer units, i.e.:
HTU ¼ H=NTU (8.16)
where H is the height of the column employed.
HTU is subjected to the effects of both radial and axial mixing,
and these are not readily quantified, so scale-up of columns of this kind is
often not based on fundamentals, but rather on correlations determined
from detailed studies of several systems in the particular design of column
chosen.
Physically, towers designed for countercurrent contact can be open,
but more usually contain some form of packing or plates. The material of
the packing is chosen so that one phase wets it preferentially, thus increas-
ing the surface area for mass transfer. Similarly, the plates are designed to
breakup droplets and increase the surface area. In addition, the contents of
the tower may be agitated either by an internal agitator or by pulsing the
fluids. The energy imparted by agitation or pulsation breaks up the droplets
of the dispersed phase. Again, further details are given in Chapter 9.
When the equilibrium curve is relatively linear, the driving force does
not vary greatly down the length of the column, and the number of transfer
units approaches the number of McCabe-Thiele theoretical stages. In this
case, it is reasonable to speak of the number of stages in the column, and to
calculate a height equivalent to a theoretical stage (HETS). However, if the
equilibrium curve and the operating line are far from parallel, the number of
Principles of Industrial Solvent Extraction 363
except that the product streams [AR]e and [AE]e are no longer at equilibrium,
but are reduced by the inefficiency to [AR]i and [AE]i. This is illustrated in
Fig. 8.10.
364 Lloyd
for some systems mutual miscibility must be taken into account, particularly
when the primary solute is organic.
The methods for doing so are described in Chapter 9. The basic prin-
ciples remain unchanged-the primary difference is the choice of a consistent
basis for calculation, such as a solvent-free basis. Graphic techniques based
on triangular coordinates provide approximate answers, but modern com-
putational techniques are to be preferred.
Some consideration should be given at this point to the need to pre-
vent loss of the organic phase in the aqueous raffinate. This loss can arise by
either solubility in the aqueous phase or by entrainment of droplets not fully
settled. The solvent lost in this way can offer a finite environmental hazard
and be an economic cost on the process.
Clearly the primary duty is good engineering practice, which is covered
especially in Chapter 9. Often, however, additional security is provided in
the following form:
Additional settler capacity for final raffinate
Extraction of residual organic phase using a third diluent, from which it
is later separated, typically by distillation
Coalescence on a solid wetted preferentially by the organic phase
Flotation with air in the presence of surfactants
8.7 SUMMARY
In this chapter, we have seen the way in which laboratory studies of solvent
extraction are adapted to industrial use. Starting from a batch extraction, it
was shown how both recovery and product purity could be markedly influ-
enced by the volumetric phase ratio, and how it was impossible to achieve
both high recovery and high purity in a single stage. Cross-flow or repeated
extraction was then evaluated, and it was shown how this could improve
366 Lloyd
Fig. 8.11 McCabe-Thiele diagram showing the effect of reduced efficiency of extrac-
tion.
both recovery and purity, yet often resulted in mixed extract solutions that
were too dilute to be processed without further upgrading.
The concept of countercurrent extraction was then introduced, and it
was shown how the minimum phase ratio for a given degree of extraction
was determined. Countercurrent extraction could yield both high recoveries
and concentrated extracts, but studies on two-component extractions soon
showed that product purity suffered.
This led to the discussion of scrubbing as an essential adjunct to
countercurrent extraction where purity was important, and it was shown that
washing the extract with a small amount of aqueous phase could improve
purity markedly. Stripping was shown to follow the same underlying prin-
ciples as extraction for achieving efficient removal of extracted species, and
the need to choose phase ratios carefully to maximize the concentration of
the desired species in the strip solution was stressed.
There followed a brief discussion of equipment for carrying out sol-
vent extraction in industrial practice, both by stagewise and differential
contact. Some of the first principles for the design of differential contactors
were outlined and the part played by the efficiency of extraction in con-
tinuous equipment was discussed. Finally there was an outline of methods
for the control of solvent loss which forms probably the most important
environmental aspect of the application of solvent extraction.
9
Engineering Design and Calculation of
Extractors for Liquid–Liquid Systems
ECKHART F. BLASS* Technische Universität Munich, Munich, Germany
9.1 INTRODUCTION
The previous chapters have demonstrated that liquid–liquid extraction is a
mass transfer unit operation involving two liquid phases, the raffinate and the
extract phase, which have very small mutual solubility. Let us assume that the
raffinate phase is wastewater from a coke plant polluted with phenol. To
separate the phenol from the water, there must be close contact with the
extract phase, toluene in this case. Water and toluene are not mutually soluble,
but toluene is a better solvent for phenol and can extract it from water. Thus,
toluene and phenol together are the extract phase. If the solvent reacts with the
extracted substance during the extraction, the whole process is called reactive
extraction. The reaction is usually used to alter the properties of inorganic
cations and anions so they can be extracted from an aqueous solution into the
nonpolar organic phase. The mechanisms for these reactions involve ion pair
formation, solvation of an ionic compound, or formation of covalent metal-
extractant complexes (see Chapters 3 and 4). Often formation of these new
species is a slow process and, in many cases, it is not possible to use columns
for this type of extraction; mixer-settlers are used instead (Chapter 8).
This chapter explains in detail how this mass separation problem can be
solved. We shall adhere to engineering symbols as listed at the end of this
chapter (see section 9.10) to facilitate comparison with other chemical
engineering texts. Reference to these symbols will be made in the text without
any extensive explanations.
*Retired.
367
368 Blass
Fig. 9.1 Spray column for liquid–liquid extraction. The water phase flows from top to
bottom, the toluene as the lighter phase from bottom to top. If the column is always filled
with water at the marked height, the toluene breaks into drops at the feeding point. When
the drops have reached the water interface, they coalesce and form a continuous toluene
phase. The phenol transfers from the water phase to the toluene phase.
Fig. 9.2 Mechanism of drop formation on a single hole is dependent on the throughput.
Individual drops periodically leave the nozzle when the volumetric flow rate V is low.
When the flow rate is higher, the liquid forms a continuous jet that breaks into droplets.
(From Ref. 2.)
where nPi represents the number of drops of the class i of the diameter dP.
Figure 9.3a lists Sauter diameters calculated from drop size measurements at
single nozzles with dN = 1.5 mm for the system toluene (dispersed phase d )–
water (continuous phase c) as a function of the toluene velocity nN in the
nozzle [1]. The results are valid for sieve trays (and perforated plates), if
the distance between the holes is more than twice the Sauter diameter. Sieve
trays are simple and frequently used dispersion devices. Experiments have
shown that the flow rates through all the holes of a sieve tray are identical
only if the flow rate is exactly high enough to form a jet. Since the sieve tray
serves to distribute the toluene equally over the cross section of the column,
only the curve for nN>22 cm s1 is of interest (Fig. 9.3a). According to this
curve, the liquid flow rate, and thus the velocity through the hole nN should
372 Blass
Fig. 9.3 Sauter mean diameter d32 calculated from drop size measurements at single
nozzles of dN = 1.5 mm for the liquid systems (a) toluene (dispersed phase d ) water
(continuous phase c); and (b) butanol (d ) water (c), is dependent on the mean velocity vN
of the dispersed phase in the nozzle. (From Ref. 5.)
be chosen such that the Sauter diameter reaches its lowest value. This gives
the greatest possible drop surface as per unit volume and, thereby, the largest
mass transfer area:
A 6"d
as ¼ (9.2)
V s d32
when ed represents the holdup of the drops in the total volume.
From experiments, equations have been derived that enable calculation
of the minimum velocity in the nozzle, the nozzle velocity, and the Sauter
diameter at the drop size minimum. They provide the basis for the correct
design of a sieve tray [3,4]. Figure 9.4a shows the geometric design of sieve
trays and their arrangement in an extraction column. Let us again consider
toluene–phenol–water as the liquid system. The water continuous phase flows
across the tray and down to the lower tray through a downcomer. The
toluene must coalesce into a continuous layer below each tray and reaches
Engineering Design and Calculation of Extractors 373
Fig. 9.4 Sieve trays with downcomers for liquid systems with (a) high interfacial tension
and dual-flow trays for liquid systems; and (b) with low interfacial tension. In the case of
downcomers, only the phase to be dispersed flows through the holes. The droplets are
formed by jet disintegration. In dual-flow trays, both liquids flow through the same holes
alternately. The larger drops split because of collision with the tray.
374 Blass
gdN2 1
Limiting criteria
3
a height such that the static pressure caused by the buoyancy presses the
liquid through the holes at the required velocity. Figure 9.3b shows the
results of another liquid system (i.e., butanol–water). Compared with the sys-
tem toluene–water, its interface tension s is much lower [5]. Therefore, the
drops are much smaller, and very low velocities at the hole cause the for-
mation of a jet. Clearly, surface tension is a very important material property
for extraction systems.
The extremely low liquid velocities at the hole require only very small
stationary layers, which cannot be implemented in technical columns without
difficulties. Therefore, sieve trays without downcomers are used (Fig. 9.4b).
Here, both liquids must flow countercurrently through the same holes.
Neither stationary layers nor liquid jets of the dispersed phase are formed.
According to experimental observations, the bigger drops split up because
they collide with the sieve trays. The drops flow stepwise through the holes in
bigger collectives and alternate with the continuous phase flowing through
the holes. Table 9.1 shows reliable, experimentally proved equations for
medium drop sizes in such dual-flow trays and also a limiting criterion
for sieve trays with or without downcomers according to Hirschmann
and Blass [5]. If the hole diameter dN obtained from the limiting criterion in
Table 9.1 is smaller than 2 mm, dual-flow plates should be selected.
Sieve trays of this type define the possible size of drops, once the sieve
tray has been designed for the intended flow rate of the dispersed phase. The
same can be applied if tower packings are placed in a column to break up the
dispersed phase by colliding with the packing. The drop size can be changed
in an extractor during operation by supplying mechanical energy. The three
most important operations for an additional energy input are mixing, puls-
ing, and centrifuging. Extractors with these systems are more flexible than
spray and simple plate columns in that they can be adapted to changing
liquid systems and operating conditions, although they are more expensive in
initial costs and in maintenance.
Engineering Design and Calculation of Extractors 375
Fig. 9.5 Sieve tray columns with pulsing of the whole liquid content (left side) or of the
whole plate package (right side, Karr column). The trays are meant to equalize the phase
flows across the column cross section and to disperse the droplets as uniformly as possible.
376 Blass
if the plate geometry is identical. It also shows the principal interface in the
upper part of the column so that the lighter phase is dispersed.
Figure 9.6 demonstrates what happens at the sieve trays when the
whole liquid is pushed up and down. Three different situations can be
observed with the aid of high-speed photographs. During the downward
motion of the liquid pulse, larger drops do not follow the motion of the
pulsing, but accumulate below the plate owing to their relatively high
buoyancy. Thus, once the upward motion starts, the relatively large drops of
this layer, together with the continuous liquid, are pushed through the holes.
Fig. 9.6 Behavior of droplets near the sieve tray in a liquid-pulsed column during
downstroke and upstroke of the liquid content.
Engineering Design and Calculation of Extractors 377
The smaller drops follow. Both liquids accelerate in the holes, because the
sum of the cross section of all the holes is less than half the column cross
section. However, this motion is retarded within a short distance, whereby a
zone of drop compaction results above the trays. These phenomena are
modeled based on a balance of maximum and minimum kinetic energy and
the cohesive energy of the droplets [1]. After that, the resulting equation for
the maximum stable drop diameter in the field of pulsing is:
4
dP;max ¼ (9.3)
d (12 22 )
where n1 is the highest nozzle velocity, n2 the velocity at a short distance
above the sieve tray with an approximate value of zero.
New modeling was performed [6] taking into consideration first the
influence of drop viscosity that produced a better agreement with the exper-
imental results than Eq. (9.3). The authors presume that bigger drops have to
be deformed in the sieve tray holes and therefore the circulation inside the
drops is accelerated. Taking into account that the kinetic energy of drops is
changing into surface energy and circulation energy, the authors derive
Eq. (9.4) by balancing the energies for the stable drop size ds at a sieve tray
hole.
dN2
12 þ 20:5 d jcN þ A j
ds2
ds ¼ (9.4)
d (cN þ A )2
where
ds = stable drop diameter
dN = hole diameter
ncN = velocity of continuous phase in the hole
nA = rising velocity of a drop, Eq. (9.15)
nd, rd = dynamic viscosity and mass density of dispersed phase
Figure 9.7a shows values of stable drop diameters that are calculated
with Eqs. (9.3) and (9.4) for special conditions for the liquid system toluene–
water as a function of the product of pulse amplitude, a, and frequency, f,
the so-called pulse intensity. Figure 9.7b is a parity plot and shows that
Eq. 9.4 describes the measured drop sizes for a wide range of velocities and
parameters of the sieve tray geometry quite well.
The mean Sauter diameter, d32, is clearly smaller than the maximum
stable diameter. As experimentally proved by many authors, the ratio of
these is between 0.4–0.6.
There are various further attempts to model the complex flow and
drop-disintegration phenomena for drop size calculation. However, it has
378 Blass
Fig. 9.7 (a) Diameters of stable drops calculated from Eqs. (9.3) and (9.4) for the liquid
system toluene-water as a function of the pulse amplitude a and frequency f. (b) Parity plot
of measured and calculated crop sizes. (From Ref. 6.)
not been possible so far to include mass transfer influences and the inter-
action of dispersion and coalescence. Only empirical equations given in Ref.
[1, Chapter 17] give the best fit to experimental data and have the broadest
range of validity.
0:10
d32 h af 1=4
rffiffiffiffiffiffiffiffiffi ¼ C1 e0:74 exp 3:00 1=4 1=4
0:05 meter g
g
af 1=4
þ exp 28:65 1=4 1=4 (9.5)
g
Engineering Design and Calculation of Extractors 379
The best values of the parameter C1 are 1.51, 1.36, and 2.01 for no mass
transfer, c ? d and d ? c direction of transfer respectively. The product af is
considered as the agitation variable in the equation, since the fit could not be
improved if a and f were treated separately. The average absolute value of
the relative deviation in the predicted values of d32 from the experimental
points is 16.3%. Even in packed columns, the separation can be substan-
tially improved by pulsing of the continuous phase resulting from greater
shear forces that reduce the drop size and increase the interfacial area [1,
Chapter 8].
If the Newton number, Ne, in relation to the Reynolds number, ReR, of the
mixer is known from experiments and the rotor speed, nR, the mixer diam-
eter, dR and the mass density rc of the continuous liquid are given.
380 Blass
Engineering Design and Calculation of Extractors 381
nR dR2
ReR ¼ (9.7)
c
3
Fig. 9.8 Sections of several stirred countercurrent columns. Horizontal stator baffles
divide the column into successive sections, each fitted with a mixer, all fixed to a central
shaft. The mixers disperse one phase into droplets and mix the two phases in the section as
efficiently as possible. The stator baffles must support the phase separation and the
countercurrent flow of the liquids from state to stage.
382 Blass
Eq. 9.11 and differ in the proportionality constant. For details, the reader is
referred to references 3 and 4. For drop size calculations, the empirical
equations given in Ref. [1, chapter 17] that take into account all the influ-
encing parameters are preferable.
Rotating Disc Contactor (RDC):
0:12 0:16 2 0:59
d32 C1 c d DR c g
¼ pffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffiffiffiffi
ffi
DR 0:07 þ FrR c DR c
0:25 0:46
h DC
ð9:12Þ
Dc DR
The optimized values of C1 are 0.63, 0.53, and 0.74 for no mass
transfer, c ? d and d ? c, respectively. The value of the holdup is ignored
due to lack of data. Equation (9.12) predicts the drop size with an average
absolute value of the relative deviation of 23%.
Kuehni columns:
d32
= C1 e0:37 n0:11
s [0:14 þ exp ( 18:73FrR )]
DR
0:20 2 0:24
c DR c g
pffiffiffiffiffiffiffiffiffiffiffiffiffiffi ð9:13Þ
c DR
The values of the constant C1 are 9.81 102 for no mass transfer
and c ? d transfer, and 0.31 for d ? c transfer. The stage number ns, which
varies from 2–17 in the present set of data, shows a rather weak effect on
drop size. Equation (9.13) predicts the drop diameter with an average
absolute value and relative deviation of 17.6%.
Fig. 9.9 Experimental centrifuge with cylindrical dual-flow perforated sheets arranged
concentrically round a vertical shaft. The heavy phase is conveyed to the outside by cen-
trifugal field force; the light phase is displaced to the inside by the heavy phase. With a
control valve in the discharge pipe of the light phase, the degree of filling of the rotor with the
two phases can be defined (i.e., the location of the principal interface). (From Ref. 9.)
displaced to the inside by the heavy phase. This causes a countercurrent flow
of the two phases. With a control value in the discharge pipe of the light
phase, the degree of filling the rotor with heavy and light phase can be defined
(i.e., the location of the principal interface). This causes a high exit pressure
on the light phase, which just compensates the higher centrifugal force of the
heavy phase. The principal interface in Fig. 9.9 can be found at a medium
radius of the rotor. The perforated sheets in Fig. 9.9 have no downcomers for
the continuous liquid, which is common for technical centrifugal extractors.
Then, both liquids must flow through the same holes, just as they do when
dual-flow plates are used in the gravity field. However, other processes take
place here. The heavy phase accumulates at the inside of the sheets as a very
thin layer that then flows through a partial cross section of the holes coun-
tercurrently to the light phase and finally breaks from the hole as a single
384 Blass
drop. Only when the layers are thicker, which can be achieved by down-
comers in the continuous phase, do jets that disintegrate into drops originate
from the holes. The drops, influenced by the centrifugal field, move outward
and form another thin layer in front of the following perforated sheet by
coalescence. The same processes take place at the periphery of the rotor,
where the light phase is dispersed.
Influenced by interfacial tension and centrifugal forces, spherical drops of
various diameters originate at the holes. If we again assume the Sauter
diameter, according to Eq. (9.1), as the mean diameter of the spectrum of
particles, the following equation for heavy and light phases results from
theoretical and experimental results [10]:
rffiffiffiffiffiffiffiffiffiffiffiffiffiffi
c
d32 ¼ 3:22 (9.14)
r!2 d
where o is the angular velocity of the rotor that influences the size of the
drops, and r is the radius of the perforated sheet. Hence, at each stage of the
perforated sheet there is a different mean diameter.
If the distance between two perforated sheets, which is the settling
distance of the drops, is higher than about 40 mm, the drops of the heavy
phase are deformed by the accelerated motion, so that they again disin-
tegrate. Technical distances of the perforated sheets, however, should be
lower than 40 mm so that as many stages as possible can be built into the
rotor to improve the mass transfer performance. Sections 9.4 and 9.5.2 deal
further with this topic. Drop swarms of the light phase, however, do not
change their mean diameter during their retarded motion to the rotor shaft,
no matter how far is the distance to the next sheet.
In addition to the rotating columns previously described, there are a
number of other designs for centrifugal extractors, many originally developed
for the separation of radioactive wastes in nuclear processes (see Chapter 12).
They are both of the mixer-settler type, as discussed in section 9.3.3, and of
the rotating column types.
In the simplest case, a liquid phase in a static mixer is dispersed into
another phase and then fed into a centrifugal separator for phase separation
(external mixer–centrifugal separator) [11,12]. The advantage of this design
is that mixer and settler are separately optimized, and the centrifugal unit
is relatively simple. At best, for each mixer-settler stage, a separation per-
formance corresponding to a theoretical stage can be achieved; for a more
efficient separation, several of this type of mixer-settler may be connected in
series. Furthermore, the mixing can be integrated with phase separation in
the centrifuge into an ‘‘internal centrifugal mixer-separator,’’ as in the
Robatel design. Figure 9.10a shows four such interconnected stages. These
centrifuges are used, for example, by the French nuclear reprocessing
Engineering Design and Calculation of Extractors 385
Fig. 9.10 (a) Robatel four-stage centrifugal extractor. The heavy and the light liquids
are introduced near the bottom and top, respectively. They flow countercurrently and leave
the extractor at opposite ends. The extractor consists of a rotating bowl divided by baffles
into horizontal compartments on top of one another. The stationary central shaft carries
mixing disks (A) that run through each compartment and serve to mix and pump the two
phases to the settling chamber of the stage. This chamber comprises a set of two overflow
chutes (B for the heavy phase; C for the light phase), that stabilize the separating area
independently from the outputs. Each compartment has connections to take the settled
phases to the previous and following stages. The standard design has three to eight stages.
(Courtesy of Robatel SLPI and Eries, Genas, France.) (b, see next page) Westfalia coun-
tercurrent centrifugal extraction decanter. The heavy liquid phase containing solids such as
microorganisms and the solvent are fed into the rotating bowl through separate inlets at
different parts of the bowl. Phase contact is effected in countercurrent flow in the feed area
and in the spirals of the scroll. The separation zone is adjusted for variable liquid densities
by means of the ring dam at the cylindrical bowl end. After mass transfer in the contact
zone, the enriched solvent is discharged at the cylindrical bowl end under pressure by a
centripetal pump. The raffinate phase discharges at the conical end of the bowl. All
particles that sediment in the bowl are conveyed by the scroll to the conical end of the
bowl and discharged together with the raffinate phase. (From Ref. 13.)
Fig. 9.10 (continued ).
Engineering Design and Calculation of Extractors 387
The basic terms nP and dP and the liquid properties are made non-
dimensional in the equations according to the similarity theory:
388 Blass
P dP
Particle Reynolds number: ReP ¼ (9.16)
d
dP3 gc
Archimedes number: Ar ¼ (9.17)
c2
2c 3
Characteristic number of the continuous liquid: KL ¼ (9.18)
g c4
the flow rate of the dispersed phase is maintained. Thus, the remaining
flow area for the flow of the continuous liquid narrows. This interaction
between flows of the two phases leads to a limit of operation when the
buoyancy is not sufficient to overcome the flow resistance. This is a complex
phenomenon on which the deformability of the drops is superimposed, so
that modeling the real processes mathematically would not be particularly
promising.
However, a simple model does work: the liquids should occupy a share of
the cross section according to their holdup and should stream toward each
other as layers. If their superficial velocity, that is, their volumetric flow
rate, related to the complete column cross section, S, is:
V_ c V_ d
_ c ¼ ; _ d ¼ (9.19)
S S
they have an effective velocity of _ c ð1 "d Þ or _ d ="d . Now the relative
velocity nr, between the two liquids is:
_ d _ c
r ¼ þ (9.20)
"d 1 "d
Either nr or ed can be calculated from Eq. (9.20), if there is one more
relationship between the two terms. In Fig. 9.12, nr related to the single-
drop velocity nP according to Eq. (9.15) is plotted as a function of the drop
holdup for droplet swarms with the Archimedes number as a dimensionless
term for the drop diameter for the measured values. It can be seen that
the relative velocity constantly decreases, as the holdup of the drops, ed,
increases and the size of the drops in the swarm decreases.
Many experiments have been published that attempt to predetermine
the dependency of Fig. 9.12. To illustrate the main point, we use the simple
equation:
r
¼ ð1 "d Þk (9.21)
P
The velocity of single drops nP can be calculated by Eq. (9.15). The
exponent k is defined by the empirical correlation:
k ¼ 4:6 0:13 ln Ar (9.22)
which describes well-published experiments in the range 2.5<k<4.6 [14].
The premises have now been provided for calculating the operation
limits (i.e., the highest possible flow rates of a countercurrent column). This
is reached as soon as no further increase of the flow rate can be performed by
the density difference. The consequence is that the light phase leaves the
column at the bottom instead of at the top and the opposite situation for the
heavy phase (see Fig. 9.1). This load limit is also called the flooding point.
The mathematical formulation of this condition makes use of the fact
that an increase of the flow rate cannot increase the drop holdup:
390 Blass
Fig. 9.12 Stationary relative velocity vr, of the drop swarms related to the single drop
velocity vp is dependent on the drop hold-up, ed, and Archimedes number Ar. (From Ref. 2.)
d"d d"d
¼ ¼1 (9.23)
d _ d d _ c
or
d _ d
¼ 0 when _ c ¼ constant (9.24a)
d"d
d _ c
¼ 0 when _ d ¼ constant (9.24b)
d"d
If Eq. (9.20) for the relative velocity nr is used to calculate the superficial
velocity, one obtains the equations for calculating the phase flow velocities
at the flooding point:
dr
_ c ¼ ð1 "d Þ2 r þ "d (9.25a)
d"d
Engineering Design and Calculation of Extractors 391
dr
_ d ¼ "2d r ð1 "d Þ (9.25b)
d"d
Fig. 9.13 Flooding point diagram of countercurrent extractors. The drawn lines for six
different Ar values are valid for circulating drops. The dotted lines for three different KL
values are valid for oscillating drops. (From Ref. 2.)
Fig. 9.14 Flooding point diagram of centrifugal extractors (see Fig. 9.9). Curve 1 marks
the lower limit, when the rotor is filled with heavy phase; curve 2 marks an upper limit,
when the rotor is filled with light phase; curve 3 marks the limit of the total throughput of
both phases.
394 Blass
Fig. 9.15 Principle of residence time measurement of the continuous heavy phase in a
countercurrent extractor. After feeding a short test signal, for example, a colored tracer, the
color intensity ca or co, respectively, is measured downstream at two cross sections a and
o as a function of time. The test signal curve widens downstream, owing to the various
mixing phenomena.
Consider now a mathematical model for the mixing process that allows
the calculation of the distribution function co out of the function ca measured
at the cross section a, where co(t) is called the response curve. This model
must provide one or more parameters that define the degree of mixing in
addition to operation and construction parameters. If the calculated and
measured response curves are almost identical for the defined parameter
values, not only is the model proved, but it will also have yielded a set of
connected operation, construction, and mixing parameters. From a sufficient
number of these measurements, one can work out a functional dependence
for all of these parameters to predetermine the mixing processes in the var-
ious extractors.
Engineering Design and Calculation of Extractors 395
@c @2c _ c @c
¼ Dac;c 2 (9.28)
@t @h 1 "d @h
The solution of this partial differential equation is known for special
boundary conditions and measuring conditions at the cross sections a and o
Fig. 9.16 Basic assumptions of the one-dimensional dispersion model. The dispersion
of the residence time of the phases is modeled by superimposing the plug profile of the
basic flow with a stochastic dispersion process in axial direction.
396 Blass
of the measuring zone [17]. The experimenter can then calculate the response
curve in the cross section o, co(t), and can change the dispersion coefficient
Dax,c as the fitting parameter so that the calculated response curve co(t)
agrees best with the measured one. Figure 9.17 for liquid pulsed sieve tray
columns shows that the dispersion model gives a fine reflection of the process
of mixing of the continuous phase. The measured and calculated response
curves for a special value of Dax,c agree very well. However, there are large
deviations for the dispersed phase, in spite of the best fit. This leads to the
conclusion that the dispersion model provides an inadequate description of
this process. Relevant literature offers correlations for dispersion coefficients
Fig. 9.17 Residence times of (a) continuous; and (b) drop phase: in a liquid pulsed sieve
tray extractor according to measurements (squares and crosses) and according to calcu-
lations by the dispersion model (drawn lines). Measured and calculated response curves
agree well for the continuous phase, but not for the dispersed phase. (From Ref. 14.)
Engineering Design and Calculation of Extractors 397
in various extractors [1, Chapter 17]. For the continuous phase, they are
usually between 1 and 5 cm2 s1 (in some cases values as high as 20 cm2 s1
are observed); for the dispersed phase they increase by a factor of 2–3.
The design engineer can use the dispersion coefficients determined in
this way for the calculation of the real course of concentrations, c, of any
component in the dispersed (d ) and continuous (c) phases along the coun-
tercurrent column. If the mass transfer between the two phases, the actual
task of an extractor, is included in the balance, the balance equations for an
element of height dh of the extractor for stationary conditions is:
dcc d 2 cc
_ c þ ð1 "d ÞDax;c 2 ¼ kod aðcd cd Þ (9.29)
dh dh
dcd d 2 cd
_ d þ "d Dax;d 2 ¼ kod aðcd cd Þ (9.30)
dh dh
The term on the right side of the equations quantitatively defines mass
transfer from the dispersed to the continuous liquid, which is explained
more fully in section 9.6. The coupled differential equation system can be
analytically or numerically integrated for the appropriate practical bound-
ary conditions.
Figure 9.18 shows schematically the influence of the flow dispersion on
the concentration profile and the possible exit concentrations. Notice that
dispersion changes the possible exit concentrations of the two phases
negatively.
Fig. 9.18 The influence of flow dispersion on the concentration profiles of the two
phases (mass transfer from the continuous phase to the dispersed phase).
the various peripheral flow velocities at different radii of the rotor and by the
influence of the drops. If sufficient ideally mixed cells are connected in series,
plug flow residence time behavior results. As mentioned at the beginning of
section 9.5, there is no negative influence on the mass transfer by mixing and
the motion of the droplet swarm approaches that of plug flow, since neither
the drop sizes nor their velocity of motion are very different.
The dispersion model introduced in section 9.5.1 cannot be used for
the demonstration of the mixing process in a multistage centrifugal ex-
tractor, because it can describe only the behavior of the extractor overall.
There is no means of distinguishing between the various contributions to
mixing from the single steps. Models, however, that consist of terms for plug
flow patterns and ideal stirring tanks describe the real mixing behavior
exactly [10].
n_ c ¼ kc cc (9.31a)
n_ d ¼ kd cd (9.31b)
where n_ is the mole flux of the transitional component and Dc the con-
centration differences between the bulk phase and interface [18]. At steady
state, the mole flux of the component from one phase to the interface must
just be able to flow into the other phase. For this transitional state, an overall
mass transfer coefficient ko can be defined. To define ko it must be assumed
that the transitional resistance of the two phases is concentrated in one phase,
Fig. 9.19 The residence times of heavy droplets in a centrifugal extractor (see Fig. 9.9)
with three concentric perforated sheets. Extremely short residence times are a main at-
tribute of centrifugal extractors.
400 Blass
so that the bulk concentration of the resistanceless phase reaches as far as the
interface, as in Fig. 9.20. An overall mass transfer coefficient can then be
defined for the dispersed phase as well as the continuous phase:
It is generally agreed that mass transfer coefficients are only correlated for
negligibly small convectional motion of the transitional component, which is
vertical to the interface. However, when the mass transfer is mutual and
equimolar, no such convections normal to the interface result; otherwise the
transfer coefficient and the driving force must be corrected with the aid of
theories of mass transfer [18]. The transitional rates and, accordingly, con-
vectional flow rates normal to the interface are only low for the extraction
process, so that the uncorrected Eq. (9.31) may be used.
The intensity of mass transfer shown by the mass transfer coefficient
depends on the flow processes inside the drop or in its surroundings and,
thereby, on the various life stages of the drops. During the drop formation,
new interfaces and high concentration gradients are produced near the
interface. The contact times between liquid elements of the drop and the
surroundings that are near the surface are then extremely short. According
to Fick’s second law for unsteady diffusion, it follows that for the phase
mass transfer coefficient [19]:
rffiffiffiffiffiffiffi
4D
k¼ (9.34)
t
nonpulsed sieve tray columns, and stirred columns, the disintegration zone
is relatively small compared with the whole distance that the drops cover in
the extractor. Thus, the influence of the mass transfer kinetics on the integral
mass transfer during the drop formation can be neglected during the lifetime
of the drops.
Next, we are interested in mass transfer during drop motion, which
according to Fig. 9.11 is strongly dependent on drop size. When drops are
very small, and therefore rigid, diffusion in the interior of the drops deter-
mines the mass transfer velocity. When the internal drops circulation occurs,
however, mass transfer is considerably enhanced by the constant transport
of liquid elements to the drop surface. The concentration gradient driving
force is constantly kept at a high level. For oscillating drops, a further rise in
the mass transfer velocity can be expected, since the interior of the drop is
turbulently mixed.
402 Blass
very short residence times in the various steps of flow through a stage indicate
the use of the mass transfer coefficients according to Eq. (9.34) given by
Highbie [19]. Otillinger and Blass [9] have proved that Eq. (9.34) can actually
be used for the two mass transfer coefficients kc and kd in drop formation and
motion, as well as in the coalescence phase, and for the mass transfer between
liquid layers in a stage [1, Chapter 14]. Figure 9.21 shows this for a special
situation, with four actual stages formed of concentric sieve trays. The
number of theoretical stages is used as a measure for the performance of mass
transfer plotted radially in the centrifugal extractor. The increase of the
theoretical number of stages during drop formation and motion within the
radial path of each stage can be clearly seen. The number of stages,
depending on the coalescence and the contact of the liquid layers, can be
observed at specific radii. The straight lines have been calculated using the
mass transfer coefficient from to Eq. (8.34) using the residence times of the
respective flow steps as contact times. The final points of the calculation
Fig. 9.21 The mass transfer performance of a centrifugal extractor (see Fig. 9.9) with
four concentric sheets (stages). This is dependent on the radial path with the field force as
parameter: ns, number of stages; hs, height of a stage; rm, radial drop path in a stage; ri,
inactive radial distance (thickness of perforated sheet and of coalesced layer in a stage).
(From Ref. 9.)
404 Blass
agree well with the measured results for the whole extractor (i.e., that for the
whole radius of the rotor).
Fig. 9.22 Marangoni convection between two drops. If the mass transfer is directed out
of the drops, the transfer component is relatively more enriched in the thin region of the
layer. This corresponds to a lower interfacial tension for most of the extraction liquid
systems (ds/dx<0). Thus, a flow of liquid near the interface out of the region of lower
tension toward the ranges of higher tension results is called Marangoni convection. The
coalesence of the two droplets is promoted.
deeper layers of the phase to the interface and thereby increase mass
transfer. This is desired. In the past, criteria have been developed that allow
the prediction of such Marangoni instabilities [26]; however, the number of
correct and incorrect predictions is equal. Consequently, experiments should
be carried out for a safe prediction of the interfacial instabilities of a par-
ticular liquid system under various conditions. The ‘‘kicking drop’’ method
proves to be very useful. Drops are produced on a capillary in the con-
tinuous phase and, if Marangoni convection appears in the chosen direction
of the mass transfer, a strong vibration of the drop can be observed.
However, strong Marangoni convection can cause interfacial turbulence and
the formation of very small drops. Emulsions like these cannot be separated
in normal settlers and should be avoided a priori.
Finally, it must be noted that tensides that are adsorbed at the inter-
face cause a stiffening of the interface. They hinder or even stop the inner
circulation and oscillation of drops, and reduce the mass transfer intensity.
Moreover, they form a barrier against the mass transfer, so that a further
resistance term should be considered in the overall mass transfer process [28]
in Eq. (9.33). Since the nature and concentration of tensides in industrial
processes cannot be predicted, such phenomena cannot be taken into con-
sideration during equipment calculations.
The wetting of solid surfaces by liquids is also due to interface ener-
gies. As a measure of the wettability, the wetting angle is used, which occurs
at the contact line of the three phases between the solid and the wetting
liquid. It can be measured with suitable optical systems by placing a drop on
a plain, horizontal surface of the material used in the industrial apparatus.
For total wetting, the wetting angle is 08, rising to 1808 if a wetting is not
406 Blass
Fig. 9.23 Settlers for separation of liquid–liquid dispersions. Vh volume flow of heavy
phase; Vl, volume flow of light phase.
Engineering Design and Calculation of Extractors 407
Fig. 9.24 Basic phenomena of drop coalescence at a horizontal interface. The drop has
to reach the interface. A thin layer of the continuous phase remains between the drop
and the interface. The thin layer has to drain until it breaks up. Then the drop can flow
into its homophase. Mostly, the drainage process is the time-determining step of this
process.
_
Vd
H¼C (9.38)
S
0%
wt
8% 0%
wt
- 0%
wt
8%
Deq for (Deq = yt/xt)
6.5%
—————— —————— ——————
0.61–0.83 1.3–1.12 0.9–0.98
9.10 NOMENCLATURE
Nomenclature (continued )
9.10.3 Subscripts
c Continuous phase
d Dispersed phase
D Flow dispersion
i Phase i
i Class of drop sizes
j Jet formation
I Interface
K Column
L Liquid phase
max Maximum
min Minimum
N Nozzle hole
o Overall
P Particle
R Stirrer
r Relative
s Swarm of droplets
t Transfer component
z Cell
a Inlet
o Outlet
* Equilibrium
REFERENCES
1. Godfrey, J. C.; Slater, M. J. Liquid-Liquid Extraction Equipment; John Wiley
and Sons: Chichester, 1994; 630.
2. Philhofer, T; Mewes, D. Siebboden-Extraktionskolonnen; Verlag Chemie:
Weinheim, 1979; 10.
3. Blass, E.; Goldmann, G.; Hirschmann, K.; Mihailowitsch, P.; Pietzseh, P. Ger.
Chem. Eng. 1986, 9, 222.
4. Lo., T. C.; Baird, M. H. I.; Hanson, C. Handbook of Solvent Extraction; John
Wiley and Sons: New York, 1983.
5. Hirschmann, K.; Blass, E. Ger. Chem. Eng., 1984, 7, 280.
6. Wagner, G.; Blass, E. Chem. Eng. Technol., 1999, 21, 475.
7. Kolmogoroff, A. N. Sammelband zur Theorie der statistischen Turbulenz;
Akademieverlag: Berlin, 1958.
8. Barson, N.; Beyer, G. H. Chem. Eng. Progr., 1953, 49, 243.
9. Otillinger, E.; Blass, E. Chem. Eng. Technol., 1988, 11, 312.
10. Otillinger, F. Fluiddynamik und Stoffiibergang in Zentrifugalextraktoren
[doctoral thesis]; TU Munchen, 1988.
414 Blass
10.1 INTRODUCTION
The separation of organic mixtures into groups of components of similar
chemical type was one of the earliest applications of solvent extraction. In
this chapter the term solvent is used to define the extractant phase that may
contain either an extractant in a diluent or an organic compound that can
itself act as an extractant. Using this technique, a solvent that preferentially
dissolves aromatic compounds can be used to remove aromatics from ker-
osene to produce a better quality fuel. In the same way, solvent extraction
can be used to produce high-purity aromatic extracts from catalytic refor-
mates, aromatics that are essentially raw materials in the production of
products such as polystyrene, nylon, and Terylene. These features have
made solvent extraction a standard technique in the oil-refining and pet-
rochemical industries. The extraction of organic compounds, however, is
not confined to these industries. Other examples in this chapter include the
production of pharmaceuticals and environmental processes.
In these applications, solvent extraction constitutes an extraction stage
during which an organic phase is in contact with an aqueous phase or
another immiscible organic phase. The extract is then recovered by dis-
tillation or washing with an aqueous or organic phase.
Using new solvents and having a better understanding of the chemistry
involved in more specific interactions, solvent extraction has become a very
interesting separation technique for high-value organic chemicals (e.g.,
amino acids). Furthermore, liquid extraction using two phases with very high
water concentration has found applications for the separations of proteins.
415
416 Wennersten
10.3.1 Theory
Two main approaches can be taken when developing thermodynamic
equations for the correlation and prediction of equilibrium data. Both are
semiempirical in that they are based on simplifications of rigorous ther-
modynamic expressions, but include parameters that have to be fitted to
experimental data.
The first method, which is the more flexible, is to use an activity
coefficient model, which is common at moderate or low pressures where the
liquid phase is incompressible. At high pressures or when any component is
close to or above the critical point (above which the liquid and gas phases
become indistinguishable), one can use an equation of state that takes into
account the effect of pressure. Two phases, denoted a and b, are in equi-
librium when the fugacity f (for an ideal gas the fungacity is equal to the
pressure) is the same for each component i in both phases:
(Reference states for both phases are chosen at the same temperature.)
From classic thermodynamics, the following relation can be derived
[2]:
Z P
fi 1 @V RT
ln ¼ ln i ¼ T; P; nj dP (10.2)
xi P RT 0 @ni P
Z 1
fi 1 @P RT P
ln ¼ ln i ¼ T; V; nj dV ln (10.3)
xi P RT V @ni V RT
These equations are general and exact and express the fugacity of
component i in a mixture as a function of the measurable quantities pressure
(P), temperature (T ), volume (V ), and concentration (x).
The equilibrium condition is thus:
i xi ¼ i xi for i ¼ 1; 2; . . . n (10.4)
To solve the integrals in Eqs. (10.2) and (10.3), it is necessary to have
volumetric data; for example, an equation of state (EOS) of type:
V ¼ f1 ðT; P; n1 ; n2 . . .Þ (10.5)
or
P ¼ f2 ðT; P; n1 ; n2 . . .Þ (10.6)
Most of the equations of state are pressure explicit, and Eq. (10.6) can be
used for equilibrium calculations. As the integration is made from V to ?
the EOS has to be valid in the density range from zero to the actual density.
where Tc and Pc are the critical temperature and pressure, TR is the reduced
temperature (TR=T/Tc), and o is the acentric factor for component i.
422 Wennersten
For mixtures, one has to apply a mixing rule; for example, for a:
X
N X
N
a¼ xi xj aij (10.13)
i j
and
pffiffiffiffiffiffiffiffi
aij ¼ ð1 ij Þ ai aj (10.14)
and
bij ¼ ð1 ij Þbðbi þ bj Þ=2c (10.16)
where Zij is an adjustable parameter determined along with dij by fitting the
equation of state to equilibrium data. The cubic equations of state are
further considered by Marr and Gamse [5].
Activity coefficient methods work fairly well at temperatures well below the
critical, at which the liquid phase is largely incompressible, and up to
moderate pressures.
Fig. 10.1 Different types of liquid–liquid systems. (a), (b) Solubility as function of
temperature for binary systems; (c), (d) ternary systems. (Dashed lines are examples of tie
lines, which connect the two phases in equilibrium located at the binodal.)
Extraction of Organic Compounds 425
which gives the same reference state for component i in both phases, and the
equilibrium condition then reduces to
i xi ¼ i xi (10.24)
where a and b denote the two phases.
The activity coefficients can be calculated using any of the existing
models if the binary parameters for all combinations of binary pairs are
known. These parameters are obtained by fitting to experimental data. For
ternary systems, one can either simultaneously fit all six parameters or
first determine the parameters using binary data for those binary systems
that have a phase separation and the rest of the parameters from ternary
data.
In principle, as the g models do not have an internal temperature depen-
dence, the extrapolations in temperature must be carried out with great
care. When extrapolating in temperature, the parameters in the g models
should be made temperature dependent by fitting to experimental data at
different temperatures. Parameter estimation is obtained by minimizing an
object function such as:
X
N X
n
Q¼ ½ðXik Xik Þ2 þ ðXik Xik Þ2 (10.25)
k¼1 i¼1
Fig. 10.2 Liquid–liquid equilibrium for the system ethyl acetate–acetic acid–water at
303 K.
As can be seen in Figs. 10.2 and 10.3, the agreement with experimental LLE
is not very good.
Generally, one can say that parameters obtained from VLE should
be used only for qualitative calculations. The result is considerably im-
proved if the parameters for the binary ethyl acetate–water are calculated
from solubility data.
When both VLE and LLE data are missing, it is not possible to
assume ideal solutions. The separation into two liquid phases indicates a
strong deviation from ideality. Here, the remaining possibility is to use
some of the existing group contribution methods, such as UNIFAC, for
which the parameters are based on LLE data.
are the treatment of lubricating oils and the separation of aromatic and
aliphatic hydrocarbons. Other examples are the production of high-purity
fiber-grade caprolactam, recovery of acrylic acid, and production of anhy-
drous acetic acid. The success of solvent extraction in the petroleum and
petrochemical industry is due to careful process integration and also the
development of large-scale column extractors.
The development of different processes and solvents for the separation
of aromatics from aliphatics has reached a rather stable state. A number of
different processes, some of them with capacities of several hundred thou-
sand tons of aromatics per year, are in operation. The more important ones
are listed in Table 10.1.
The Union Carbide Process
The process equipment consists mainly of two extraction columns with
pulsating trays and four distillation columns according to Fig. 10.4 [7]. The
feed, with a high content of aromatics, is pumped to the middle of the first
extraction column where the aromatics are extracted with the solvent S1
(tetraethylene glycol). In the lower part, the extracted aromatics are washed
with S2 (dodecane). The outgoing raffinate phase R1 (containing aliphatics,
S2, and a small amount of S1) is distilled in the fourth distillation column.
The distillate, D4 consists of pure aliphatics, whereas the bottom stream
Table 10.2 Some Operating Conditions for the Union Carbide Process
First extraction Second extraction
column column
Equipment type Pulsed plate Pulsed plate
column column
Number of ideal stages 10 7
Extraction temperature 373 K 373 K
Solvent Tetraethylene glycol Dodecane
Reflux Dodecane -
Sl/S2 9 1.43
Sl/F 7 -
Aromatics in feed - 87 vol%
Recovery of benzene and - >99.5%
toluene (D1, D2) - -
Recovery of C8-aromatics (D3) - >98.5%
Purity of benzene - 99.9%
Here the authors found that the penicillin anion (P) could be
extracted efficiently with a secondary amine (Amberlite LA-2) in the pH
range 5–7 where the product is most stable. This type of process is used
extensively in hydrometallurgy (Chapter 11) and can be used to extract both
anionic species using cations as shown earlier, or cationic species using
organic acid anions. In hydrometallurgy, the system normally uses a hydro-
carbon diluent, but for pharmaceutical applications more polar diluents are
generally required.
Extraction of Organic Compounds 431
active carbon in
Solvent solvent
Water Solvent
acid
Penicillin
G or V
containing
broth
to distillation
solvent + impurities
40
10ºC
22ºC
30 60ºC
Extract phase conc., g\l
80ºC
20
10
10 20 30 40 50
Raffinate phase conc., g\l
However, similar problems can also occur with the alkylammonium ex-
tractants, especially as the pH of the solution must be sufficiently high to
produce an adequate concentration of the amino acid anions, i.e., more than
two units above the pK of the amino acid. Thus, coextraction of hydroxyl
ions competes with the amino acid anions and in addition lowers the pH of
the aqueous phase and reduces the concentration of amino acid anions. To
avoid such problems, a large buffer capacity is required in the aqueous phase
with the buffer chosen to minimize coextraction of anions.
10.5.1.3 Ethanol
One of the problems associated with the production of ethanol by fermen-
tation is the inhibition of the yeast activity with ethanol concentrations
greater than about 12 wt%, thus defining an upper practical limit for
ethanol production. Removing ethanol during fermentation by solvent
extraction would allow a greater total product yield from a given batch of
feed. The application of in-line extraction of ethanol from the fermentation
broth has attracted interest. Problems do arise with direct extraction from
broths with the effect of entrained organic solvents on microbial growth and
the reduction in mass transfer of the ethanol by the presence of cell debris
and other biological material. These problems can be minimized by suitable
choice of the extracting solvent, e.g., long-chain aliphatic alcohols such as
butanol, hexanol, etc., and removal of cell debris by filtration [26]. The
process operates as a closed loop with the filtered broth circulating through
Extraction of Organic Compounds 439
Fig. 10.12 Phase diagram and phase compositions of the dextran–polyethylene glycol
system D 48-PEG 4000 at 293 K. All values are % w/w.
Extraction of Organic Compounds 441
sensitive to product inhibition. The reaction may stop when only a small
fraction of the substrate has been converted into product. In such cases,
there is a strong need for continuous extractive procedures during which the
product is continuously removed. The integration of bioconversion and
extraction is called extractive bioconversion.
The large differences in surface tension and dielectric constants
between water and common organic solvents cause enzymes to unfold when
they are exposed to the interface between the two solvents. This problem can
Extraction of Organic Compounds 443
Fig. 10.17 (a) Extraction and separation through reducing the pressure; (b) extraction
and separation through a change in temperature; and (c) extraction and separation through
adsorption.
448 Wennersten
10.6.3 Applications
Much of the interest in SFE has been focused on using carbon dioxide to
extract different natural products from solid materials. Examples of large
industrial processes in this area are decaffeinating coffee beans and hop
extraction.
In the area of extracting solutes from aqueous solutions, many systems
have been screened in feasibility tests that have used carbon dioxide as a
solvent. A partial list of the solutes includes ethanol, acetic acid, dioxane,
acetone, and ethylene glycol. The reason for these efforts has been potential
low energy costs compared with distillation and the environmental advan-
tages of using carbon dioxide.
Other systems that have been investigated are the separation of bio-
cides from edible oils and fractionation of different components in vegetable
oils. An example of the latter is given next. For an extensive literature survey
of work done in the area of SFE, see references [31–34].
Separation of Mono-from Triglycerides
In Fig. 10.18, the quasi-ternary system carbon dioxide–acetone–glycerides
at 13 MPa is shown [34]. As shown in Fig. 10.18, acetone increases the
solubility of glycerides in the carbon dioxide phase. The entrainer has
the function of increasing the solubility of the solute and enhancing the
selectivity.
In Fig. 10.19, a simplified process scheme is given to illustrate the
separation of a nonvolatile compound from a mixture, using an entrainer.
The process consists of two columns. In the first column, the nonvolatile
monoglycerides are separated from the mixture, and in the second column,
they are separated from the solvent.
The mixture to be separated is fed, together with the entrainer, to the
middle of the first column. Here, the solvent, carbon dioxide and acetone, is
supercritical to provide high solubility of the monoglycerides. The super-
critical phase leaves the top of column I and goes to the lower part of
column II. In column II, the binary solvent entrainer is subcritical and in
the bottom of this column, the monoglyceride leaves, together with the
entrainer. Part of it is returned as reflux to column I, whereas the rest goes
to distillation for the separation of acetone. With a bottom temperature of
450 Wennersten
353 K and a pressure of 13.5 MPa in column I and 383 K at the top of
column II, 95% of the monoglycerides could be separated selectively from
the mixture.
Supercritical fluid extractions with solid feed stocks are industrially
carried out batchwise because of lack of equipment for feeding solid
materials to a pressurized extractor.
With liquid feed solutions, however, it is possible to work in a
manner analogous to traditional solvent extraction. Pressurized columns
can be of the packed-bed type or agitated by magnetic stirrers. Because
of the efforts of pilot plant tests, much of the scale-up work has to
be carried out in laboratory extractors. From solubility measurements, it
is possible to determine parameters in thermodynamic models (e.g.,
equations of state), which can be used for the simulation of large-scale
applications.
Extraction of Organic Compounds 451
10.7 NOMENCLATURE
The symbols used follow standard textbooks for chemical engineering [see,
Ref. 35].
fi Fugacity of component i
xi Mole fraction of component i in liquid phase
P Pressure
V Volume
v Molar volume
T Temperature
fi Fugacity coefficient of component i
ni Number of moles of component i
a Adjustable parameter in equations of state that accounts for
interaction between species in a mixture
b Adjustable parameter in equations of state that accounts for differences
in size between species in a mixture
oi Accentric factor for component i
Hi Henry’s constant for component i
GE Gibbs excess energy
gi Activity coefficient for component i
452 Wennersten
BIBLIOGRAPHY
Kennedy, J. F.; Cabral, J. M. S. Recovery Processes for Biological Molecules; Wiley:
1993.
Hatti-Kauk, R. Ed., Aqueous Two-Phase Systems; Humana Press, 2000.
Schugerl, K. Solvent Extraction in Biotechnology; Springer-Verlag: 1994.
Thornton, J. D. Ed., Science and Practice of Liquid-Liquid Extraction; Oxford Sci-
ence Publications: Vol. 2, chapters 3–5, 1992.
REFERENCES
1. Dortmunder Datenbank DDB (DETHERM database) [online]. Available from
STN International, DECHEMA e.V. server, 2001.
2. Smith, J. M.; Van Ness, H. C. Introduction to Chemical Engineering Thermo-
dynamics; McGraw-Hill: New York, 1975.
3. Peng, D. Y.; Robinson, D. B. Ind. Eng. Chem. Fundam., 1976, 15, 59.
4. Soave, G. Chem. Eng. Sci., 1972, 27, 1197.
5. Marr, R.; Gamse, T. Chem. Eng. Process., 2000, 39, 19.
6. Abrams, D. S.; Prausnitz, J. M. A.I.Ch.E., 1975, 21(1), 116.
7. Somekh, G. S.; Proc. ISEC 71, Lyon; Society of Chemical Industry: London
1971, 1, 323.
8. Kane, L.; Romanow-Garcia, S.; Nakamura, D. Hydrocarb. Process., 1995, 74,
36.
9. Verrall, M.S. in Science and Practice in Solvent Extraction, J. D. Thornton, Ed.,
Oxford Science Publications, Vol. 2, Chapter 3, 1992.
10. Kirk-Othmer Encyclopedia of Chemical Technology, 4th Edition 1993; Vol. 10,
pp. 125ff.
11. Reuben, B. G.; Sjoberg, K. Chemtechnol., 1981, 11, 315.
12. Reschke, M.; Schurgel, K. Chem. Eng. J. 1994, 28, B1-29.
13. Bailes, P. J. Chem. Ind., 1977, 69.
14. Bender, M. L.; Komiyama, M. Cyclodextrin Chemistry; Springer-Verlag: Ber-
lin, 1978.
15. Uemasu, I. Value Adding Through Solvent Extraction (Proc ISEC’96) edited
Shallcross, D. C., Paimin, R.; and Prvcic, L. M. Ed., Univ. Melbourne Press,
1966; Vol. 2, 1635p.
16. Pratt, M. W. T. Handbook of Solvent Extraction. Lo, T. C., Baird, M. H. I., and
Hanson, C. John Wiley and Sons: New York, 1983, p. 605.
17. Wohler, E. Removal and Recovery of Phenol and Ammonia from Gas Liquor;
Lurgi Kohle Mineraloltechnik GmbH, 1978.
18. Wennersten, R. J. Chem. Tech. Biotechnol., 1983, 33, 85.
Extraction of Organic Compounds 453
11.1 INTRODUCTION
The use of solvent extraction as a unit operation in hydrometallurgy now
extends to a wide range of metals from a variety of feed materials including
low-grade ores, scrap and waste, and dilute aqueous solutions. The tech-
nology was pioneered in the 1940s for the extraction of uranium from its ores
and, later, for the treatment of wastes from spent reactor fuel, still an
important use of the technique today (see Chapter 12). The knowledge
gained led to processes for the recovery of other high-value metals and the
separation of elements such as the rare earths, zirconium–hafnium, and
niobium–tantalum that, before the introduction of this technology, could be
separated only by lengthy batch techniques with many recycle steps to obtain
the desired purity. Gradually, solvent extraction was seen to have applica-
tions for the recovery of other less valuable, but important, metals such as
cobalt and nickel. At the time, the process was confined to rather small
operations. It was only after the development of selective chelating acidic
reagents in the 1960s that liquid–liquid extraction was seen to be a com-
mercially viable addition to the unit operations of hydrometallurgy and was
able to compete with alternative processes like cementation.{ Liquid–liquid
*Retired.
{
Cementation is the process of recovery of metals from dilute aqueous solution by reductive
precipitation using another metal with a more negative electrode potential, e.g., Cu2++
Feo?Cuo + Fe2+. The product, in this case ‘‘cement’’ copper, is relatively impure because of
iron contamination. However, cementation can be used in conjunction with a solvent extraction
flow sheet to remove small amounts of a metallic impurity, for example, removal of copper from
a nickel solution by cementation with nickel powder. Here the dissolved nickel conveniently
augments the nickel already in solution.
455
456 Cox
MXðnpÞ
p þ ðn pÞR3 NHþ X ! R3 NHðnpÞ MXp þ ðn pÞX
(11.2)
458 Cox
Versatic acids
R1 = R2 = C5 = Versatic 10
R1 = R2 = C4–C5Versatic 911
Phosphorus
acids
R = 5,5,7,7-tetramethyl-1-octenyl- (pre-1976)
= 4-ethyl-1-methyloctyl- (post 1976) Kelex 100
R = unknown side chain Kelex 108
R = unknown saturated alkyl LIX26
b-Diketones R1COCH2COR2
R1 = R-C6H5 R2 = CH3(CH2)5 LIX54
Solvent Extraction in Hydrometallurgy 461
solution, as no commercial flow sheet has yet been devised to remove nickel
selectively from cobalt. Two different solvent extraction systems are used
commercially involving (1) anion exchangers and (2) acidic chelating ex-
tractants, and these will be considered in turn.
þ
4 þ 2R3 NH Cl
CoCl2 ! ðR3 NHÞ2 CoCl4 þ 2Cl (11.4)
Solvent Extraction in Hydrometallurgy 463
Fig. 11.4 Extraction of cobalt and nickel with naphthenic acid (0.5 mol dm3), dashed
lines, and a mixture of naphthenic acid (0.5 mol dm3) and isotridecanal oxime (1.0 mol
dm3), both systems in toluene. (From Ref. 10.)
treatment usually to produce ferronickel, but the lower content of nickel and
higher levels of cobalt in the limonitic ores makes them more suitable to
hydrometallurgical processing. As the metal values are widely distributed
through the limonitic ore, preconcentration is not feasible so that large
amounts of the ore have to be leached, giving large volumes of a dilute leach
liquor. Two reagents are used for leaching, ammonia in the Caron process,
which dissolves the metal as ammoniates e.g., [Ni(NH3)6]2+, and acid, which
gives hydrated cations. Details of the complete hydrometallurgical flow
sheets are outside the scope of this chapter and can be found elsewhere [13].
Here discussion will concentrate on the solvent extraction flow sheets to
illustrate how the technology can separate the desired values from a range of
other metals present in the leach liquor.
11.3.2.1a Ammonia Leachate
The ammonia leach leaves the iron as an insoluble residue that is removed
by filtration. The filtrate is contacted with a ketoxime reagent (e.g., LIX84I,
Cognis) to extract the nickel [14].
þ
NiðNH3 Þ2þ
x þ 2HðoximeÞ ! NiðoximeÞ2 þ 2NH4 þ ðx 2ÞNH3
(11.5)
However, this will also extract cobalt(II) under the same conditions and once
extracted the cobalt is oxidized to cobalt(III) and cannot be stripped.
Therefore, to avoid this problem the cobalt is oxidized prior to extraction. In
this way, nickel can be selectively extracted from a feed solution (9 g dm3 Ni
and 0.3 g dm3 Co) leaving about 0.01 g dm3 Ni in the raffinate. Nickel is
subsequently stripped from the loaded organic phase with strong ammonia.
One of the main disadvantages of this flow sheet is that the cobalt(III) causes
some oxidation of the reagent and a process of reoximation is required [14].
Also, ammonia transfers to the organic phase, which again represents an
operating loss. However the introduction of solvent extraction raises the
nickel content of the products to >99.5% nickel as compared to 85–90%
nickel with the earlier flow sheet that did not use solvent extraction [13].
An alternative route to cobalt/nickel separation following ammonia
leaching in the Caron process is to produce a mixed Co/Ni carbonate
product. This can then be dissolved in sulfate media and processed using one
of the acidic extractants described later [13].
11.3.2.1b Acid Leaching
Acid leaching is usually carried out using sulfuric acid under pressure to
dissolve the majority of the iron minerals and to release the cobalt and
nickel. If this is carried out at 150–2508C, then the iron(III) is precipitated as
haematite or jarosite, reducing the amount of iron in the leachate. The
468 Cox
*Electrowinning is an electrolytic process where cathodic reduction is used to recover metal from
an electrolyte using an inert anode. The process differs from electro-refining where the anodic
reaction is dissolution of the metal, that is then reversed at the cathode.
Solvent Extraction in Hydrometallurgy 469
from which the copper is removed with a more concentrated acid strip and
cobalt by stripping in the presence of a reducing agent. One other major
problem is the oxidative degradation of the oxime extractant by the
cobalt(III) requiring the addition of a reoximation step to maintain ex-
tractant capacity. The cobalt in the raffinate is recovered by sulfide pre-
cipitation. The Cawse mine produced 8500 t/a nickel and 1900 t/a cobalt in
2000 at a cost in 1999 of $0.11 kg1 after cobalt credits, the lowest cost
production in the world.
11.3.2.1e Murrin-Murrin Process
The Murrin-Murrin mine is also near Kalgoorlie in Western Australia.
This is the largest of the three projects, with a first-phase production of
45,000 t/a nickel and 3000 t/a cobalt. Future expansion has been suggested
to 115,000 t/a nickel and 9000 t/a cobalt. The pressure acid leachate, fol-
lowing removal of the insoluble residue, is treated with hydrogen sulfide
under pressure to precipitate nickel and cobalt and reject impurity ele-
ments. The sulfides are then dissolved in a pressure oxygen leach to pro-
duce a sulfate liquor that is treated with CYANEX 272. Initially zinc is
removed at low pH, followed by a pH adjustment and extraction of cobalt
(Fig. 11.7). Any coextracted nickel is scrubbed from the loaded organic
phase with a dilute cobalt sulfate solution and the cobalt stripped with
dilute sulfuric acid. Both the cobalt and nickel are recovered from solution
as powders by hydrogen reduction that also produces ammonium sulfate
as a by-product. The main problem in this flow sheet has been the crys-
tallization of nickel ammonium sulfate because of the high nickel con-
centration (100 g dm3) in the feed to extraction. This severely reduced the
production from the plant and the problem was solved by preloading the
extractant with nickel prior to extraction by contact with the nickel sulfate
raffinate.
11.3.3 Summary
Cobalt and nickel can be separated by several solvent extraction processes,
the choice depending on the total flow sheet requirements, on the ratio of the
metals in the feed solution, and on the required purity of the products.
The various systems can be summarized as follows:
1. Extraction of cobalt from chloride media by amines
2. Extraction of cobalt by organophosphorus acids, especially CYANEX
272, from acidic solutions
3. Extraction of nickel after prior oxidation to cobalt(III) using hydro-
xyoximes
4. Extraction of nickel by mixtures of simple oximes and carboxylic acids
472 Cox
11.4 COPPER
There has been a long history of the extraction of copper by a hydro-
metallurgical route, with the Rio Tinto process being operated in Spain
from the mid-18th century. In the 19th century, the use of acid leaching of
low-grade oxidized ores was practiced in Russia and the United States, with
the resulting copper solution being reduced to metal by scrap iron. This
process of leaching–cementation requires an adequate supply of acid and
iron, and it produces only a low-grade copper powder that requires further
treatment by smelting before marketing. The introduction of electrowinn-
ing of copper from acid solution provided a number of advantages over
cementation [2] in that the product, cathode copper, had a much higher
value than the cement copper and also, during electrowinning some of the
sulfuric acid was regenerated for leaching. However, the demands placed on
the feed liquor to electrowinning were severe enough to limit the intro-
duction of the leach-electrowin route. Since high concentrations of other
metals such as iron and magnesium have effects on the current efficiency
and nature of the copper deposit, concentrations of these and several other
elements should be as low as possible. This is difficult for iron and mag-
nesium, in particular, because of their wide occurrence in mineral deposits.
Solvent extraction provides a possible process whereby a pure feed to
electrowinning could be obtained if a suitable extractant for copper was
available.
Until recently, the favored leaching acid had been sulfuric acid,
in which the metals exist as cationic species in solution. Examination
of selectivity series for the common extractants available in the 1950s
(Fig. 11.8) indicates that separation of copper from magnesium, for
example, would be easily achieved by carboxylic or phosphoric acids, but
with both these extractants, iron(III) would also be extracted with the
copper. Although it would be possible to precipitate iron as the hydroxide
by addition of alkali, this would complicate the process. However, as
copper extraction does not occur below pH 3, the addition of alkali is
necessary to achieve optimum extraction. This increases the overall costs
of the extraction circuit and prevents the direct return of the extraction
raffinate to leaching. Thus, it was obvious that to take advantage of the
benefits of the process, new extractants that could selectively extract
copper from iron at the pH of the leach liquor would be required. The
search for such reagents was based on extractants used in analyti-
cal chemistry, such as benzoin oximes, 8-quinolinol, and b-diketones (see
Table 11.3). The use of these reagents in commercial operations is
described individually.
474 Cox
Fig. 11.8 Extraction of metals from sulfate media by DEHPA. (From Ref. 4.)
Solvent Extraction in Hydrometallurgy 475
Table 11.5 Production Capacity of Acid Leach Copper SX-EW Plants (2001)a
Capacity
Name / location Leaching method (tons / year) Startup
USA
Nord Copper Johnson Camp Heap (oxide) 3,000 1990
Arimetco, NVb Heap (oxide) 2,500 1989
Asarco Ray, AZ Dump/Heap (oxide/sulf ) 41,000 1979
Asarco Silver Bell Heap/Dump (oxide/sulf ) 20,000 1997
BHP, Miami, AZ In situ/Agit. (sulf/tails) 12,500 1976
BHP, Pinto Valley, AZ Dump (oxide/sulf ) 8,750 1981
BHP, San Manuel Heap/In situ (oxide) 18,000 1986
Burro Chief, Tyrone, NM Heap/Dump (oxide/sulf ) 71,500 1984
Cyprus Bagdad. AZ Dump (oxide/sulf ) 14,000 1970
Cyprus Miami, AZ Heap (oxide/sulf ) 76,000 1979
Cyprus Sierrita, AZ Dump/Heap (oxide/sulf ) 26,000 1987
Cyprus Tohono, AZb In situ (oxide) 21,900 1981
Phelps Dodge, Chino, NM Heap (oxide/sulf ) 59,700 1988
Phelps Dodge, Morenci, AZ Heap/Dump (oxide/sulf ) 355,000 1987
Equitorial Mineral Park, AZ Dump (oxide/sulf ) 2,500 1992
Equatorial Tonopahb Heap (sulfide) 26,000 2000
Peru
Cyprus Cerro Verde Heap (oxide/sulf ) 73,000 1977
Southern Peru (Toquepala) Dump/Heap (oxide/sulf ) 56,000 1995
Centromin Mine Water (oxide/sulf ) 5,000 1980
Mexico
Cananea I and II, Mexico Dump (sulfide) 40,000 1980
La Caridad, Mexico Dump (oxide/sulf ) 24,000 1994
Canada
Gibraltar, Canadab Dump (oxide/sulf ) 5,475 1986
Africa
ZCCM, Zambia Agitation (oxide/tails) 90,000 1974
Sanyati Project (Reunion) Heap (oxide) 3,500 1995
Bwana Mkubwa Agitation (tails) 10,000 1998
Australia
BHAS Agitation (sulfide) 4,380 1984
Girilambone Heap (sulfide) 20,000 1993
Gunpowderc Heap (sulfide) 1990
Mt. Gordon Pressure (sulfide) 45,000 1998
Olympic Dam Agitation (oxide) 21,900 1989
Nifty Heap 16,500 1993
Cloncurryb Heap (oxide/sulf ) 1996
Murchison United (Mt.Cuthbert) Heap (oxide/sulf ) 5,500 1996
(continued)
478 Cox
CuðNH3 Þ2þ
4 þ 2HA ! CuA2 þ 2NHþ
4 þ 2NH3 (11.6)
Kinetics of extraction are fast and stripping is possible with solutions of low
acidity and high copper concentration, as found with spent tankhouse
liquors. A process using the reagent LIX54 for the recovery of copper from
printed circuit board etch liquors has been very successful commercially (see
Chapter 14). With these reagents, a higher copper transfer is achieved than
with the hydroxyoximes, as illustrated by a tenfold reduction in solvent
extraction plant size when changing from LIX64 to LIX54 for a leachate
from a lead dross process. Some problems have been found with the long-
term use of LIX54 in ammoniacal circuits due to the formation of a ketimine
[Eq. (11.7)] that reduces the capacity of the LIX54 and could not be ade-
quately removed by a normal clay cleanup treatment. Only one compound
was identified, suggesting that the reaction could be inhibited by steric
hindrance about the carbonyl group. Synthesis of compounds where the a-
hydrogen atoms on the C7H15 group were substituted by alkyl groups
removed this problem.
Unlike the amine extractants, this compound does not protonate easily and
provides excellent selectivity over those metals and metalloids that exist as
chloroanions in this medium. Extraction of the latter species becomes sig-
nificant only above 8 mol dm3 chloride, as protonation becomes significant.
This extractant [Acorga CLX 50], which contains 50% of the active ingre-
dient, has been proposed for an integrated leach, extraction, electrowinning
circuit (Cuprex process) [19].
480 Cox
Oxidation Major
Metal state chloro species Comments
Silver (Ag) I AgCl Insoluble
AgCl2 High HCl concentration
Gold (Au) I AuCl2 Very stable
III AuCl4
Platinum (Pt) II PtCl42 Conversion IV to II slow
IV PtCl62 Most common species, kinetically inert
Palladium (Pd) II PdCl42 Most common
IV PdCl62 Conversion II to IV difficult
Iridium (Ir) III IrCl63 Both species stable and
IV IrCl62 conversion IV to III easy
Rhodium (Rh) III RhCl63
Ruthenium (Ru) III RuCl63 Complex equilibria between III and IV
IV RuCl62 depends on redox potential and
chloride concentration
Osmium (Os) III OsCl63
IV OsCl62 Os(IV) more stable than Ru(IV)
Ir(IV), Pt(IV), with the states from Rh(III) being termed inert. Thus,
kinetic factors tend to be more important, and reactions that should be
possible from thermodynamic considerations are less successful as a result.
On the other hand, the presence of small amounts of a kinetically labile
complex in the solution can completely alter the situation. This is made even
more confusing in that the basic chemistry of some of the elements has not
been fully investigated under the conditions in the leach solutions. Conse-
quently, a solvent extraction process to separate the precious metals must
cope with a wide range of complexes in different oxidation states, which
vary, often in a poorly known fashion, both in kinetic and thermodynamic
stability. Therefore, different approaches have been tried and different flow
sheets produced.
It can be seen that the three commercial flow sheets (Figs. 11.10–11.12)
do have common features. Hence, because gold and silver can be easily
separated from the remaining precious metals, they tend to be removed from
the flow sheet at an early stage. In addition, significant quantities of gold
and silver exist, in either primary ores or waste materials that are not
associated with the platinum group metals. Silver often occurs as a residue
from the chloride leaching operation and so is removed first, followed by
gold, which can be extracted as the tetrachloroauric(III) acid, HAuCl4, by a
solvating reagent.
482 Cox
Fig. 11.11 Matthey Rustenberg flow sheet for extraction of precious metals.
484 Cox
Notice that none of the flow sheets uses solvent extraction exclusively.
Because the aqueous chemistry of osmium and ruthenium is very complex,
most operators remove these elements by distillation of the tetraoxides,
MO4. Also, it has been advantageous to use ion exchange to separate and
concentrate rhodium. The various extraction routes for individual elements
are discussed in the following sections.
11.5.1 Gold
Gold exists in the leach solution as AuCl 4 , and all the processes rely on the
extraction of this ion or the parent acid by ion-pair or solvating extractants,
respectively. The early studies by Morris and Khan [20] led to the adoption
of dibutylcarbitol (diethyleneglycol dibutylether) by INCO in a gold refinery
several years before the complete solvent extraction refining flow sheet was
produced. The extractant has a high selectivity for gold (Fig. 11.13) and
coextracted elements can be easily scrubbed with dilute hydrochloric acid
[22]. Because of the high distribution coefficient, recovery of gold is best
achieved by chemical reduction. Thus, oxalic acid at 908C produces easily
filtered gold grain. However, the slow kinetics necessitate a batch process.
An alternative solvating extractant, 4-methyl-2-pentanone (methyl iso-
butylketone; MIBK) is used by Matthey Rustenberg Refiners (MRR). This
has a lower selectivity for gold (Fig. 11.14), so a simple acid scrub is not
sufficient to remove impurities [22]. However, there is an advantage to the
overall flow sheet in that some of the base elements, such as tellurium,
arsenic, and iron, are removed at an early stage. Again, gold is recovered by
chemical reduction with, for example, iron to give an impure product for
subsequent refining. A third type of solvating extractant, long-chain alco-
hols have also been proposed, the most effective being 2-ethylhexanol,
which can be easily stripped at ambient temperature with water [24]. Gold,
like silver, being classified as a soft acid, should prefer interaction with soft
donor atoms like sulfur. Therefore, it is no surprise to find that extraction
with the trialkylphosphine sulfide, CYANEX 471X, is very efficient and
provides excellent separation from elements such as copper, lead, bismuth,
and other metalloids found in anode slimes [22]. Stripping can be achieved
using sodium thiosulfate, which allows the separation of recovery and the
reduction to the metal.
Amine salts can also be used to extract anionic species, but, in general,
these are less selective for gold. However, Baroncelli and coworkers [25]
have used trioctylamine (TOA) to extract gold from chloroanions of, for
example, copper, iron, tin, and zinc, in aqua regia solution. The selectivity is
attributed to the presence of TOA HNO3 in the organic phase. Coextracted
metals are scrubbed with dilute nitric acid, and gold is recovered by
486 Cox
a hydrochloric acid-thiourea solution. The same authors also used N,N0 -di-
n-butyloctanamide as a solvating extractant with aqua regia solutions. The
protonated forms of Kelex 100 and LIX26, both acting as ion pair extrac-
tants, have been proposed by Demopoulos et al. [26] for gold extraction,
with recovery as gold grain by hydrolytic stripping.
Alkaline cyanide solutions are another common lixiviant for gold,
giving complex cyano species in solution. These anionic complexes may be
Solvent Extraction in Hydrometallurgy 487
Fig. 11.14 Extraction of elements from chloride medium by MIBK. (From Ref. 23.)
488 Cox
11.5.2 Silver
Although silver is not treated by solvent extraction in any of the flow sheets,
silver is recovered from aqueous solution in several other situations. For
these processes, Cytec developed reagents with donor sulfur atoms to extract
this ‘‘soft’’ element. For example, tri-isobutylphosphine sulfide (CYANEX
471X) extracts silver from chloride, nitrate, or sulfate media selectively from
copper, lead, and zinc [32]. The silver is recovered from the loaded organic
phase by stripping with sodium thiosulfate, and the metal recovered by
cementation or electrolysis. Silver can also be extracted from chloride
solution by a dithiophosphinic acid (CYANEX 301) [33].
4 þ 2R2 S
PdCl2 ! ðR2 SÞ2 PdCl2 þ 2Cl (11.9)
ðR2 SÞ2 PdCl2 þ 4NH3 ! PdðNH3 Þ2þ
4 þ 2R2 S þ 2Cl (11.10)
Neutralization of the strip solution with hydrochloric acid gives Pd(NH3)2-
Cl2 as product. One of the problems that has emerged is the formation of di-
n-hexylsulfoxide [34] by oxidation of the sulfide. This may cause several
problems including extraction of iron(III) that is strongly dependent on the
HCl concentration. The iron can easily be stripped by water. There have also
been indications of a buildup of rhodium in the extract phase that again can
be explained by the extraction of anionic rhodium species by the sulfoxide.
One benefit from the presence of the sulfoxide is that the rate of palladium
extraction is increased by the presence of the protonated sulfoxide at high
acidities; however, this kinetic enhancement is less that found with
TOA HCl, which remains protonated even at low acidities.
In this flow sheet, the aqueous raffinate from extraction is acidified to
5–6 mol dm3 with hydrochloric acid to optimize platinum extraction by the
solvating extractant TBP. The coextraction of iridium is prevented by
reduction with sulfur dioxide, which converts the iridium(IV) to the (III)
species, which is not extractable. Once again, kinetics are a factor in this
reduction step because, although the redox potentials are quite similar,
[Ir(IV)/(III) 0.87 V; Pt(IV)/(II) 0.77 V], iridium(IV) has a relatively labile
d 5 configuration, whereas platinum(IV) has the inert d 6 arrangement. The
species H2PtCl6 is extracted by TBP, from which platinum can be stripped
by water and recovered by precipitation as (NH3)2PtCl2.
The second selective extractant route (that of MRR) uses a hydroxy-
oxime to remove palladium. The actual compound used has not been spe-
cified, but publications refer to both an aliphatic a-hydroxyoxime and an
aromatic b-hydroxyoxime. The a-hydroxyoxime LIX63 has the faster
extraction kinetics, but suffers from problems with stripping. For the b-
hydroxyoxime, a kinetic accelerator in the form of an amine (Primene
JMT*) has been proposed. The precise mode of operation of this accelerator
is unknown, but it may be a similar process to that proposed for the sulfide
4 þ R3 NHCl
PdCl2 ! ðR3 NHÞ2 PdCl4 þ 2HA
! PdA2 þ 2R3 NHCl þ 2Cl (11.11)
Good loading of palladium and palladium–platinum separation fac-
tors of 103–104 are found. Stripping of the palladium is easy with hydro-
chloric acid, and (NH3)2PdCl4 is obtained by precipitation. The raffinate
from palladium extraction (see Fig. 11.10) is treated to remove osmium and
ruthenium and, again, sparged with sulfur dioxide to reduce iridium(IV),
following PtCl62 species removed by an amine. Stripping of the platinum is
difficult because of the high-distribution coefficient, but techniques that
have been successful include the use of strong acids, such as nitric or per-
chloric, to break the ion pair, or alkali to deprotonate the amine salt. The
choice depends very much on the amine chosen; for example, if a quaternary
amine is used, the deprotonation reaction is not available, and a strong acid
has to be used. On the other hand, with other amines, the possibility of
+4(6x)
forming inner sphere complexes such as Pt(RNH2)xCl(6x) is an
important consideration, as these are nonextractible and cause a lockup of
platinum in the organic phase. The formation of such complexes is enhanced
by a deprotonation step that releases the free amine into the organic phase.
As the ease of formation of platinum–amine complexes decreases in the
order primary>secondary>tertiary, the reason for the choice of the latter
for platinum extraction can be understood.
The phosphine sulfide CYANEX 471X has been proposed as an
alternative to the foregoing reagents for palladium–platinum separation.
Here, the faster extraction of palladium affords separation factors of 10:1,
with a 10 min contact time. The chemistry of the process is similar to that of
the alkyl sulfides. Various other separations have been published; for
example, TBP will extract platinum(IV) chloroanions preferentially to pal-
ladium(II).
An alternative general process involves coextraction of palladium and
platinum, followed by selective stripping (see Fig. 11.12). A novel amino acid
extractant, made by the reaction between chloracetic acid and Amberlite*
LA-2, a secondary amine, is used to extract the two elements from the leach
liquor. However, in the given flow sheet, no selective stripping is employed,
both elements being stripped by hydrochloric acid. The resulting chlor-
oanions are then separated by using di-n-hexylsulfide to extract PdCl42.
4 þ 2H2 ACl
PdCl2 ! PdA2 þ 6Cl þ 4Hþ (11.16)
6 þ 2H2 ACl
PtCl2 ! ðH2 AÞ2 PtCl6 þ 2Cl (11.17)
The reaction to form the palladium complex is similar to that reported for
amine salts, although here, because a bidentate chelating ligand is used, no
chlorine atoms are retained in the complex, and the system is easy to strip.
Also, as both reactions involve initial ion pair extraction, fast kinetics are
observed with 3–5 min contact time to reach equilibrium at ambient tem-
perature. The extraction conditions can be easily adjusted in terms of acidity
to suit any relative metal concentrations and, because the reagent is used in the
protonated form, good selectivity over base metals, such as iron and copper,
*Trademark of Cognis.
492 Cox
must precede the extraction of platinum and palladium, as with other sys-
tems. The platinum–palladium separation factors depend on the amide
structure, but are rather small, about 25–55, so that scrubbing of the loaded
organic phase is required to give a good separation [36].
used commercially, but it suffered from the restrictions of low resin capacity
that required large columns, and also the system had to use a batch opera-
tion. This process has now been largely superceded by solvent extraction.
Both ion exchange and solvent extraction separations are based on the steady
decrease in size across the lanthanide elements, which result in an increase in
acidity or decrease in basicity with increasing atomic number. This causes a
variation in formation coefficient of any metal–extractant complex, allowing
preferential binding to an ion exchange resin, or extraction of the complex
into an organic phase (see Chapters 3 and 4). The variations in formation
coefficients and hence, separation factors, between adjacent elements, are
small; so many equilibria and thus a large number of extraction units are
required. Even then, separation is possible only by recycling both the exit
streams. This configuration, rarely used outside lanthanide production, is
essential to the preparation of pure products or the separation of adjacent
elements. The operation is shown in Fig. 11.17. The bank of mixer-settlers is
fed close to the midpoint, with a mixture of components, and by returning
the exit streams to the process, 100% reflux is obtained. The components are
force-fed to build up their concentration, and the reflux is continued until
the desired degree of separation is achieved. Then, the operation is stopped
and the individual mixer-settlers emptied to give the required products.
Solvent Extraction in Hydrometallurgy 495
Fig. 11.18 Concentration profiles for mixer-settler bank under total reflux: system TBP
(50%) from nitrate medium. (From Ref. 42.)
been studied to a small extent and these show some processing advantages
with, for example, an organophosphonate showing higher separation factors
than TBP.
Table 11.8 Separation Factors for Adjacent Lanthanides Using TBP from Nitrate
Medium
Separation factors
Aqueous phase
conc. (Ln2O3 g dm3) Sm/Nd Gd/Sm Dy/Gd Ho/Dy Er/Ho Yb/Er
460 2.26 1.01 1.45 0.92 0.96 0.81
430 - - 1.20 0.96 0.65 -
310 2.04 1.07 1.17 0.94 0.82 -
220 1.55 0.99 1.08 0.89 0.78 -
125 1.58 0.82 0.92 0.83 0.72 -
60 1.40 0.78 0.89 0.77 0.70 0.63
Fig. 11.19 Distribution coefficients for lanthanide extraction with Aliquat 336 from
nitrate (solid squares) and thiocyanate (solid circles) media. (From Ref. 1.)
Ce/La Pr/Ce Nd/Pr Sm/Nd Eu/Sm Gd/Eu Tb/Gd Dy/Tb Ho/Dy Er/Ho Tm/Er Yb/Tm Lu/Yb
a
Versatic 911 3.00 1.60 1.32 2.20 - 1.96 - - 1.17 1.28 - - -
Solvent Extraction in Hydrometallurgy
DEHPAa - - 1.40 4.00 2.30 1.50 5.70 2.00 1.80 2.40 3.40 - -
DEHPAb 5.24 1.86 1.32 9.77 2.14 1.99 - - - - - - -
Phosphonic acidb 13.8 3.47 1.51 13.80 2.45 1.86 6.92 3.23 2.17 2.82 2.63 3.71 1.74
Phosphinic acidb 16.98 1.86 1.32 13.80 2.14 1.32 6.02 2.82 1.74 2.63 3.47 2.63 1.51
a
Source: Ref. 42.
b
Source: Ref. 43. (All organophosphorus extractants used in this work had 2-ethylhexyl substituents.)
499
500 Cox
Fig. 11.20 Effect of DEHPA concentration on extraction of lanthanides. (From Ref. 2.)
(0.2 mol dm3) than either CYANEX 272 (0.08 mol dm3), DEHPA, or the
phosphonic acid ester (P507) (0.17 mol dm3) [45]. In addition, stripping from
the synergistic mixture with either hydrochloric or nitric acid is easy. The
latter enables the production of high-purity products free from chloride ions.
11.6.4 Summary
Liquid–liquid extraction provides one of the easiest methods for the
separation of these commercially important elements. It is more convenient
than ion exchange, allowing both semicontinuous operation and the use of
more concentrated feed solutions. However, the large number of mixer-
settlers required imposes a considerable capital investment in the plant, and
their use in the reflux mode entails that some elements are locked into the
process for long periods, which also has economic implications. Therefore,
flow sheets are developed to provide elements with a ready market at the
earliest opportunity, leaving the remaining elements as intermediates for
separation at a later stage or for sale as mixed lanthanides.
The choice of reagent depends very much on the nature of the feed, but
the introduction of organophosphinic acids either alone or in combination
with carboxylic acids in a synergistic mixture offer great potential for
commercial development.
Further discussion on the extraction of lanthanides, especially con-
cerning their separation from actinides, can be found in Chapter 12.
11.7 CONCLUSION
Solvent extraction is now a proven technology for the commercial extrac-
tion, separation, and concentration of a wide range of metals both from
primary and secondary sources (see Chapter 14). In recent years, there has
been a reduction in the development, production, and marketing of new
commercial extractants as the overall costs of such activities increases.
However, the use of established reagents in new hydrometallurgical appli-
cations continues to expand.
REFERENCES
1. Lo, T. C.; Baird, M. H. I.; Hanson, C. Handbook of Solvent Extraction; Wiley-
Interscience: New York, 1983.
2. Ritcey, G.; Ashbrook, A. W. Solvent Extraction, Part 2; Elsevier; Amsterdam,
1979.
3. Ritcey, G.; Ashbrook, A. W. Solvent Extraction, Part 1; Elsevier; Amsterdam,
1984.
4. Cox, M.; Flett, D. S. Chem Ind., 1987, p. 188.
Solvent Extraction in Hydrometallurgy 503
BIBLIOGRAPHY
In addition to the list at the end of Chapter 1, texts concerned with appli-
cations of solvent extraction in hydrometallurgy include the following:
Li, N.N. and Navratil, J.D. Eds.; CRC Press: Boca Raton, Florida,
1986, Vol. 8, 1 p.
Osseo-Asare, K.; Miller, J. D. Eds. Hydrometallurgy; Research, Development
and Plant Practice; Metallurgical Society AIME; New York, 1982.
Proceedings of Extraction Metallurgy Conferences, published by Institution
of Mining and Metallurgy, London, 1981, 1985, 1989.
Schulz, W. W.; Navratil, J. D. Eds. Science and Technology of Tributyl
Phosphate, Vols. 1 and 2; CRC Press: Boca Raton, Florida, 1987.
Szymanowski, J. Hydroximes and Copper Hydrometallurgy; CRC Press:
Boca Raton, Florida, 1993.
Yamada, H.; Tanaka, M. Solvent extraction of metal carboxylates. Adv.
Inorg. Chem. Radiochem. 1985, 29, 143.
12
Solvent Extraction in Nuclear Science
and Technology
CLAUDE MUSIKAS* Commissariat à l’Energie Atomique, Paris, France
WALLACE W. SCHULZ Consultant, Albuquerque, New Mexico, U.S.A.
JAN-OLOV LILJENZIN* Chalmers University of Technology, Göteborg,
Sweden
12.1 INTRODUCTION
In the year 2000, 15% of the world’s electric power was produced by
433 nuclear power reactors: 169 located in Europe, 120 in the United States,
and 90 in the Far East. These reactors consumed 6,400 tons of fresh enri-
ched uranium that was obtained through the production of 35,000 tons of
pure natural uranium in 23 different nations; the main purification step was
solvent extraction. In the reactors, the nuclear transmutation process yielded
fission products and actinides (about 1000 tons of Pu) equivalent to the
amount of uranium consumed, and heat that powered steam-driven turbines
to produce 2,400 TWh of electricity in 2000.
Different countries have adapted various policies for handling the
highly radioactive spent fuel elements:
1. in the once through cycle, the fuel elements are stored in temporary
facilities, usually with the intention to finally deposit them in under-
ground vaults; and
2. in the recycling concept, the spent fuel pins are dissolved in acid and
‘‘reprocessed’’ to recover energy values (i.e., unused uranium and other
fissionable isotopes, notably plutonium), while the waste is solidified
and stored in underground vaults.
In some countries, the main purpose of reprocessing is to recover
plutonium for weapons use. The main separation process in all known
reprocessing plants is solvent extraction.
*Retired.
507
508 Musikas et al.
12.2 BACKGROUND
12.2.1 Historical Aspects
Nuclear science and technology have played key roles in development of
liquid–liquid solvent extraction for the industrial-scale separation of metals.
The high degree of purity of materials needed for nuclear applications,
together with the high value of these materials, favors solvent extraction as
the separation method of choice when compared with other methods. The
need for a continuous, multistage separation system is unique to reproces-
sing of spent nuclear reactor fuels because of the requirement for nearly
quantitative recovery and separation of fissile uranium and plutonium from
some 40 or more fission products belonging to all groups of the periodic
table. Only a brief account of the historical development and application of
solvent extraction processes and technology in the nuclear field is included
here. Numerous other authors have reviewed various aspects of this history;
readers interested in historical details should consult references [1–8].
In 1942, the Mallinckrodt Chemical Company adapted a diethylether
extraction process to purify tons of uranium for the U.S. Manhattan Project
[2]; later, after an explosion, the process was switched to less volatile ex-
tractants. For simultaneous large-scale recovery of the plutonium in the
spent fuel elements from the production reactors at Hanford, United States,
methyl isobutyl ketone (MIBK) was originally chosen as extractant/solvent
in the so-called Redox solvent extraction process. In the British Windscale
plant, now Sellafield, another extractant/solvent, dibutylcarbitol (DBC or
Butex), was preferred for reprocessing spent nuclear reactor fuels. These
early extractants have now been replaced by tributylphosphate [TBP], diluted
in an aliphatic hydrocarbon or mixture of such hydrocarbons, following the
discovery of Warf [9] in 1945 that TBP separates tetravalent cerium from
Solvent Extraction in Nuclear Science 509
Fig. 12.1 Oxidation states of the actinide elements: . most stable ions in aqueous
solutions; ++ oxidation states observed in aqueous solutions; + , unstable ions observed
only as transient species. *In solids precipitated from alkaline solutions.
than oxygen have more covalency. In addition, the trivalent actinides have
slightly more covalency than the lanthanides [19]. This has led to the search
for new extractants containing more covalent donor atoms, such as sulfur or
nitrogen. Such extractants have been shown to provide better separation of
the trivalent ions of the 4f and 5f transition series (see section 12.9.1).
After the separation of the actinides from the high-level waste, it is
desirable to remove certain other fission products from the nuclear wastes.
Some Cs+ and Sr2+ are low-charged cations that react well with macro-
cyclic ligands (e.g., crown ethers, calixarenes). Research to synthesize and
investigate the properties of macrocyclic ligands for application in nuclear
waste treatment has been an active effort internationally. Some of the results
obtained are discussed in section 12.7.
Extractants that lead to the formation, in the organic phase, of com-
plexes in which the metal is completely surrounded by oxygen atoms are
often quite selective for the actinides. This is true with TBP, which selec-
tively extracts actinides from nitrate solutions, forming complexes such as
Th(NO3)4(TBP)2 and UO2(NO3)2(TBP)2 (see Fig. 12.2). Nitric acid is pre-
ferred over other strong inorganic acids because it is the least corrosive acid
for the stainless steel equipment used in large-scale reprocessing of spent
nuclear fuels. Its low complexing affinity for the d transition ions and its
ability to fill two coordination sites favors binding with large actinide ions.
Fig. 12.2 Distribution ratios of actinide ions and some selected metals between un-
diluted TBP and HNO3 or HCl solutions.
Table 12.2 Chemical Basis for the Amex and the Dapex Processes for U(VI) Recovery from H2SO4 Leach Liquors
Amex Dapex
Point of
comparison Amex process Dapex process
widely used than the Dapex process because of the greater selectivity of
trialkylamines than HDEHP for uranium in H2SO4 solution.
Table 12.4 Main Industrial Processes for the Recovery of U from Wet Phosphoric Acid
Extractant Mechanism of extraction Stripping Remark
Today, phosphoric acid for use in the pharmaceutical or food industry must
be free from uranium and other harmful metals. In the future, the processing
of phosphate rock for production of fertilizers may be an important source
of uranium because uranium removal may become compulsory to avoid its
dissemination into the environment from the use of phosphate-based fertil-
izers.
Fig. 12.3 Developed formula for calixarene(6) selective extractants for U(VI) extraction
from sea water. (From Ref. 23.) (1) Carboxylate functions; (2) hydroxamate functions.
Note: The Organic Phase is 104 mol dm3. Calixarene in CHCl3. Aqueous
Phase is 2 105 mol dm3 UO2(CH3COO)2 Buffered to pH 5.9 by
102 mol dm3 acetate. Phase Volume Ratio Org/Aq = 0.2.
520 Musikas et al.
Table 12.6 Commercial Processes for the Recovery of Th from Its Main Ores
Types of ores Ore
(main locations) composition Leaching Purification
Fig. 12.4 Simplified flow sheet used in the recovery of thorium from its ores.
522 Musikas et al.
Table 12.7 Composition of Dissolver Solution from Several Reactor Fuel Types
Graphite Pressurized Fast breeder
Fuel type characteristics gaseous reactor water reactor reactor
Irradiation MW day per metric ton 4,000 33,000 80,000
Activity b,g after 150 days of cooling 1,400 4,500 18,000
(Ci per kg of fuel)
U(VI), mol dm3 1.05 1.05 0.85
Pu(IV), mol dm3 0.00275 0.01 0.22
HNO3 mol dm3 3 3 3
Zr, mol dm3 0.0013 0.0095 0.0184
Ru, mol dm3 0.0005 0.0058 0.02
Fission products (kg per metric ton) 4.16 35 87
Transplutonium elements, mol dm3 0.0002 0.0028
retention of metals and use of less concentrated Na2CO3 in the solvent wash,
which helps to limit the amount of sodium salts that must be vitrified.
12.5.1.3 Purex Process Improvements
Reprocessing of spent reactor fuels has been the subject of much interna-
tional debate [34]. Strategies ranging from no reprocessing at all to complete
reprocessing with, in some cases, recovery and transmutation of long half-
life nuclides into stable or shorter half-life nuclides, have been proposed.
Several countries, such as the United Kingdom, France, and Japan, have
adopted an intermediate policy that consists of recovery of fissile material
from spent fuels and storage of the highly radioactive wastes after vitrifi-
cation in deep geological repositories.
To increase the economic benefits of reprocessing, some improvements
in the Purex process appear to be highly desirable. Two approaches can be
followed:
1. make changes within the Purex process, or
2. develop new processes that are more advantageous.
The chemistry of the Purex process has been extensively investigated,
and it seems very difficult to find new major improvements. However, some
improvements in some process details may contribute to a decrease in re-
processing costs, especially if several changes are made simultaneously.
Examples of such changes are presented in next sections.
Solvent Extraction in Nuclear Science 525
Fig. 12.7 Volumes of wastes containing long half life isotopes generated by the re-
processing of spent nuclear fuels as a function of time. (From Ref. 38.) Comparison with
the nonreprocessing option.
528 Musikas et al.
Fig. 12.8 Flow sheet of the first purification cycle of U and Pu in a process using an
N,N-dialkylamide as extractant. (DOBA is C3H7CON(CH2CHC2H5C4H9)2; DOiBA,
(CH3)2CHCON(CH2CHC2H5C4H9)2; diluent TPH (branched dodecane).
Fig. 12.9 Two-stage acid Thorex process for highly irradiated fuels. Numbers in the frames indicate stage number, whereas numbers on
Musikas et al.
the lines indicate flow volumes relative to the feed volume (DOD and FP are dodecane and fission products).
Solvent Extraction in Nuclear Science 531
Thorex two-stage
Fig. 12.10 Generic Truex process flow sheet for use with nitric acid wastes (NPH is a
mixture of aliphatic hydrocarbons).
Fig. 12.11 Distribution ratios of various metallic species between 0.5 mol dm3
(C4H9CH3NCO)2CHC2H4OC2H4OC6H13 [DMDB(2-3,6-OD,1,3-DA)P] t-butylbenzene solu-
tions and HNO3 solutions.
of Pu(IV) and U(VI) from Diamex will not be a difficult task. In addition,
the degradation products of the solvent do not retain the actinides in the
organic phase, as happens with the organophosphorus extractants.
Inactive metals in the wastes, such as Fe(III), Mo(VI), and Zr(IV), are
retained in the aqueous feed by addition of suitable quantities of oxalic acid.
As with CMPO, some Ru is extracted. Currently, the extraction of Pd, Tc,
Np and their poor stripping remain a problem for which process mod-
ifications are necessary. The flow sheet in Fig. 12.13 and Table 12.11 has
536 Musikas et al.
of trivalent actinides (e.g., 241Am) into TBP and DBBP are only high
enough for practical applications from either highly salted or low-acid
solutions, or very concentrated (i.e., 15–16 mol dm3 HNO3) solutions.
Actual plant-scale experience with DBBP was very disappointing [56]
because of the inability to adequately control the feed pH values in the range
required for effective americium extraction.
The use of HDEHP for extraction of actinides from waste solutions
also has several drawbacks, including extraction of trivalent ions at a pH
at which tetravalent ions such as Zr(IV) and Pu(IV) are hydrolyzed,
and difficulties in stripping tetravalent and hexavalent actinides. All in all,
monofunctional organophosphorus reagents are vastly inferior to their
bifunctional counterparts for extracting Am(III) and other actinides from
strong HNO3 media.
Many of these drawbacks were circumvented in a process developed at
Chalmers University, which was successfully demonstrated in continuous
operation using the old high-level raffinate concentrate from Purex pro-
cessing of low burn-up fuel. Although operation was very easy and stable,
the process flow sheet was complicated using bromoacetic acid, HDEHP,
TBP, and lactic acid [57–59].
Table 12.11 Decontamination Factors for Extraction (Mass in the Feed to Mass in
Raffinate) and for the Stripping [(Mass in the Feed Minus Mass in the Raffinate)/Mass in
the Organic Out] in DIAMEX Hot Test with Flow Sheet in Fig. 12.13
Element Extraction Stripping Element Extraction Stripping
a
Zr(IV) 1 - Mo(VI) 1 -a
b
Tc(VII) 120 30 Ru(II) 1.3 61
Pd(II) 157 96 La(III) 167 >46000
Ce(III) 211 >22500 Pr(III) 216 >44000
Nd(III) 157 >50000 Sm(III) 110 -a
Eu(III) 113 >300000 Gd(III) 49 >300000
Np(V, VI) 29 29b Am(III) 258 >25100
Cm(III) 301 3150
a
Below ICP-MS determination limit.
b
Accumulation in the back extraction section.
538 Musikas et al.
Fig. 12.13 Schematic drawing for the hot test of the DIAMEX process in a 16-stage
centrifugal extractor battery. (From Ref. 53.)
provides such removal from highly acidic solutions. The process employs an
n-octanol solution of the commercially available macrocyclic ether (cf.
Appendix D, Fig. 21), di-tert-butyldicyclohexyl-18-crown-6 (DtBuCH18C6)
as the extractant for 90Sr. Batch contacts with simulated waste solutions
show that the SREX process solvent is highly selective for strontium; of the
many contaminants only barium and technetium coextract with strontium to
any extent. Irradiation tests confirm that the SREX process solvent is
satisfactorily resistant to radiolysis and hydrolytic attack is also minimal.
The SREX process has been successfully tested in both batch contacts and
continuous countercurrent tests with simulated waste solutions. An overall
90
Sr decontamination factor (SREX process feed to SREX process raffinate)
of 4600 was achieved from Truex wastes.
In the past, the extraction of 90Sr and 137Cs was investigated for some
practical applications (heat and gamma ray sources) but the main interest
today is for decreasing the thermal power and the potential hazard of
nuclear waste in underground repositories. Results of extractions with some
Solvent Extraction in Nuclear Science
Fig. 12.14 Schemes for partitioning Purex high-level waste. (a, data from Ref. 54; b, data from Ref. 55.)
539
540 Musikas et al.
other macrocyclic extractants are collected in Table 12.12, and the formulae
of some of the extractants are shown in Fig. 12.15.
0.01 mol dm3 dicarbollide (CoDC) in nitrobenzene 1 mol dm3 HNO3 + 0.03 mol dm3 Sr2 + 15.3 [61]
polyethylene glycol
Mn + + nHCoDC ! M(CoDC)n + nH + 0.1 mol dm3 HNO3 + 0.03 mol dm3 Sr2 + 569 [61]
polyethylene glycol
0.01 mol dm3 dicarbollide (CoDC) 0.5 mol dm3 HNO3 Cs + 22 [61]
in nitrobenzene 0.1 mol dm3 HNO3 Cs + 102.6
1 mol dm3 TBP + 0.1 mol dm3 HDNNS + 0.05M crown 1(C1) 3 mol dm3 HNO3 Sr2 + 0.06 [62]
into kerosene: Mn + + m(HNO3)C1 + nHDNNS ! 0.5 mol dm3 HNO3 Cs + 1.59
M(C1)m(DNNS)n + nH + + mHNO3 Cs + 1
Solvent Extraction in Nuclear Science
1 mol dm3TBP + 0.1 mol dm3 HDNNS + 0.05 mol dm3 3 mol dm3 HNO3 Sr2 + 1.98 [62]
crown 2(C2) in kerosene: Mn + + m(HNO3)C2 + nHDNNS ! Cs + *0
M(C2)m(DNNS)n + nH + + mHNO3 0.5 mol dm3 HNO3 Sr2 + 2.57
2
10 mol dm3 di n-octyloxy calix[4]arene crown 6 1 mol dm3 HNO3 Cs + 33 [63]
(CA1) in orthonitrophenyl hexyl ether:
Csþ þ NO 3 þ CA1 ! CsNO3 CA1
102 mol dm3 di n-octyloxycalix[4]arene crown 6 1 mol dm3 HNO3+ Cs + 25 [63]
(CA1) in orthonitrophenyl hexyl ether: 4 mol dm3 NaNO3
Csþ þ NO 3 þ CA1 ) ! CsNO3 CA1
102 mol dm3 benzyloxycalix[8]arene N-diethylamide 1 mol dm3 HNO3 Sr2 + 20 [63]
(CA2) into orthonitrophenyl hexyl ether: Cs+ <.001
Sr2þ þ 2NO 3 þ CA2 ! SrðNO3 Þ2 CA2
102 mol dm3 benzyloxycalix[8]arene N-diethylamide 1 mol dm3 HNO3+ Sr2 + 5 [63]
(CA2) in orthonitrophenyl hexyl ether: 4 mol dm3 NaNO3 Cs + 0.1
Sr2þ þ 2NO 3 þ CA2 ! SrðNO3 Þ2 CA2
541
Fig. 12.15 Formulae of the ligands used for Cs(I) or Sr(II) extraction from acidic media
and cited in Table 12.12.
542
Fig. 12.16 Flow sheets used for separation of transplutonium elements from irradiated Pu: (a) France (Fontenay-aux-Roses); (b) United States
(Oak Ridge National Laboratory). (Data from (a) Ref. 64; (b) Ref. 65.)
543
544 Musikas et al.
Fig. 12.17 Determination of the oxidation state of submicro traces of 256Lr (open cir-
cles) and 255No (solid circles). Solvent: 0.2 mol dm3 TTA–MIBK; aqueous phases: buf-
fers at various pHs. (From Ref. 67.)
Fig. 12.19 Distribution ratios of Am(III) and Eu(III) between 1 mol dm3 a-bromocapric
acid + 0.02 mol dm3 nitrogen ligand into TPH as a function of aqueous HNO3 con-
centration. (From Ref. 52.)
tested, but their low solubility in TPH is another drawback [53]. Altogether
78 nitrogen donor molecules were synthetized. Among them are the tri-
dentate nitrogen donors from the family of the 2,6-bis-(5,6-dialkyl-1,2,4-
triazine-3-yl)-pyridine prepared by Kolarik et al. [78].
Sulfur donors belonging to the dialkyl-thiophosphorous acids were
also investigated, and nine molecules have been considered. Separation
factors between Am(III) and Eu(III) as high as 17,000 have been found and
confirm the ability of sulfur donors to promote intergroup 5f-4f separations
of trivalent ions. As indicated by the Pauling electronegativities of N and S,
the S donors should bind better to the 5f trivalent ions than the N donors.
With the individual sulfur donors or synergistic mixtures investigated, the
extraction reactions were:
Fig. 12.21 Flow sheet for the hot test of SANEX 3 process using n-Pr-BTP.
Solvent Extraction in Nuclear Science 551
12.10 SUMMARY
Solvent extraction technology has an unchallenged role in the nuclear fuel
cycle and such technology has been aggressively developed and implemented
over the last five decades to a fully mature and complete status. Incentives,
either technical or economic, to make any short-term changes or perceived
improvements to present advanced engineering-scale fuel cycle practice are
not apparent and likely do not exist. Long-term changes in fuel processing
strategies, e.g., implementation and transmutation of actinides and selected
fission products, have been suggested by some. But even in this case, workable
Solvent Extraction in Nuclear Science 553
REFERENCES
1. Schulz, W.; Burger, L. L.; Navratil, J. D. Eds. Science and Technology of
Tributyl Phosphate, Vol. 3, Applications of Tributyl Phosphate in Nuclear Fuel
Processing; CRC Press: Boca Raton, Florida, 1990.
2. Hewlett, G.; Anderson, O. E. The New World 1939/1946; Pennsylvania State
University Press: State College, Pennsylvania, 1962.
3. Long, T. Engineering for Nuclear Fuel Reprocessing; American Nuclear Society:
La Grange Park, Illinois, 1978.
4. Wymer, R. G.; Vondra, B. L. Eds. Light Water Nuclear Reactor Fuel Cycle;
CRC Press: Boca Raton, Florida, 1981.
5. Stoller, S. M.; Richards, R. B. Eds. Reactor Handbook, 2nd Ed.; Interscience
Publishers: New York, 1961.
6. Lo, T. C.; Baird, M. H. I.; Hanson, C. Eds. Handbook of Solvent Extraction;
John Wiley and Sons: New York, 1983.
7. McKay, H. A. C.; Miles, J. H.; Swanson, J. L. Science and Technology of
Tributyl Phosphate; Schulz, W. W., Burger, L. L., Navratil, J. D., Eds.; CRC
Press: Boca Raton, Florida, Vol. 3, Chap. 1, 1990.
8. Danesi, P. R. Development in Solvent Extraction Alegret, S. Ed.; Ellis Horwood:
Chichester, U.K., 1988; Chap. 12.
9. Warf, J. C. J. Am. Chem. Soc., 1949, 71, 3257.
10. McKay, H. A. C. Science and Technology of Tributyl Phosphate; Schulz, W. W.
Navratil, J. D. Eds.; CRC Press: Boca Raton, Florida, 1990; Vol. 1, Chap. 1.
11. Smith, L. E.; Page, J. E. J. Chem. Soc., 1948, 67.
12. Coleman, C. F.; Brown, K. B.; Moore, J. G.; Crouse, D. J. Ind. Eng. Chem.,
1958, 50, 1756.
13. Chesne, A.; Koehly, G.; Bathellier, A. Nucl. Sci. Eng., 1963, 17, 557.
554 Musikas et al.
35. Kolarik, Z.; Shuler, R. Proceedings of Extraction’84; Inst. Chem. Eng. Sym-
posium Series No. 88, Pergamon Press: London, 1984; p. 83.
36. Drake, V. A. Science and Technology of Tributyl Phosphate; Schulz, W. W.;
Burger, L. L.; Navratil, J. D. Eds.; CRC Press: Boca Raton, Florida, 1990;
Vol. 3, Chap. 3.
37. Baron, P.; Boullis, B.; Germain, M.; Gué, J. P.; Miquel, P.; Poncelet, F. J.;
Dormant J. M.; Dutertre, F. Extraction cycles design for La Hague plants,
Proceedings of the International Conference on Future Nuclear Systems Emerg-
ing Fuel Cycles and Waste Disposal Options, Global ’93; Sept.12–17, Seattle
Washington, 1993.
38. Hugelmann, D.; Pradel, Ph. Revue Générale Nucléaire, 1995, 1, 30.
39. Musikas, C. Sep. Sci. Technol., 1998, 23, 1211.
40. Gasparini, G. M.; Grossi, G. Solv. Extr. Ion Exch., 1986, 4, 1233.
41. Küchler, L.; Schäfer, L.; Wojtech, B. Kerntechnik, 1970, 12, 327.
42. Schulz, W. W.; Navratil, J. D.; Recent Developments in Separations Science;
Li, N. N. Ed.; CRC Press: Boca Raton, Florida, 1980; Vol. 7, p. 31.
43. Schulz, W. W.; Horwitz, E. P. Sep. Sci. Technol., 1988, 23, 1191.
44. Nash, K. L.; Rickert, P. G.; Horwitz, E. P. Solv. Extr. Ion Exch., 1989, 7,
655.
45. Ozawa, M.; Nemoto, S.; Toghashi, A.; Kawata, T.; Onishi, A. Solv. Extr. Ion
Exch., 1992, 10, 829.
46. Prybylova, G. A.; Chmutova, M. K.; Nesterova, N. P.; Myaesoedov, B. F.;
Kabachnik, M. I. Radiokhimiya 1991, 33, 70.
47. Myaesoedov, B. F.; Chmutova, M. K.; Smirnov, I. V.; Shadrin, A. U. Global
’93: Future Nuclear Systems: Emerging Fuels Cycles and Waste Disposal Op-
tions, American Nuclear Society: La Grange Park, Illinois, 1993; p. 581.
48. Babain, V. V.; Shadrin, A. Yu. New Extraction Technologies for Management of
Radioactive Wastes in Chemical Separation Technologies and related Methods of
Nuclear Waste Management, Choppin, G. R.; Khankhasayev M. Kh. Eds.;
NATO Science Series, 2- Environmental Security; Kluwer Academic Publishers
1999; Vol. 53.
49. Horwitz, E. P.; Shulz, W. W. Metal Ion Separation and Preconcentration Pro-
gress and Opportunities Bond, A. H. Dietz, M. L. Rogers, R. D. Eds., American
Chemical Society, 1998.
50. Musikas, C.; Hubert, H. Solv. Extr. Ion Exch., 1987, 5, 877.
51. Nigond, L.; Musikas, C; Cuillerdier, C. Solv. Extr. Ion Exch., 1994, 12, 297.
52. Madic, C.; Hudson, M. J. High Level Liquid Waste Partitioning by Means of
Completely Incinerable Extractants; European Commission, Nuclear Science
and Technology, EUR18038 EN 1998; p. 208.
53. Madic, C.; Hudson, M. J.; Liljenzin, J. O.; Glatz, J. P.; Nannicini, R.;
Facchini, A.; Kolarik, Z.; Odoj, R. New Partitioning Techniques for Minor
Actinides; European Commission, Nuclear Science and Technology, EUR19149
EN 2000; p. 286.
54. Liljenzin, J. O.; Rydberg, J.; Skarnemark, G. Sep. Sci. Technol., 1980, 15,
799.
556 Musikas et al.
55. Cecille, L.; Dworschak, H.; Girardi, E.; Hunt, B. A.; Mannone, F.; Mousty, E.
Actinides Separations; Navratil, J. D. Schulz, W. W. Eds.; American Chemical
Society: Washington D.C., 1980; p. 427.
56. Schulz, W. W. The Chemistry of Americium; U.S. Department of Energy
Technical Information Center; Oak Ridge, Tennessee, 1976; p. 203.
57. Liljenzin, J. O.; Persson, G.; Svantesson, I.; Wingefors, S. The CTH-process for
HLLW treatment, Part I–General description and process design. Radiochim.
Acta. 1984, 35, 155.
58. Persson, G.; Wingefors, S.; Liljenzin, J. O.; Svantesson, I. The CTH-Process for
HLLW treatment, Part II-Hot test. Radiochim. Acta. 1984, 35, 163.
59. Persson, G.; Svantesson, I.; Wingefors, S.; Liljenzin, J. O. Hot test of a TAL-
SPEAK procedure for separation of actinides and lanthanides using
recirculating DTPA-Lactic acid solution. Solv. Extr. Ion Exch., 1984, 2(1), 89.
60. Horwitz, E. P.; Dietz, M. L.; Fisher, D. E. Solv. Extr. Ion Exch., 1990, 8, 557.
61. Rais, J.; Selucky, P.; Kyrs, M. J. Inorg. Nucl. Chem., 1976, 38, 1376, 1742.
62. Shuler, R. G.; Bowers, C. B. Jr.; Smith, J. E. Jr.; Van Brunt, V.; Davis, M. W.
Jr. Solv. Extr. Ion Exch., 1985, 3, 567.
63. Dozol, J. F.; Scwing-Weill, M. J.; Arnaud-Neu, F.; Böhmer, V.; Ungaro, R.;
van Veggel, F. C. J. M.; Wipff, G.; Costero, A.; Desreux, J. F.; de Mendoza, J.
Extraction and Selective Separation of Long-Lived Nuclides by Functionalized
Macrocycles; European Commission, Nuclear Science and Technology series,
EUR19605 EN 2000; p. 198.
64. King, L. J.; Bigelow, J. E.; Collins, E. D. Transplutonium Elements-Production
and Recovery; Navratil, J. D. Schulz, W. W. Eds.; ACS Symposium Series 161,
American Chemical Society: Washington, D.C. 1981; p. 133.
65. Berger, R.; Koehly, G.; Musikas, C.; Pottier, R.; Sontag, R. Nucl. Appl.
Technol., 1970, 8, 371.
66. Schulz, W. W. The Chemistry of Americium; U.S. Department of Energy
Technical Information Center: Oak Ridge, Tennessee, 1976, p. 30.
67. Silva, R.; Sikkeland, T.; Nurmia, M.; Ghiorso, A. Inorg. Nucl. Chem. Lett.,
1970, 6, 733.
68. Skarnemark, G.; Skälberg, M.; Alstadt, J.; and Björnstad, T. Nuclear Studies
with the Fast In-Line Chemical Separation System SISAK, Physica Scripta,
1986, 34, 597.
69. McDowell, W. J. Radioactivity Radiochem., 1992, 3(2), 26.
70. Raghavan, R. S. Phys. Rev. Lett. 1997, 78, 3618.
71. Choppin, G.; Liljenzin, J. O.; Rydberg, J. Radiochemistry and Nuclear Chem-
istry; 3rd Ed.; Butterworth-Heinemann, 2002; p. 593.
72. CURE: Clean Use of Reactor Energy, U.S. Department of Energy Report
WHC-EP-0268; Westinghouse Hanford Co.: Richland, Washington, 1990.
73. Schulz, W. W.; Bray, L. A. Sep. Sci. Technol., 1987, 22, 191.
74. Mukaiyama, T.; Kubota, M.; Takizuka, T.; Ogawa, T.; Mizumoto, M.; and
Yoshida, Y. Partitioning and Transmutation Program OMEGA at JAERI,
Proceedings of the GLOBAL’95 conference; Versailles, France, 1995, 1, 110.
Solvent Extraction in Nuclear Science 557
75. Barré, J. Y.; Bouchard, J. French R&D strategy for the back-end of the fuel
cycle. Proceedings of the Global ’93; Seattle, Washington, September 1993; 12–
17.
76. Musikas, C.; Vitart, X.; Pasquiou, J. Y.; Hoel, P. Chemical Separations; King
C. J.; Navratil, J. D. Eds.; Litarvan Literature: Denver, 1986; Vol. 2, p. 359.
77. Musikas, C.; Vitorge, P.; Pattee, D. Proceedings International Solvent Extrac-
tion Conference ISEC 83; American Institute of Chemical Engineers: Denver,
1983; p. 6.
78. Kolarik, Z.; Müllich, U.; Gassner, F. Solv. Ext. Ion Exch., 1999, 17, 23, 1155.
79. Musikas, C. Proceedings of the International Symposium on Actinides/Lantha-
nides Separations; Honolulu, Hawaii, 1984; and Actinide-Lanthanide Separa-
tion; World Scientific: Singapore, 1985 p. 19.
80. Zhu, Y. Radiochim. Acta. 1995, 68, 95–98.
81. Modolo, G.; Odoj, R. J. Radioanal. Nucl. Chem. 1998, 228(1,2), 83.
82. Miquel, P. Bull. Inf. Sci. Tech., CEA 1973, 57, 184.
13
Analytical Applications of Solvent Extraction
MANUEL AGUILAR, JOSÉ LUIS CORTINA, and ANA MARÍA
SASTRE Universitat Politècnica de Catalunya, Barcelona, Spain
13.1 INTRODUCTION
The role of solvent extraction in analytical chemistry has steadily increased
since the mid-1950s as a powerful separation technique applicable both to
trace and macro levels of materials. Work in this field has provided the
basis for a rich store of analytical methodology characterized by high sen-
sitivity and selectivity as is described in Chapters 2–4. Developments of new
extractants and their application to separation of a growing variety of
compounds are recognized as an important area of analytical chemistry.
Advances in this field over the last 50 years have been reported in a large
number of publications, among them several monographs and reviews [1–5].
Because of the great range of concentrations (from weightless trace levels of
carrier-free radioisotopes to macro levels of several weight percent of metal
ions) for which quantitative separations by solvent extraction are applicable,
this technique is equally useful in analytical, preparative, and process
chemistry. Solvent extraction has been also used as a separation step in
many analytical techniques and methods in response to the new problems
posed in many other fields such as medicine, biology, ecology, engineering,
etc. Among these, automatic methods of analysis have gained a notable
momentum and have motivated the development of a large number of
commercial instruments.
Solvent extraction has also played a major role in sample pre- or
posttreatment to improve selectivity and sensitivity. Initially simple schemes
of liquid–liquid extraction were used as separation methods for the clean-
up and preconcentration of samples, mainly because of its simplicity,
559
560 Aguilar et al.
1. It receives the two streams of immiscible phases and combines them into
a single flow with alternate and regular zones of the two phases (solvent
segmenter).
2. It facilitates the transfer of material through the interfaces of segmented
flow in the extraction coil, the length of which, together with the flow
rate, determines the duration of the actual liquid–liquid extraction.
3. It splits, in a continuous manner, the segmented flow from the extrac-
tion coil into two separated phases (phase separator).
The functions of these three parts are based on the same fundamental
principle, i.e., the selective wetting of internal component surfaces by both
organic and aqueous phases. In general, it is found that organic solvents wet
Teflon surfaces whereas aqueous phases wet glass surfaces.
S ¼ D1 =D2 (13.1)
where D12 denotes the distribution coefficient of the species for the mixture
of extractants. In analytical chemistry, for an extraction system based on a
pH-dependent extraction reaction, as in the case of acidic extractants, the
value of SC has been estimated by the following equation:
SC ¼ n pH50 (13.3)
where n is the charge of the metal ion and DpH50 is the difference of pH
corresponding to 50% extraction when the total concentration of the
extraction system is the same for the single system and for the mixtures.
where [AT] and [AoT ] are the concentrations of the microcomponent in the
concentrate and in the sample respectively. The need for preconcentra-
tion of trace compounds results from the fact that instrumental analytical
methods often do not have the required selectivity and/or sensitivity. Thus,
the combination of instrumental techniques with concentration techniques
significantly extends the range of application of the instrument. The chem-
ical techniques used in preconcentration can provide, in many cases, analyte
isolation, as well as high enrichment factors. The concept of sample pre-
concentration prior to determination could apply to many situations other
than just concentration enrichment and minimization of matrix effects.
Selection of a preconcentration scheme based on any liquid–liquid extrac-
tion step will depend upon on the type of analyte and/or sample matrix.
Concentration factors reasonably achieved by solvent extraction have
practical limits. However, considering a reasonable practical limit of ex-
tracting a 100 cm3 sample with 1–2 cm3 of organic phase, a concentration
factor of 100 to 50 can be achieved. One of the limitations of the pre-
concentration factor in batch solvent extraction is the difficulty of obtaining
good separation of the small volume of organic phase. Significant portions
of the organic solvent may adhere to the walls of the vessel, requiring
repeated washings, which decreases the concentration factor.
Analytical Applications of Solvent Extraction 565
color-forming reagents. The role of the most important organic reagents has
been discussed [17] and a comprehensive dictionary is available [18].
Chelating reagents are the most popular. Although most of these reac-
tions were initially developed for aqueous phase measurement, in many cases
taking into account the low aqueous solubility of the analyte-chromogenic
reagents, extraction into an organic phase was used for direct spectroscopic
measurements. Among the most important reagents used are the follow-
ing [19].
Dithizone (diphenylthiocarbazone) is a weak acid insoluble in water at
pH <7 but readily soluble in CCl4 and CHCl3. It reacts with many metal
ions to form chelates that can be extracted into the organic phase from
which the excess of the green reagent is stripped with dilute NH3. The most
stable dithizonates (Pt, Pd, Au, Ag, Hg, and Cu) are extracted from strongly
acid solutions. Other metals (Bi, Ga, In, Zn) are extracted from weakly acid
media and some (Co, Ni, Pb, Tl, Cd) from neutral or alkaline media. Some
compounds are extracted rapidly (Ag, Hg, Pb, Cd) and others more slowly
(Pd, Cu, Zn), while a few (Rh, Ir, and Ru) require prolonged heating to be
formed.
Azo dyes contain an azo link between two aromatic rings possessing a
orthohydroxy group. The most important reagents include PAN, PAR, and
Arsenazo III and generally they offer high sensitivity for the majority of
transition metals.
Chelating dyes include triphenylmethane reagents (e.g., Pyrocatechol
Violet, Eriochrome Cyanine R, Chrome Azurol S, Xylenol Orange) and
xanthene reagents (fluorones, e.g., Gallein, Pyrogallol Red, phenylfluorone,
and salicylfluorone). They form chelates with most metals. Ionic surfactants
make it easier to dissociate the protons of chelating triphenylmethane
reagents and facilitate reactions with easily hydrolyzable metals (Be, Al, Fe,
Se, Ti, Zr), leading to very sensitive but poorly selective methods.
Nonchelating dyes include basic triphenylmethane dyes (e.g., Brilliant
Green, Malachite Green, Crystal Violet), xanthene dyes (e.g., Rhodamine B,
Rhodamine 6G), azine dyes (e.g., Methylene Blue), and acid dyes (e.g.,
Eosin, Erythrosin). These are intensely colored and when paired with an
oppositely charged analyte ion lead to high sensitivities.
Crown ethers are not chromogenic unless they contain a pendant
chromogen able to dissociate a proton in a basic medium. The resulting
anion interacts strongly with the crown-complexed cation compensating the
electric charge. The formation of a zwitterion leads to a hydrophobic ex-
tractable species with a considerably shifted absorption maximum compared
with the protonated species. This allows the same spectrophotometric
determination to be used for a large number of metal ions, provided the
appropriate crown compound is used in each case. Another method involves
Analytical Applications of Solvent Extraction 569
assembled the channels face each other. The channels are about 0.10–
0.25 mm deep, 1.5 mm wide, and differ in length from 15 cm to 250 cm to
provide volumes from 12 mL to 1000 mL. In the smallest devices the channels
can be U-shaped grooves, while in other separators they are arranged in
spirals with the feed inlet on the periphery and the outlet in the center
(Fig. 13.3). The impregnated membrane, which is prepared by soaking the
Fig. 13.3 Schematic diagram of two different membrane units. (a) Membrane separator
unit composed of two machined blocks of PTFE or Ti (A), and PTFE membrane (B),
impregnated with stationary liquid. (b) Membrane unit. The PTFE membrane is placed
between the two blocks made of titanium. The two channels (donar and acceptor) that are
formed have a nominal volume of 12 mL.
578 Aguilar et al.
support in the organic phase, is clamped tightly between the planar surfaces
of the two blocks. Such devices can be coupled to chromatographic col-
umns, but are too large for on-line connection to packed capillary liquid
chromatography or capillary electrophoresis. To decrease the volume while
preserving the enrichment efficiency requires shallower channels with the
possibility of clogging problems. In such cases, porous hollow fiber modules
as supports for the liquid membrane can be used. The hollow fiber module
may have only one fiber of a microporous polymer about 15 mm long with
an internal diameter of 300–500 mm giving a lumen volume for the acceptor
phase of 1.9 mL and annular donor volume of approximately 0.5 mL.
13.5.2.1 Applications of Liquid Membranes in
the Analysis of Selected Samples
The most frequently used pretreatment methods for extraction and enrich-
ment of analytes are liquid–liquid solvent extraction and solid-phase
extraction (SPE). Liquid–liquid solvent extraction often gives a good cleanup
from the matrix. It allows, by incorporating different specific reagents,
improvement of the separation of the analyte from the sample. However,
liquid–liquid solvent extraction has some drawbacks. It is laborious and
difficult to automate and connect on-line to analytical instruments. In
addition, large amounts of organic solvents should be avoided for environ-
mental and health reasons. Supported liquid membranes (SLM) are an
attractive alternative based on the efficient cleanup of liquid–liquid extrac-
tion, using small amounts of organic solvents and avoiding the possibility of
emulsion formation. Using appropriate carriers, the SLM technique offers
very selective extraction of analytes in very complex samples. Furthermore,
preconcentration is often required in trace analysis to improve detection
limits and this can be also achieved using liquid membranes [79–81] with the
possibility of obtaining over 100 times enrichment of heavy metal ions as well
as various organic pollutants.
The most important features of liquid membranes are that they offer
highly selective extraction, efficient enrichment of analytes from the matrix
in only one step, and the possibility of automated interfacing to different
analytical instruments such as liquid chromatography, gas chromatography,
capillary zone electrophoresis, UV spectrophotometry, atomic absorption
spectrometry, and mass spectrometry [82].
13.5.2.1a Application in Cleanup Procedures
Many analyses of organic compounds in liquid samples require selective
cleanup and concentration. Direct on-line coupling of sample preparation to
the analytical instrumentation minimizes sample handling and thereby
the risk for contamination or loss of analyte. Also, on-line coupling makes
Analytical Applications of Solvent Extraction 579
the acceptor phase, it is possible to back-extract the metal ions into the
acceptor phase.
Metal ions such Cu2+, Cd2+, and Pb2+ can be preconcentrated
from water samples using liquid membranes containing 40% w/w of di-2-
ethylhexylphosphoric acid in kerosene diluent in a PTFE support. The
liquid membrane can be coupled on-line to an atomic absorption spectrom-
eter and has been shown to be stable for at least 200 h with extraction
efficiencies over 80%, and enrichment factors of 15 can be obtained. A
liquid membrane has also been used for sample cleanup and enrichment of
lead in urine samples prior to determination by atomic absorption spec-
trometry [100]. The experimental setup for metal enrichment is shown in
Fig. 13.4. Lead was enriched 200 times from urine [80] and several
metals were enriched 200 times from natural waters [88]. Using hollow fiber
of fatty acids) that demonstrate cloud point behavior with increasing solution
temperature, and zwitterionic surfactants (ammonioethylsulfates, ammo-
niopropylsulfates, ammoniopropanesulfonates, phosphobetaine, dimethyl-
alkylphosphine oxides) that show cloud point behavior on decreasing
solution temperature. Thus, cloud point temperature depends on the struc-
ture of the surfactant and its concentration and range from 0.5C to 120C
[105,106]. In a homologous series of polyoxyethylated surfactants, the
cloud point temperature increases with the hydrocarbon chain length and
increasing length of the oxyethylene chain. At a constant oxyethylene con-
tent in the surfactant molecule, the cloud point temperature is lowered by
decreasing the molecular mass of the surfactant and by branching of the
hydrophobic group.
The most important advantage of cloud point extraction is that only
small amounts of nonionic or zwitterionic surfactants are required and
consequently the procedure is less costly and more environmentally benign
than other conventional extraction techniques such as liquid–liquid extrac-
tion and solid–liquid extraction [107,108]. Moreover, CPE offers the pos-
sibility of combining extraction and preconcentration in one step.
13.5.3.1 Experimental Protocols
Operation procedures are performed by adding a small volume of a con-
centrated nonionic or zwitterionic surfactant solution to an aqueous sample
(50–100 cm3) containing the analyte, taking into account that the final sur-
factant concentration must be greater than its CMC value so that micelles are
present in solution. The analyte, depending on its affinity to the micelles, is
incorporated into the micellar aggregates in solution. The solution is heated
(in the case of nonionic surfactants) or cooled (with zwitterionic surfactants)
until the cloud point temperature is reached and the solution is allowed to
settle (settling temperature) in a thermostatted bath set at a temperature
above or below that of the cloud point (for nonionic and zwitterionic sur-
factants respectively) until the phases separate. As the density of both phases
in some cases is quite similar, centrifugation is often recommended to facil-
itate the physical separation. The analyte is concentrated in the surfactant-
rich phase in a small volume (50–400 mL) [105,106]. Figure 13.5 shows the
steps involved in CPE prior to analysis.
The surfactant selected for CPE technique should not have too high a
cloud point temperature. In practice, it is possible to obtain almost any
desired temperature by choosing an appropriate mixture of surfactants, as
cloud point temperatures of mixtures of surfactants are intermediate
between those of the two pure surfactants, or by the choice of an appro-
priate additive (i.e., salts, alcohols, organic compounds) [105].
584 Aguilar et al.
Fig. 13.5 Steps involved in cloud point extraction (CPE) prior to HPLC, GC, and CE
analysis.
Table 13.2 Summary of Cloud Point Extractions of Metals Chelates Using Nonionic
Surfactants Micelles
Metal ion Ligand Nonionic surfactant Experimental conditions
pH; CFa, %Eb
Ni(II) TANc Triton X-100 pH 7.0 (phosphate); CF = 30
PAN Triton X-100 pH 5–6
Zn(II) PAN PONPE-7.5 pH 10 (carbonate); CF = 40
QADI PONPE-7.5 pH 9; NH3; CF = 40
Ni(II), Zn(II), Cd(II) PAMP PONPE-7.5
Au(III) HCl PONPE-7.5 HCl; %E > 95
Ni(II), Cd(II), Cu(II) PAMP PONPE-7.5 pH 5.6
Ni(II) PAN OPd pH 6.0; CF = 15–25
Transition metal ions TAN PONPE-7.5
Fe(III), Ni(II) TAC Triton X-100
U(VI) PAN Triton X-114 pH 9.2; %E = 98
U(VI), Zr(IV) Arsenazo Tween 40e pH 3; %E > 96
Cu(II), Zn(II), Fe(III) Thiocyanate PONPE-7.5 %E = 72.5–96.8
Er(III) CMAP PONPE-7.5 CF = 20
Gd(III) CMAP PONPE-7.5 CF = 3.3; %E = 99.88
Cd(II) PAN Triton X-114 CF = 60
Ni(II), Zn(II) PAN Triton X-114
Ru(III) Thiocyanate Triton X-110 CF = 5–10
Au(III) HCl PONPE-10f %E > 90
Ga(III) HCl PONPE-7.5f %E & 90
Ag(I), Au(III) DDTP Triton X-114f CF = 9–130
Cu(II) LIX54 Igepal CO-630f
a
Concentration factor.
b
Percent extracted.
c
Abbreviations for ligands: CMAP: 2-(3,5-dichloro-2-pyridylazo)-5-dimethylaminophenol; DDTP:
o,o-diethyldithiophosphoric acid; LIX54: dodecylbenzoylacetone; PAN: 1-(2-pyridylazo)-2-naphthol;
QADI: 2-(8-quinolazo)-4,5-dipheylbenzimidazole; PAMP: 2-(2-pyridylazo)-5-methylphenol; PAP: 2-
(2-pyridylazo)-phenol; TAC: 2-(2-thiazoylazo)4-methylphenol; TAN: 1-(2- thiazoylazo)-2-naphthol.
d
OP surfactants refer to (polyethyleneglycol octylphenyl ethers).
e
In this extraction, the concentrated surfactant-rich phase was a solid rather than a liquid.
f
Abbreviations for surfactants: Igepal CO-630: nonylphenoxypoly(ethylenoxy)ethanol; PONPE-7.5:
polyoxyethylene(7.5)nonylphenyl ether; PONPE-10: polyoxyethylene(10)nonylphenyl ether; TRI-
TON X: t-octylphenoxypolyoxyethylene ether.
Fig. 13.6 Schematic diagrams of two liquid membrane electrodes: (a) a commercial
Onion electrode; (b) an improved version developed by Szczepaniac and Oleksy [125].
that, in the latter technique, species are separated, not by discrete extraction
steps, but by continuous equilibration between two phases, one that is
stationary and the other mobile. The same fundamental laws of phase
equilibrium apply to both, extraction and chromatography. In principle, the
latter is a multiple-extraction process in which the mobile phase moves
continuously over the fixed phase in a chromatographic column.
solvent, whereas the mobile liquid is the polar solvent, most frequently
water or an aqueous mixture with polar organic solvents [139–143].
For conventional or ‘‘normal,’’ in contrast with reversed-phase,
LLPC, many materials have been used as the solid support for the sta-
tionary liquid. In addition to silica gel, which was the first and is still the
most popular material, a variety of other adsorbents that adsorb the polar
solvent such as cellulose powder, starch, alumina, and silicic acid have been
used. The more recent practice of HPLC has greatly simplified the technique
in providing column stability for repeated use and for treatment of large
volumes.
Reversed-phase chromatography is the predominant technique in
HPLC, and chemically bonded silica gel supports are made specifically for
the nonpolar stationary phase. In the last decade, as many as 60% of the
published LLPC techniques refer to RPC. The reasons for this involve
the significantly lower cost of the mobile liquid phase and a favorable elu-
tion order that is easily predictable based on the hydrophobicity of the
eluate.
The selection of solvents for LLPC is similar to the selection of sol-
vents in liquid–liquid extraction systems. The solid support has little effect
upon the selection of the solvent pair, except for the obvious fact that a
hydrophilic support for a polar stationary phase requires a hydrophobic
mobile phase with the opposite for a reversed-phase system. Both solvents
should exhibit good solubility for the solute(s); otherwise the column
loading capacity would be too low. Frequently, the separation potential of
LLPC columns can be additionally enhanced by using solvent mixtures
rather than a single solvent as the mobile phase. In RPC, typical mobile
phases are water, aqueous electrolyte solutions, or aqueous mixtures of
one or more water-immiscible organic solvents. Water, the most polar of
mobile phases employed, is the weakest eluent. For specific purposes, such
as the separation of closely related acidic or basic substances, the mobile
phase is adjusted by buffers.
A large variety of analytical separation processes has been reported in
the literature. Table 13.4 [144–146] demonstrates the range of organic com-
pounds for which LLPC has been applied. The method has been limited
essentially to organic compounds, with much less use in the field of separa-
tion of metal ions or complexes. However, chromatographic separation pro-
cedures have been successfully used to separate metals with a combination of
a cation exchanger as a stationary phase and a solution of a chelating reagent
as a selective mobile phase. Thus chromatographic separations of many of
the rare-earth elements [Gd(III), Y(III), La(III), Pr(III), Nd(III), Ho(III),
Er(III), Sc(III)] from acidic solutions have been achieved using different
acidic organophosphorous extractants (e.g., DEHPA, PC88A) as well as
Analytical Applications of Solvent Extraction 593
transfer and collection of extracted analytes from the SFE to the chromato-
graphic system), and the final chromatographic conditions be understood
and controlled. With many on-line approaches the sample extract is com-
mitted to a single analysis, and the extract is not available for analysis by
other methods. While off-line SFE has the advantages of simplicity and
availability of commercial instrumentation, and provides extracts that can
be analyzed by several instrumental methods, on-line techniques have
greater potential for enhanced sensitivity and automation.
Off-line SFE is conceptually a simple experiment to perform and
requires only relatively basic instrumentation. The instrumental components
necessary include a source of fluid, most often CO2 or CO2 with an organic
modifier, a means of pressurizing the fluid, an extraction cell, a method of
controlling the extraction cell temperature, a device to depressurize the
supercritical fluid (flow restrictor), and a device for collecting the extracted
analytes.
13.6.2.2 On-Line Coupling SFE with
Analytical Techniques
On-line supercritical fluid extraction/GC methods combine the ability of
liquid solvent extraction to extract efficiently a broad range of analytes with
the ability of gas-phase extraction methods to rapidly and efficiently transfer
the extracted analytes to the gas chromatograph. The characteristics of
supercritical fluids make them ideal for the development of on-line sample
extraction/gas chromatographic (SFE-GC) techniques. SFE has the ability
to extract many analytes from a variety of matrices with recoveries that rival
liquid solvent extraction, but with much shorter extraction times. Addi-
tionally, since most supercritical fluids are converted to the gas phase upon
depressurization to ambient conditions, SFE has the potential to introduce
extracted analytes to the GC in the gas phase. As shown in Fig. 13.8, the
required instrumentation to perform direct coupling SFE-GC includes suit-
able transfer lines and a conventional gas chromatograph [162,163].
Finally, supercritical fluid chromatography, in which a supercritical
fluid is used as the mobile phase, was introduced by Klesper [164–166].
SFE directly coupled to SFC provides an extremely powerful analytical tool.
The efficient, fast and selective extraction capabilities of supercritical fluids
allows quantitative extraction and direct transfer of the selected solutes
of interest to be accomplished to the column, often without the need for
further sample treatment or cleanup. Extraction selectivity is usually
achieved by adjusting the pressure of the supercritical fluid at constant
temperature or, less often, by changing the temperature of the supercritical
fluid at constant pressure. SFE coupled with packed column SFC has found
596 Aguilar et al.
Fig. 13.8 Schematic diagram of a simple SFE-GC system showing all the required
components. (Several manufacturers supply suitable components and specific suppliers are
listed only for the reader’s convenience.) Components are: (A) SFE grade extraction fluid
source; (B) 1.5 mm o.d. stainless steel tubing (0.77 mm or smaller i.d.); (C) shut-off valves
(SSI model 02–120 or equivalent, Supelco, Bellefonte, PA); (D) SFE pump; (E) SFE cell
heater; (F) approx. 0.5 m long coil of 1.5 mm stainless steel tubing for fluid preheater; (G)
1.5 mm1.5 mm tubing union (e.g., Parker or Swagelok brand); (H) finger-tight con-
nectors (e.g. Slip-Free connectors from Keystone Scientific Bellefonte, PA, USA); (I) SFE
cell; (J) restrictor connector ferrule (Supelco M2-A, Bellefonte, PA) which is used to
replace the stainless steel ferrule in the outlet end of the tubing union ‘G’; (K) 15–30 mm
i.d. fused silica tubing restrictor (Polymicro Technologies, Phoenix, AZ); (L) GC injection
port.
Analytical Applications of Solvent Extraction 597
Fig. 13.9 Schematic diagram of a single-channel flow injection manifold, showing the
transient nature of the signal output. (--) Liquid flow; (.....) data flow.
could take place in different locations of the schemes defined previously, for
example:
1. In a coil after the sample is merged with a reagent stream.
2. In the solvent segmenter when it is necessary to avoid precipitation of
the extractable compound in the aqueous phase before it reaches the
continuous extractor (SS).
3. In the extraction coil (EC), when the extraction reaction is based on the
formation of metal chelates and the ligand is dissolved in the organic
solvent. Here the reaction and the transfer of the chelate take place
simultaneously.
4. In a coil located after the extractor using a reagent stream that is merged
prior to the detector with the phase containing the analyte.
The derivatizing reactions used in these systems are not substantially
different from those described in classical analytical determinations. The
most common are ion-pair formation reactions, where the analyte forms an
ion pair directly, e.g., determination of anionic surfactants by using ethyl
violet or methylene blue, and the determination of codeine with picric acid
and enalapril with bromothymol blue; or after reacting with a reagent to
form a bulkier, charged species e.g., tetracyanatocobaltate or phosphomo-
lybdate. Metal cations are typically transformed into extractable metal che-
lates by using such common ligands as described in section 13.4.1 (e.g., oxine,
dithizone, dimethylglyoxime, etc.). Some of the most important applications
of the technique are in the determination of surfactants in waters, quinizarin
in hydrocarbons, amines in pharmaceuticals, and drugs in urine [177].
Typical FIA manifolds are shown in Fig. 13.10 with two general
alternatives depending on whether injection takes place before or after the
continuous extractor device. The most common situation is when prior
injection of the sample take place. Figure 13.10(a) depicts a manifold for the
determination of vitamin B1 in pharmaceuticals [178], based on the oxida-
tion of thiamine to thiochrome in a carrier of potassium ferricyanide in a
basic medium (NaOH). The thiochrome is continuously extracted into a
chloroform stream and the fluorescence of the organic phase is measured
continuously.
Continuous multiextraction could be applied for the separation and
determination of carcinogenic polynuclear aromatic compounds (Fig. 13.10b)
in crude-oil-ash residues [179]. The ash sample dissolved in a cyclohexane
stream merges with a DMSO stream in the first solvent segmenter. After
passing through a glass extraction coil, the DMSO phase is mixed with a
water stream of cyclohexane segmented with the aqueous DMSO phase.
After passing through a Teflon extraction coil, the organic phase is carried
through the flow cell of a fluorimeter.
Fig. 13.10 Typical FIA extraction manifolds. Injection in front of the extractor com-
ponents: (a) simple; (b) multiextraction; and (c) injection after the extraction has been
carried out. (Based on systems from Refs. 178, 179.)
600
Analytical Applications of Solvent Extraction 601
ACRONYM LIST
AAS atomic absorption spectrophotometry
ASE accelerated solvent extraction
CCFA completely continuous flow analysis
CMC critical micelle concentration
CPE cloud point extraction
CZE capillary zone electrophoresis
FAAS furnace atomic absorption spectrophotometry
FIA flow injection analysis
GC gas chromatography
GC-MS mass chromatography
GF-AAS graphite furnace atomic absorption spectrophotometry
HPLC high-performance liquid chromatography
ICP inductively coupled plasma spectrophotometry
ICP-AES inductively coupled plasma with atomic emission spectrophoto-
metry
ISE ion-selective electrode
ISFET ion-selective field effect transistors
LLPC liquid–liquid partition chromatography
ME micelle-mediated extraction
MIMS membrane introduction mass spectrometry
PAH polynuclear aromatic hydrocarbon
PCDF polychlorinated dibenzofuran
PCB polychlorinated biphenyl
PF preconcentration factor
PONPE polyoxyethylene nonyl phenyl ether
PTFE polytetrafluoroethene
PVC porous polyvinyl chloride
REE rare earth elements
602 Aguilar et al.
REFERENCES
1. Morrison, M. G.; Freiser, H. Solvent Extraction in Analytical Chemistry; John
Wiley and Sons; New York, 1957.
2. Stary, J. The Solvent Extraction of Metal Chelates; Pergamon Press: Oxford,
1964.
3. De, A. K.; Khopkar, S. M.; Chalmers, R. A. Solvent Extraction of Metals;
Van Nostrand Reinhold: London, 1970.
4. Alders, L. Liquid-Liquid Extraction; Elsevier: New York, 1995.
5. Zolotov, Y. A.; Bodnya, V. A.; Zagruzina, A. N. CRC Crit. Rev. Anal.
Chem., 1982, 14, 2.
6. Keller, R.; Mermet, J. M.; Otto, M.; Widmer, H. M. Eds. Analytical Chem-
istry; Wiley-VCH, Verlag GmbH: Weinheim, Germany, 1998; p. 25.
7. Valcarcel, M. Liquid-Liquid Extraction in Continuous Flow Analysis. In
Developments in Solvent Extraction; S. Alegret, Ed., Ellis Horwood Series in
Analytical Chemistry: West Sussex, UK, 1988; p. 135.
Analytical Applications of Solvent Extraction 603
36. Jauniaux, M.; De Meyer, M.; Lejeune, W.; Levert, J. M. Bull. Soc. Chim.
Belg., 1975, 84, 565.
37. Luke, C. L. Anal. Chem., 1956, 28, 1443.
38. Marczenko, Z.; Kalowska, H. Chem. Anal. (Pol.), 1976, 21, 183.
39. Shtefak, M.; Kirsh, M.; Rais, I. Zh. Anal. Khim., 1976, 32, 1364.
40. Takeda, Y.; Suzuki, S.; Ohyagi, Y.; Chem. Lett., 1978, 12, 1377.
41. Roberts, A. H. C.; Turner, M. A.; Syers, T. K. Analyst, 1976, 101, 574.
42. Karisson, R.; Gorton, L. Talanta, 1976, 23, 672.
43. Carry, J.; Menon, M. P. J. Radioanal. Chem., 1975, 34, 319.
44. Gentry, C. H. R.; Sherrington, L. G. Analyst, 1946, 71, 432.
45. Argollo, R. M.; Schilling, J. G. Anal. Chem. Acta., 1978, 96, 117.
46. Fujinaga, T.; Puri, B. K. Talanta, 1975, 22, 71.
47. Bode, H.; Z. Anal. Chem., 1955, 144, 165.
48. Kaiyanarman, S.; Khopkar, S. M. Anal. Chem., 1977, 49, 1192.
49. Charib, A.; Morris, D. F. C. Talanta, 1978, 25, 569.
50. Tomazic, B.; Branica, M. J. Radioanal. Chem., 1976, 30, 361.
51. Pollock, E. N. Anal. Chim. Acta., 1977, 88, 399.
52. Fantova, J.; Kliz, Z.; Cermakova, Z.; Suk, V. Chem. Listy, 1980, 74, 291.
53. Shimomura, S.; Sakurai, H.; Morita, I.; Mino, Y. Anal. Chem. Acta., 1978,
96, 69.
54. Steinnes, E. Radiochem. Radioanal. Lett., 1978, 33, 205.
55. Shamaev, Y. I.; Bogdanov, N. V. Zh. Anal. Khim., 1976, 31, 72.
56. Hasemann, K.; Bock, R. Z. Anal. Chem., 1985, 274, 185.
57. Huffman, E. H.; Beaufait, L. J. J. Amer. Chem. Soc., 1949, 71, 3179.
58. Wadelin, C.; Mellon, M. G. Anal. Chem., 1953, 25: 1668.
59. Fetchett, A. W.; Daughtrey, D.; Hunter, E. Anal. Chim. Acta., 1975, 79, 93.
60. Kamada, T.; Yamamoto, Y. Talanta, 1977, 24, 330.
61. Nakamura, K.; Ozawa, T. Anal. Chim. Acta., 1976, 86, 147.
62. Svehla, G.; Tolg, G. Talanta, 1976, 23, 755.
63. Lapenko, L. A.; Gibalo, I. M. Zh. Anal. Khim., 1976, 31, 481.
64. Spivakov, B. Y.; Orlova, Y. A.; Malyutina,T. M.; Kirilova, T. L. Zh. Anal.
Khim., 1979, 34, 161.
65. Kamada, T. Shiroishi, T.; Yamamoto, Y. Talanta, 1978, 25, 15.
66. Chao, T. T.; Sanzolone, R. F.; Hubert, A. E. Anal. Chim. Acta., 1978, 96, 251.
67. Kudo, K.; Shigernatsu, T.; Kobayashi, K. J. Radioanal. Chem., 1977, 36, 65.
68. Bhowal, S. K.; Umiand, E. Anal Chem., 1976, 282, 197.
69. Musil, J.; Dolezal, J. Anal. Chim. Acta, 1977, 92, 301.
70. Roberts, R. F. Anal. Chem., 1977, 49, 1861.
71. Tribalat, S.; Beydon, J. Anal. Chim. Acta., 1953, 8, 22.
72. Fujinaga, T.; Lee, H. L. Talanta, 1977, 24, 395.
73. Fujinaga, T.; Puri, B. K. Indian J. Chem., 1976, 14, 72.
74. Zolotov, Y. A.; Petrukhin, O. M.; Shevehenko, U. N.; Dunina, U. Y.; Ru-
khadze, E. G. Anal. Chim. Acta., 1978, 100, 613.
75. Matsuoka, S.; Yoshimura, K.; Waki, H. Mem. Fac. Sci. Kyushu Univ., Ser.
C., 1991, 18, 55–62.
Analytical Applications of Solvent Extraction 605
76. Compaño, R.; Ferrer, R.; Guiteras, J.; Prat, M. D. Analyst, 1994, 119, 1225.
77. Kirkbright, G. F.; Narayanaswamy, R.; Welti, N. A. Analyst, 1984, 109, 15.
78. Narayanaswamy, R.; Sevilla, F. Anal. Chim. Acta., 1986, 189, 365.
79. Jönsson, J. Å.; Mathiasson, L. Trac-Trend. Anal. Chem., 1999, 18, 318.
80. Jönsson, J. Å.; Mathiasson, L. Trac-Trend. Anal. Chem., 1999, 18, 325.
81. Jönsson, J. Å.; Mathiasson, L. Trac-Trend. Anal. Chem., 1992, 11, 106.
82. Audunsson, G. Anal. Chem., 1986, 58, 2714.
83. Norberg, J.; Zander, Å.; Jönsson, J. Å. Chromatographia, 1997, 46, 483.
84. Nilvé, G.; Stebbins, R. Chromatographia, 1991, 32, 269.
85. Nilvé, G.; Audunsson, G.; Jönsson, J. J. Chromatogr., 1989, 471, 151.
86. Audunsson, G. Anal. Chem., 1988, 60, 1340.
87. Lindegård, B.; Jönsson, J. Å.; Mathiasson. L. J. Chromatogr., 1992, 573, 191.
88. Thordarson, E.; Pálmarsdóttir, S.; Mathiasson, L.; Jönsson, J. Å. Anal.
Chem., 1996, 68, 2559.
89. Norberg, J.; Emnéus, J.; Jönsson, J. Å.; Mathiasson, L.; Burestedt, E.;
Knutsson, M.; Marko-Varga, G. J. Chromatogr. B, 1997, 701, 39.
90. Pálmarsdóttir, S.; Lindegård, B.; Deiningr, P.; Edholm, L. E.; Mathiasson, L.;
Jönsson, J. A. J. Cap. Electrophor., 1995, 2, 185.
91. Pálmarsdóttir, S.; Mathiasson, L.; Jönsson, J. Å.; Edholm, L. E. J. Chro-
matogr. B, 1997, 688, 127.
92. Pálmarsdóttir, S.; Thordarson, E.; Edholm, L. E.; Jönson, J. Å.; Mathiasson,
L. Anal. Chem., 1997, 69, 1732.
93. Bauer, S.; Solyom, D. Anal. Chem., 1994, 66, 4422.
94. Cisper, M. E.; Hembeergeer, P. H. Rapid Commun. Mass Sp., 1997, 11,
1454.
95. Johnson, R. C.; Koch, K.; Cooks, R. G. Ind. Eng. Chem. Res., 1999, 38, 343.
96. Johnson, R. C.; Koch, K.; Cooks, R. G. Anal. Chim. Acta., 1999, 395, 239.
97. Shen, Y.; Obuseng, V.; Grönberg, L.; Jönsson, J. Å. J. Chromatogr. A, 1996,
725, 189.
98. Grönberg, L.; Shen, Y.; Jönsson, J. Å. J. Chromatogr. A, 1993, 655, 207.
99. Djane, N. K.; Bergdahl, I. A.; Ndung’u, K.; Schütz, A.; Johansson, G.; Ma-
thiasson, L. Analyst, 1997, 122, 1073.
100. Djane, K. N.; Ndung’u, F.; Malcus, G.; Johansson, G.; Mathiasson, L. Fre-
senius J. Anal. Chem., 1997, 358, 822.
101. Cox, J. A.; Bhatnagar, A. Talanta, 1990, 37, 1037.
102. Parthasarathy, N.; Pelletier, M.; Buffle, J. Anal. Chim. Acta., 1997, 350, 183.
103. Djane, N. K.; Ndung’u, K.; Johnson, K.; Sartz, H.; Törnström, T.; Ma-
thiasson, L. Talanta, 1999, 48, 1121.
104. Scamerhorn, J. F.; Harwell, J. H. Surfactant-Based Separations Processes;
Marcel Dekker: New York, 1989; p. 139.
105. Hinze, W. L.; Pramauro, E. Criti. Rev. Anal. Chem., 1993, 24, 133.
106. Carabias-Martı́nez, R.; Rodrı́guez-Gonzalo, E.; Moreno-Cordero, B.; Pérez-
Pavón, J. L.; Garcia-Pinto, C.; Fernández Laespada, E. J. Chromatogr. A,
2000, 902, 195.
107. Quina, F. H.; Hinze, W. L. Ind. Eng. Chem. Res., 1999, 38, 4150.
606 Aguilar et al.
108. Huddleston, J. G.; Willauer, H. D.; Griffin, S. T.; Rogers, R. D. Ind. Eng.
Chem. Res., 1999, 38, 2523.
109. Watanabe, H.; Tanaka, H. Talanta, 1978, 25, 585.
110. Garcı́a Pinto, C.; Pérez Pavón, J. L.; Moreno Cordero, B.; Romero Beato, E.;
Garcı́a Sánchez, S. J. Anal. Atomic Spectrometry, 1996, 11, 37.
111. Akita, S.; Rovira, M.; Sastre, A. M.; Takeuchi, H. Sep. Sci. Technol., 1998,
33, 2159.
112. Okada, T. Anal. Chem., 1992, 64, 2138.
113. Horvath, W. J.; Huie, C. W. Talanta, 1993, 9, 1385.
114. Garcı́a Pinto, C.; Pérez Pavón, J. L.; Moreno Cordero, B. Anal. Chem., 1995,
67, 2606.
115. Stangl, G.; Niessner, R. Mikrochim. Acta., 1994, 113, 1.
116. Ferrer, R.; Beltran, J. L.; Guiteras, J. Anal. Chim. Acta., 1996, 330, 199.
117. Eiguren, A.; Sosa, Z.; Santana, J. J. Analyst 1999, 124, 487.
118. Calvo Seronero, L.; Fernández Laespada, M. E.; Pérez Pavón, J. L.; Moreno
Cordero, B. J. Chromatogr. A, 2000, 897, 171.
119. Eiguren, A.; Sosa, Z.; Santana, J. J. Anal. Chim. Acta., 1998, 358, 145.
120. Moreno, B.; Pérez, J. L.; Garcı́a, C.; Fernández, M. E. Talanta, 1993, 40,
1703.
121. Carabias-Martı́nez, R.; Rodrı́guez-Gonzalo, E.; Domı́nguez-Alvarez, J.;
Hernández-Méndez, J. Anal. Chem., 1999, 771, 2468.
122. Eisenman, G. Ion Selective Electrodes; National Bureau of Standards (USA)
Special Publications: Washington, D.C., 1969; p. 314.
123. Janata, J.; Josowitz, M.; de Vaney, D. M. Anal. Chem., 1994, 66, 207R.
124. Camman, K.; Lemke, U.; Rohen, A.; Sander, J.; Wilken, H.; Winter, B.
Angew. Chem. 1991, 109, 519.
125. Szczepaniac, W.; Oleksy, J. Anal. Chim. Acta., 1986, 189, 237.
126. Worf, W. E. The Principles of Ion Selective Electrodes and of Membrane
Transport; Akademiai Kiadó: Budapest, 1981; Elsevier: Amsterdam, 1981.
127. Koryta, J. Ion-Selective Electrodes; Cambridge University Press: London,
1975.
128. Freiser, H. Ion-Selective Electrodes in Analytical Chemistry; Plenum Press:
New York, Vol. 1, 1978.
129. Pretsch, E.; Buchi, R.; Amman, D.; Simon, W. Essays on Analytical Chem-
istry; Pergamon Press: New York, 1977.; p. 321.
130. Svehia, G. Comprehensive Analytical Chemistry; Elsevier: New York, Vol. 11,
Chap. 3, 1981; p. 346.
131. Hobby, P. C.; Moody, G. J.; Thomas, J. D. R. Analyst, 1983, 108, 581.
132. Geissler, M.; Kunze, R. Anal. Chim. Acta., 1986, 189, 245.
133. Simon, W.; Pretsch, E.; Morf, W. E.; Amman, D.; Oesch, U.; Dinten, O.
Analyst, 1984, 109, 207.
134. Simon, W.; Morf, W. E.; Meier, P. C. Struct. Bonding, 1973, 16, 113.
135. Eisenman, G. Anal. Chem., 1968, 40, 311.
136. Xie, R. Y.; Christian, G. D. Analyst, 1987, 112, 61.
137. Pioda, L. A. R.; Stankova, Y.; Simon, W. Anal. Lett., 1969, 2, 665.
Analytical Applications of Solvent Extraction 607
14.1 INTRODUCTION
Recently the protection of the environment has become increasingly
important for industry with the requirement that the potential impact on the
environment is considered for all aspects of industrial processes. Such
considerations are supported by environmental legislation that controls all
types of emissions as well as the treatment of wastes. Such legislation is
based on global standards that have largely resulted from developments
within the European Union, Japan, and the United States in collaboration
with international conventions. Of these, the Basel Convention (1989) and
the Earth Summit in Rio de Janeiro (1992) were significant in the control
and prevention of wastes. In the case of liquid wastes that are most
appropriate for treatment by liquid–liquid extraction, limits for discharge
into the aqueous environment have been established by the three countries
already mentioned. These limits depend on the particular country and
sometimes on the industry. (See section 14.6.)
Environmental limits vary according to country, and in the future
pressure will be applied to those countries where limits are considered too
high. Indeed, there is a general trend for discharge limits to be reduced, with
the concept of zero industrial discharge for certain metals being the ultimate
aim for the future.
Metal pollution can arise from a number of different sources, for
example:
*Retired.
609
610 Cox and Reinhardt
remove entrained organic compounds and residual toxic metals. Thus the
main applications of solvent extraction in treatment of wastes are in the
removal and recovery of values from waste streams to provide a process
water for recycling. This reduces the effect on the environment by mini-
mizing the amount of effluent for treatment and reduction in fresh water
consumption.
The most important metalliferous liquid effluents where solvent
extraction could be applied are from the various metal finishing operations:
plating, pickling, etching, and the wash waters arising from the cleaning of
work pieces. In the case of solids, in addition to scrap metal and alloy wastes
from manufacturing operations, a number of other products use valuable
and toxic metals and offer potential applications, e.g., spent automobile
catalysts, Ni/Cd batteries, etc.
In the treatment of both types of waste, the objectives must be:
To remove the metals effectively so as to produce a liquid waste stream
capable of reuse or finally meeting environmental discharge limits
To recover the metals for recycling within the plant, or at appropriate
purity for sale
To separate other impurities in a form for resale as by-products or safe
environmental discharge
These principles are embodied in the concepts of zero discharge and
sustainable technology.
containing 40 g dm3 Zn, emanating from the leaching of steel furnace dust.
The resulting raffinate, with a residual Zn concentration of 15–20 g dm3,
was recycled to leaching. Comparable net transfer conditions were obtained
[5] in the modified Zincex process (MZP) for the processing of spent
domestic batteries.
Similar high transfer of copper in a solvent extraction procedure has
been recently reported [6]. After pressure leaching to liberate copper
(60 g dm3) under conditions that minimize the free acid (pH 1.4) and iron
in solution, copper was extracted to 10 g dm3 in three stages with loading in
the organic phase of 12.5 g dm3.
The situation described is valid in sulfuric acid solutions, when no
strong metal complexes are formed. However with hydrochloric acid the
formation of metal chloride complexes makes the situation quite different.
The influence of the chloride ion concentration on the extraction of some
metal chloride complexes of general interest is shown in Fig. 11.2. For
example, the separation of zinc and iron(III) can be achieved at a low
chloride concentration, and later the separation of cobalt and nickel is
described. The disadvantage of using the chloride system is concerned with
stripping. For example, when a zinc chloride complex ZnCl42 is extracted
with a tertiary amine, the stripped species corresponds to ZnCl2, resulting in
an increased chloride concentration and an interruption of the mass trans-
fer. The consequence of this is that it is difficult to achieve concentrated zinc
chloride strip liquors. Two examples are given next.
14.4.1.1 Extraction of Zinc from Weak Acid Effluents
The first example describes the extraction of zinc from weak acid solutions.
In the manufacture of rayon, rinse waters and other zinc-containing liquid
effluents are produced. The total liquid effluent in a rayon plant may amount
to several m3 per minute with a zinc concentration of 0.1–1 g dm3 and pH
normally 1.5–2. In addition to zinc, the effluent contains surface-active
agents and dirt (organic fibers and inorganic sulfide solids). The use of both
precipitation (OH and S2) and ion exchange has been reported to remove
zinc from such effluents. In addition, solvent extraction has successfully
been used to recycle the zinc back to the operation.
In the process [3,7], zinc is extracted to greater than 95% (pH > 2)
with DEHPA (25%) in kerosene in two or three stages. Stripping is per-
formed with a sulfuric acid solution (Figs. 14.1 and 14.2). By adjusting the
net flow rate of the solution, the concentration of zinc in the strip solution
may be increased to 50 g dm3 or more. Thus, the resulting zinc sulfate
solution can be reused directly in the spinning bath.
The extraction of zinc is pH dependent. The lower the pH the lower
the extraction of zinc. In addition, the extraction of each zinc ion (Zn2+)
Solvent Extraction in the Recovery of Waste 615
liberates two hydrogen ions (H+). Thus, the more zinc in solution, the less is
extracted. This disadvantage means that using solvent extraction without
some neutralization (pH > 2) results in a residual zinc concentration in the
raffinate, which is too high to meet with environmental standards of today.
14.4.1.2 Extraction of Nickel from Cadmium Electrolytes
In the electrowinning of cadmium, nickel is an interfering element that has
to be continuously removed from the electrolyte, a weak acidic cadmium
sulfate solution. To handle the undesirable buildup of nickel contamination
and at the same time obey environmental demands, the industry installed a
solvent extraction ‘‘kidney’’ [8].
The nickel extraction can be performed with DEHPA under defined
conditions. As described earlier for zinc, the extraction of nickel is pH
dependent. The D-value for nickel decreases drastically with decreasing pH
below 3.5. The extraction of each nickel ion (Ni2+) liberates two hydrogen
ions (H+) from the extractant H+DEHPA, which means that only a very
small amount of nickel can be transferred to the solvent before the extrac-
tion stops. However, by adding a neutralization reagent (NaOH) into the
616 Cox and Reinhardt
Fig. 14.2 Zinc extraction plant at Enka AKZO, Netherlands. Full-scale plants have been
built and operated at Svenska Rayon AB in Sweden and at Enka AKZO in The Netherlands.
metal sulfate solution. The operation can be regarded as two separate inter-
connected processes, one for copper recovery and one for zinc (Fig. 14.4).
In the copper circuit, copper-rich material is leached in a pH-controlled
leach at 608C to a final pH of about 2. Oxidizing conditions are maintained,
e.g., by addition of MnO2, to promote copper dissolution and to prevent
cementation of metallic copper when the feed material contains metallic zinc
(brass or iron). The leach solution, containing about 20 g dm3 zinc and
4 g dm3 copper, is filtered and fed to a solvent extraction circuit. Copper is
extracted in four stages with a kerosene solution containing H+R (10%).
Stripping is performed in three stages with a sulfuric acid electrolyte.
Copper metal is recovered from the electrolyte by electrowinning.
The zinc circuit consists of a similar pH-controlled leach at 608C under
oxidizing conditions. Zinc flue dust with low copper content is leached with
the copper barren raffinate and with part of the zinc raffinate. The zinc
leaching operation is maintained at about pH 2 for most of the leaching time
and then slowly raised to a final pH of 4.5, reducing the iron level to below
10 ppm. The leach solution is filtered and cleaned from impurity metals such
as Cu, Ni, and Cd by an ordinary cementation procedure, again filtered and
finally fed to a solvent extraction circuit. Zinc is extracted in three stages with
Fig. 14.4 Extraction of copper and zinc from weak sulfuric acid leach solutions.
620 Cox and Reinhardt
1. Iron (II) chloride solution, for possible treatment in a pyrolysis plant or,
more probably, for the production of flocculation chemicals used in
sewage water treatment
2. Dilute hydrochloric acid condensate, mainly used as strip solution
Solvent Extraction in the Recovery of Waste 621
Fig. 14.5 Extraction of zinc from spent hydrochloric acid pickling liquors.
extract nickel and to control the buildup of impurities. Owing to the high
vanadium concentration in the feed, a large extraction capacity can be ob-
tained and, consequently, the process flows will be less for a given vanadium
throughput. The economical importance of this effect is decisive.
Vanadium is stripped from the organic solution with 1.5 mol dm3
sulfuric acid to a concentration of about 50 g dm3 V. After an iron removal
step with strong sulfuric acid, the organic solution is washed with water and
recycled. From the vanadium strip liquor, ammonium polyvanadate (APV)
is precipitated by oxidation and addition of ammonia. The APV slurry is
thickened and pumped to a vacuum belt filter, where the APV cake is
carefully washed with fresh water. The APV filter cake is dried and then
calcined to vanadium pentoxide.
To the second half of the raffinate, ammonia is added to form a nickel
ammonium complex, which is extracted with a hydroxyaryloxime (LIX84).
Due to simultaneous formation of metal hydroxides, the solution is carefully
Fig. 14.7 Extraction of vanadium and nickel from soot and flue ash.
624 Cox and Reinhardt
Fig. 14.8 Recovery of vanadium and nickel from soot, Stenungsund, Sweden.
Solvent Extraction in the Recovery of Waste 625
The amine-chloride system is well suited for the separation of the metals.
Using an evaporator for raffinate concentration and production of strip
solutions, a closed wet cycle is obtained and no aqueous waste pollutes
the environment.
The low consumption of chemicals and energy indicate an economic
advantage over similar open cycle processes.
A limited amount of available waste and a price increase of the raw material
resulted in the discontinuation of the project. This situation is not unique.
Many projects suffer from the same problem, i.e., a waste material becomes
a valuable raw material when it is in demand.
14.4.1.8 Extraction of Chromium(VI) from Surface
Finishing Wastewater
Hexavalent chromate [Cr(VI)] is still used within the industry to meet crit-
ical high corrosion control and other metal surface finishing requirements.
Cr(VI) is toxic and its control generates a hazardous, costly waste.
Recently, a solvent extraction process [14] for the recovery of hexa-
valent chromium from surface finishing process water, the A-LLX system,
Solvent Extraction in the Recovery of Waste 627
organic diluents like kerosene, the acids can be extracted from an aqueous
solution by solvent extraction. This procedure has been described in more
detail (See Chapter 10).
14.4.2.1 Recovery of Hydrofluoric and
Nitric Acids from Stainless Steel
Pickling Baths
A stainless steel pickling bath initially contains 2.2 mol dm3 HNO3 and
1.6 mol dm3 HF. The bath is used until the iron concentration reaches
40–80 g dm3. At this stage, the bath contains about 50% unused acids in
addition to the dissolved metals: iron, nickel, chromium, and molybdenum.
In a used pickling bath, most of the acids are bound to the metals as
complexes. By adding sulfuric acid to the pickling bath, sulfate will replace
some of the nitrate and fluoride, leading to the formation of extractable,
undissociated nitric and hydrofluoric acids.
The general procedure to remove monovalent acids from the aqueous
solution by adding sulfuric acid and then extracting the released acids with
an organic donor molecule, dissolved in an organic diluent, has been called
acid exchange or the AX process [15,16]. The procedure was first presented
for the treatment of spent stainless steel pickling liquids. The process con-
sists of three main steps (Fig. 14.10):
1. Addition of H2SO4 and extraction of HNO3 and HF with an organic
solution, containing 75% TBP in kerosene, in a pulsed column operation
2. Precipitation of the metals (Cr, Fe, and Ni) left in the aqueous raffinate
3. Stripping HNO3 and HF from the organic solution with water in a
second pulsed column operation and recycling the acids back to the
pickling bath
A plant for treating 600 dm3 h1 spent pickling liquid (corresponding
to an annual production of 25,000 t stainless steel) was in operation for some
years in Sweden (Fig. 14.11). The recovery operation was successful; how-
ever, the steel works are now closed.
Similar extraction procedures [17] using a mixture of trialkylphos-
phine oxides (CYANEX 923) for extraction of the mineral acids from elec-
troplating and stainless steel pickling baths are reported as the ARSEP
procedure.
The possibility of using nitric acid as an oxidizing agent in a leaching
operation is often prohibitive because many reagents used in solvent
extraction are not stable in an oxidizing solution. By the AX method,
however, a nitric acid (nitrate) leach solution can be converted to a weak
sulfuric acid (sulfate) solution, well suited for a solvent extraction treat-
ment.
Solvent Extraction in the Recovery of Waste 629
Fig. 14.11 Acid recovery by pulsed plate column operation at Söderfors Steel Works,
Sweden.
of the contaminating substances, together with the slurry from the clarifier,
is returned to the phosphoric acid plant. In this way, the impurities in the
green acid will be returned to fertilizer production without creating addi-
tional environmental problems.
Solvent Extraction in the Recovery of Waste 631
Fig. 14.13 Centrifugal extractor arrangement with five CENTREK EC-320 units.
Fig. 14.17 Copper recovery units (MECER) integrated in the production of printed
circuit boards. This photograph shows the solvent extraction unit.
Solvent Extraction in the Recovery of Waste 639
a carbon dioxide purge. After filtration of the cadmium carbonate, the fil-
trate is used for absorption of the stripped ammonia and the resulting
solution is recycled to leaching. The cadmium carbonate precipitate is used
directly in the main production.
It is necessary to bleed about 25% of the filtrate from the carbonate
precipitation to maintain material balances. This large bleed is due to the
sulfate added in the sulfuric acid leach and because the waste material
contains considerable amounts of alkali (KOH) and water. The bleed needs
special treatment to recover ammonia and to produce an environmentally
acceptable effluent.
The process has been tested at pilot plant scale with good technical
results. At the time, however, the amount of waste material available was
not sufficient to justify the feasibility of a full-scale plant.
Alternatively, an acid route has also been reported [26], where the
leaching is accomplished using a mixture of HNO3 and HF, and separation
and recovery of the nickel and cadmium is performed with solvent extraction.
First, in the cadmium circuit the leaching solution is adjusted to pH 3 prior to
640
Cox and Reinhardt
Fig. 14.19 Recovery of nickel, vanadium, and molybdenum from spent desulfurizing catalyst.
Cox and Reinhardt
Solvent Extraction in the Recovery of Waste 643
REFERENCES
1. Andersson, S. O. S.; Reinhardt, H. Handbook of Solvent Extraction; Lo, T.C.
Baird, M. H. I. Hanson, C. Eds.; John Wiley and Sons, 1982; 751 pp.
2. Reinhardt, H. Chem. Ind., 1975, 210.
3. Reinhardt, H. Proc. Int. Waste Treatment and Utilization Conf., Waterloo,
Canada, 1978; 83 pp.
4. Reinhardt, H. Proc. Hydrometallurgy ’81, Soc. Chem. Ind.: London, Section
F1, 1981.
5. Martin, D.; Diaz, G.; Garcia, M. A.; Sánchez, F. Proc. ISEC 2002; Cape
Town, South Africa, 2002, 1045.
6. Sole, K. C. Proc. ISEC 2002; Cape Town, South Africa, 2002, 1033.
7. Reinhardt, H.; Ottertun, H.; Troeng, T. I. Chemical Engineering Symposium
Series 41, I. Chem. E., Section W1, 1975.
8. MEAB Report. Extraction of nickel with simultaneous neutralization, 1989.
www.meab-mx.se.
9. Flett, D. S.; Pearson, D. Chem. Ind., 1975, 639.
10. Sze, Y. K. P.; Xue, L. Sep. Sci. Tech., 2003, 38, 405.
11. Andersson, S. O. S.; Reinhardt, H. Proc. ISEC ’77; Canadian Inst. Min. Metal.:
Toronto, 1979; Vol. 1, 798.
12. Ottertun, H.; Strandell, E. Proc. ISEC ’77; Canadian Inst. Min. Metal.:
Toronto, 1979; Vol. 1, 501.
13. Aue, A.; Skjutare, L.; Björling, G.; Reinhardt, H.; Rydberg, J. Proc. ISEC ’71;
The Hague, Soc. Chem. Ind. London, 1971; Vol. 1, 447.
14. Monzyk, B.; Conkle, H. N.; Rose, J. K.; Chauhan, S. P. Proc. ISEC 2002; Cape
Town, South Africa, 2002; 755.
15. Martin, D.; Garcia, M. A.; Diaz, G.; Falgueras, J. Proc. ISEC ’99; Barcelona,
Spain, Soc. Chem. Ind. London, 2001; Vol. 1, 201.
16. Rydberg, J.; Reinhardt, H.; Lundén, B.; Haglund, P. Proc. Int. Symp. Hydro-
metallurgy; Chicago, 1973; 589.
17. Demster, J. H.; Björklund, P. Proc. 4th Annual Meeting Hydromet. Sec. Met-
allurgical Soc.; CIM; Toronto, 1974; 68.
18. Dias, G.; Martin, D.; Frias, C.; Pérez, O. Proc. ISEC ’99; Barcelona, Spain,
Soc. Chem. Ind. London, 2001; Vol. 2, 1449.
19. Kalujta, V. V.; Kuznetsov, G. I.; Kravchenko, A. N.; Lanin, V. P.; Miraevshy,
G. P.; Pushkov, A. A.; Travkin, V. F.; Shklyar, L. I. Proc. ISEC ’88; Moscow,
1988; Vol. 1, 252.
20. Dreisinger, D. B.; Leong, B. J. Y.; Balint, B. J.; Beyad, M. H. Proc. ISEC ’93;
York, UK: Soc. Chem. Ind. London, 1993; Vol. 3, 1271.
Solvent Extraction in the Recovery of Waste 649
21. Wang, C.; Jiang, K.; Liu, D.; Wang, H. Proc. ISEC 2002; Cape Town, South
Africa, 2002; 1039.
22. Cox, M.; Flett, D. S.; Velea, T.; Vasiliu, C. Proc. ISEC 2002; Cape Town,
South Africa, 2002; 995.
23. Dreisinger, D.; Wassink, B.; Ship, K.; King, J.; Hames, M.; Hackl, R. Proc.
ISEC 2002; Cape Town, South Africa, 2002; 798.
24. Reinhardt, H. Proc. ISEC ’93; York, UK: Soc. Chem. Ind. London, 1993;
Vol. 3, 1625.
25. Reinhardt, H.; Ottertun, H. Eur Patent Nr. 0 005 415, 1982. www.sigmamecer.
com.
26. Ritcey, G. M. Proc. ISEC ’93; York, UK: Soc. Chem. Ind. London, 1993;
Vol. 1, 189.
27. Tsuboi, I.; Kunugita, E.; Komasawa, I. Proc. ISEC ’93; York, UK: Soc. Chem.
Ind. London, 1993; Vol. 3, 1319.
15
Recent Advances in Solvent Extraction
Processes
SUSANA PÉREZ de ORTIZ and DAVID STUCKEY Imperial College,
London, United Kingdom
15.1 INTRODUCTION
Recent developments in separation techniques are a response to the growing
demands imposed by extraction processes that cannot be carried out eco-
nomically using conventional technologies. Typical examples are the treat-
ment of wastewaters containing low concentrations of pollutants, such as
heavy metals, and the downstream separation of biological products. In the
case of wastewaters, environmental regulations impose discharge con-
centrations of the order of a few parts per million, or even per billion, for
certain pollutants. Therefore, treatment using conventional solvent extrac-
tion would require large volumes of solvent and several extraction stages to
achieve the target concentrations, even in systems with large metal dis-
tribution coefficients (see Chapter 4).
The suitability of using solvent extraction for a given separation is
determined by thermodynamic and kinetic considerations. The main ther-
modynamic parameter is the solute distribution ratio, DM, between the
organic and the aqueous phase. This is given by [Eq. (4.3), Chapter 4]:
DM ¼ ½MT;org =½MT;aq (15.1)
where [M]T is the sum of the concentrations of all M-species in a given
phase, and the second subscript indicates the organic and the aqueous
phase. The magnitude of DM determines the feasibility of the separation as
an industrial process; the higher DM the better the solute separation.
Another consideration affecting the design of extraction processes is
the extraction rate as it determines the residence time of the phases in the
651
652 Pérez de Ortiz and Stuckey
Microemulsions
Colloidal liquid aphrons
(2) diffusion of the extracted species through the membrane; and (3) strip-
ping into the third phase. As in conventional solvent extraction, solute
transfer into the membrane can be achieved by selective solubility or by
chemical reaction with a component in the liquid membrane. Stripping at
the interface of the liquid membrane with the third phase can be equally
achieved by either physical or chemical mechanisms. The mechanism of
solute separation requires analysis as it affects the conditions that control
the overall rate of transfer between the first and the third phase.
Application of Eq. (15.1) to the liquid membrane process highlights
one of the main advantages of the process, i.e., the high solute distribution
coefficient that can be obtained between phases 3 and 1. However, another
factor that must be considered when evaluating a separation process per-
formance is the kinetics of transfer, which is given in a general form by
Eq. (15.4). This equation indicates that the transfer rate in the contactor
increases with both the interfacial flux and the specific interfacial area.
The two main mechanisms of solute separation by liquid membranes
involve chemical reactions. In the case illustrated in Fig. 15.2a, the solute
first dissolves in the liquid membrane, then diffuses toward phase 3 due to the
buildup of a concentration gradient, and finally transfers to phase 3 at the
Fig. 15.2 Basic mechanisms of liquid membrane extraction: (a) type I facilitated
transport (A + B?AB); (b) type II facilitated transport (A + B , B + C).
656 Pérez de Ortiz and Stuckey
with the liquid membrane phase; (2) dispersion of the emulsion into the feed;
(3) separation of the emulsion from the raffinate phase; and (4) demulsifi-
cation. This final stage separates the stripping solution that contains the
species extracted from the feed, from the liquid membrane phase, which is
recycled to the emulsification stage.
In common with the supported liquid membrane, the emulsion liquid
membrane yields a solute partition coefficient of a higher order of magni-
tude than that obtained with the conventional solvent extraction process,
thus allowing a high separation percentage from dilute feeds, and con-
centrated stripping solutions in just one contact stage. However, the process
has its disadvantages; one is that the need to produce a stable emulsion
requires the use of additives that slow down the rate of extraction and, even
if their solubility is negligible, they may contaminate the raffinate. Another
disadvantage is the problem posed by emulsion rupture in the contactor.
Emulsion rupture is usually due to emulsion swelling caused by the
transport of the external phase into the emulsion. Three different mecha-
nisms have been identified as causes of emulsion swelling: (1) occlusion due
to entrainment of the external phase [8]; (2) secondary emulsification of the
external phase caused by an excess of surfactant in the liquid membrane
phase; and (3) external phase permeation through the liquid membrane
[9,10]. The latter includes osmosis and, in the case of external aqueous feed,
the transport of hydration water attached to complexes and water transport
due to the presence of reverse micelles in the organic membrane. Swelling
and emulsion rupture can be greatly decreased by including additives and
especially designed components to the membrane phase.
the dispersion. The aggregates formed in the aqueous phase, called micelles,
have their molecules orientated with their hydrophobic tails pointing to the
interior of the aggregate and their hydrophilic head toward the continuous
aqueous phase, whereas the aggregates in the organic phase, called reverse
micelles, have the opposite orientation, as shown in Fig. 15.4. The micro-
emulsion droplets are therefore the cores of either micelles or reverse
micelles, stabilized by a surfactant layer. If more components besides the
surfactant are present in the system, they may also be incorporated into
the micelles or reverse micelles; these are then called mixed micelles or
reverse micelles.
The ternary equilibrium diagram in Fig. 15.4 illustrates the effect of
component concentration on the structure and the number of phases in
substantially reduce the metal distribution coefficient and the extraction rate,
or leave them unchanged. An interesting example is found in the extraction
with di(2-ethylhexyl)phosphoric acid (DEHPA). DEHPA does not form
microemulsions in aliphatic solvents at pH 4 or below; however, it forms
microemulsions on addition of a surfactant and a cosurfactant, e.g., sodium
dodecylbenzene sulfonate and n-butanol. Bauer et al. [20] reported sub-
stantial improvements in the extraction of trivalent and quadrivalent metals
with respect to the conventional system with a DEHPA microemulsion.
However, Brejza and Pérez de Ortiz [21] obtained improvements in the
extraction of Al(III) using the same microemulsion, in contrast with the
extraction of Zn(II), which was reduced significantly with respect to the
single DEHPA system. A similar different behavior of a microemulsion in
the extraction of Bi(III) and Zn(II) was observed by Pepe and Otu [22]; they
observed an increase in the extraction of the trivalent metal, but not for zinc.
Brejza and Pérez de Ortiz [21] attempted to explain the contrasting effect of
the microemulsion in the extraction of aluminum and zinc with DEHPA
based on the different interfacial behavior of their complexes: the aluminum
complex is more hydrophilic, therefore it may have a greater desorption
energy than the zinc one, making the interface its preferred location, whereas
the zinc complex is more soluble in the organic phase. Therefore, in the
microemulsion system it would be preferentially solubilized in the palisade
rather than in the micellar cores, leading to an enhancement of extraction due
to the large increase in interfacial area produced by the microemulsion. On
the other hand, the zinc complex is more soluble in the organic phase than in
the aqueous phase or the palisade, and consequently the presence of the
aqueous microphase does not improve its solubility in the extracting phase.
This hypothesis remains to be confirmed.
One important point regarding microemulsion extraction is that the
complexity of the system with its three phases, two interfaces, and usually
unknown phase morphology makes the prediction of its performance quite
difficult, particularly when dealing with solutes of diverse properties. The
literature indicates that it does not always improve metal extraction, and in
cases it may even hinder it due to the effect of the emulsifying additives.
15.5.3.2 Extraction of Biological Molecules
The mechanism of separation of biological molecules such as proteins and
amino acids, and the parameters that affect the extraction distribution coef-
ficient and the kinetics of extraction have been studied more extensively than
the extraction of inorganic solutes. This is mainly due to the variety of size
and structure of these molecules and, furthermore, to the fact that their
characteristics may be adversely affected by their contact with solvents and
surfactants.
Recent Advances in Solvent Extraction Processes 665
increase in the number and size of the reverse micelles. The work of Hentsch
et al. [30] showed that the back transfer of a-chymotrypsin decreased with
decreasing AOT concentration. They suggested that this decrease was due
either to a-chymotrypsin being trapped in the emulsion, or to denaturation.
Salt type and concentration: For back-extraction, increases in pH are
not enough to strip the protein out from reverse micelles; this is also due to
the size exclusion effect resulting from a decrease in the reverse micelle size
[31,32]. This means that high salt concentration and salts that form small
reverse micelles favor back transfer. Most of the work reported in the lit-
erature used KCl solution, normally 1.0 mol dm3 KCl coupled with a pH
around 7.5. Marcozzi et al. [23] also showed that the back transfer efficiency
of a-chymotrypsin depends on the salt type and concentration used in the
forward transfer.
Counterion extraction: Due to the relative slowness of back extraction
based on the methods above, the back-extraction of proteins encapsulated in
AOT reverse micelles was evaluated by adding a counterionic surfactant,
either TOMAC or DTAB, to the reverse micelles [33]. This novel backward
transfer method gave higher backward extraction yields compared to the
conventional method. The back-extraction process with TOMAC was found
to be 100 times faster than back-extraction with the conventional method,
and as much as three times faster than forward extraction. The 1:1 complexes
of AOT and TOMAC in the solvent phase could be efficiently removed using
adsorption onto montmorillonite so that the organic solvent could be reused.
negligible levels [44]. Brejza found that the effect of the microemulsion on
the kinetics of extraction depends on the characteristics of the metal ion and
its complex, e.g., valence and degree of complex hydration, and on the
solubility of the complex in the organic phase [44]. Thus there are no general
rules as to the advantage of forming a microemulsion in a system that
contains a reagent.
In the case of protein transfer, early work by Bausch et al. [42] found
that transport was controlled by convective processes in the aqueous phase,
and the mass transfer coefficient increased with increasing surfactant but
depended strongly on stirring speed. Poppenborg et al. [45] evaluated the
kinetic separation of lysozyme and cytochrome in a Graesser contactor and
Lewis cell, and with a low rotor speed (2–3 rpm), low temperature (48C),
and a pH close to the pI of both proteins (pH 9–10), about 83% of the
lysozyme and only 11% of the cytochrome c was extracted into the reverse
micellar phase after 30 minutes. This optimum separation was based on the
effect pH changes have on the extraction kinetics, and the rate of cyto-
chrome c extraction was reduced much more than for lysozyme when the
pH approached the pI of cytochrome c, and differed markedly from the
effect the pH had on phase distribution. The extraction rate measured
in the Graesser contactor differed from that measured in the Lewis
cell, i.e., lysozyme was extracted faster than cytochrome c and this obser-
vation indicates that different steps of the reverse micellar transfer mecha-
nism are controlling the transfer, depending on the way the phases are
contacted.
The kinetics of back extraction are equally important to obtain a
better understanding of the mechanism of solute transfer, and to determine
the rate-limiting step for the process. Such information is crucial for the
rational design of an extraction apparatus and, as discussed above, Jar-
udilokkul et al. [33] showed that counterion extraction resulted in remark-
able increases in the back-extraction of proteins.
effects on aphron stability. First, it would reduce the surface charge on the
aphrons and hence the energy barrier to droplet coalescence. It had pre-
viously been found that the zeta potential of a CLA suspension falls from
45 mV at pH 8.4 to 36 mV at pH 4 [68]. Secondly, protonation of the
head groups reduces the polarity of the surfactant monomers, making it
energetically less favorable for the hydrophobic tails of the surfactant
molecules to remain in an aqueous environment. Experimentally, it is easy
to show that the solubility of SDS in 20 mmol dm3 buffer solutions rapidly
decreases below pH 7, which may also be responsible for decreasing CLA
stability at low pH. If the collision-coalescence mechanism proposed is
correct, then the stability of CLAs should depend upon the temperature at
which this process occurs. This was indeed found experimentally with t1/2
values falling from 96 to 4 min for a corresponding increase in temperature
of the continuous phase from 108C to 608C [65]. An Arrhenius plot resulted
in a linear relationship between Ln k and 1/T. Performing linear regression
yields a single value of the activation energy, Ea, for the collision process of
50 kJ mol1. A value for Ea of this magnitude would suggest that, over the
temperature range investigated, the collision-fusion process of the CLAs is
controlled by both diffusion and chemical reaction.
Concerning the structure of dispersed CLAs, the model originally
proposed by Sebba [57] of a spherical oil-core droplet surrounded by a thin
aqueous film stabilized by the presence of three surfactant layers is, in our
opinion, essentially correct. However, there is still little direct evidence for
the microstructure of the surfactant interfaces. From an engineering point of
view, however, there is now quantitative data on the stability of CLAs
which, together with solute mass transfer kinetics, should enable the suc-
cessful design and operation of a CLA extraction process.
should not occur due to the strong negative charge on the CLA, and the
generally negative charge on most bacterial cells. This situation is commonly
faced in new biotransformation reactions, and even when the product is
polar the use of CLAs will enhance the transfer of substrate.
The final intriguing use of CLAs is in the immobilization of enzymes in
the soapy shell in order to carry out an enzymatic reaction. Thus the hydrol-
ysis of p-nitrophenyl acetate to p-nitrophenol has been demonstrated by
immobilizing a lipase into the shell of a CLA. The CLAs were then pumped
through a cross-flow membrane, where they were separated and recycled,
with the product appearing in the permeate [70].
REFERENCES
1. Li, N. N. U.S. Patent 3,310,794, November 12, 1968.
2. Li, N. N. Ind. Eng. Chem. Process Des. Dev., 1971, 10(2), 215–221.
3. Cahn, R. P.; Franfeld, J. W.; Li, N. N.; Naden, D.; Subramanian, K. N. Recent
Developments in Separation Science; Li, N. N. Ed., CRC Press: Boca Raton,
Florida, Vol. VI, 1981; p. 51.
4. Cahn, R. P.; Li, N. N. Sep. Sci., 1974, 9(6), 505–519.
5. Sharma, A.; Goswami, A. N.; Rawat, B. S. J. Membr. Sci., 1991, 60, 261.
6. Gallego Lizon, T.; Perez de Ortiz, E. S. Ind. Chem. Eng. Res. Dev., 2000, 39,
5020–5028.
7. Marr, R.; Kopp, A. Int. Chem. Eng., 1982, 22, 44–60.
8. Kinugasa, T.; Tanahashi, S. I.; Takeuchi, H. Ind. Eng. Chem. Res.,1991, 30,
2470.
9. Colinart, P.; Delepine, S.; Trouve, G.; Renon, H. J. Membr. Sci.,1984, 20, 167–
187.
10. Ding, X. C.; Xie, F. Q. J. Membr. Sci.,1991, 59, 183–188.
11. Bloch, R. Hydrometallurgical Separations by Solvent Membranes; Flynn, J. E.
Ed., Membrane Science and Technology, Plenum Press; New York, 1970.
12. Danesi, P. R. Sep. Sci. Technol., 1984–1985, 19(11–12), 857–894.
Recent Advances in Solvent Extraction Processes 677
13. Danesi, P. R. Proc. ISEC ’86, Munich: DECHEMA, 1986; 527–535 pp.
14. Pabby, A. K.; Sastre, A-M. Ion Exchange and Solvent Extraction; Marcus, Y.
and Sengupta, A. K. Eds., Marcel Dekker: New York, 2002; vol. 15, 331–469.
15. Winsor, P. A. Solvent Properties of Amphiphilic Compounds; Butterworth Sci-
entific: London, 1954.
16. Zulauf, M.; Eicke, H. F. J. Phys. Chem., 1979, 83(4), 480.
17. Leodidis, E. B.; Hatton, T. A. Langmuir, 1989, 5, 741.
18. Leodidis, E. B.; Hatton, T. A. The Structure and Reactivity in Reverse Micelles;
Pileni, M. P. Ed., Elsevier; Amsterdam, 270, 1989, 270p.
19. Osseo-Asare, K.; Kenney, M. E. Proc ISEC ’80, Liege, 1980, 1, paper no. 80,
121.
20. Bauer, D.; Cote, G.; Komornicki, J.; Mallet-Faux, S. Can. Soc. Chem. Eng.,
1989, 2, 425.
21. Brejza, E. V.; Perez de Ortiz, E. S. J. Colloid Interface Sci., 2000, 227, 244–246.
22. Pepe, E. M.; Otu, E. O. Solv. Extr. Ion Exch., 1996, 14(2), 247.
23. Marcozzi, G.; Correa, N.; Luisi, P. L.; Caselli, M. Biotechnol. Bioeng., 1991,
38, 1239.
24. Woll, J. M.; Hatton, T. A. Bioprocess Eng., 1989, 4, 193.
25. Jarudilokkul, S.; Poppenborg, L. H.; Stuckey, D. C. Sep. Sci. Tech., 2000, 35,
503–517.
26. Luisi, P. L.; Bonner, F. J.; Pellergrini, A.; Wiget, P.; Wolf, R. Helv. Chim. Acta,
1979, 62, 740.
27. Dekker, M.; van’t Riet, K.; Van Der Pol, J. J.; Baltussen, J. W. A.; Hilhorst, R.;
Bijsterbosch, B. H. J. Chem. Eng., 1991, 46, B69.
28. Woll, J. M.; Hatton, T. A.; Yarmush, M. L.; Biotechnol. Prog., 1989, 5(2), 57.
29. Kelley, B. D.; Wang, D. I. C.; Hatton, T. A. Biotechnol. Bioeng., 1993, 42(10),
1199.
30. Hentsch, M.; Menoud, P.; Steiner, L.; Flaschel, E.; and Renken, A. Bio/
Technol., 1992, 6(4), 359.
31. Goklen, K. E. Ph.D. Thesis, Massachussetts Institute of Technology, 1986.
32. Dekker, M.; van’t Riet, K.; Baltussen, J. W. A.; Bijsterbosch, B. H.; Hilhorst,
R. Laane, C. in Proc. 4th European Conf. on Biotechnology, Amsterdam, 1987,
2, 507.
33. Jarudilokkul, S.; Poppenborg, L. H.; Stuckey, D. C. Biotech. Bioeng., 1999,
62(5), 593–601.
34. Battistel, E.; Luisi, P. L. J. Colloid Interface Sci., 1989, 128(1), 7.
35. Nitsch, W. Plucinski, P. J. Colloid Interface Sci., 1990, 136(2), 338.
36. Albery, W. J.; Choudhery, R. A.; Atay, N. Z.; Robinson, B. H. J. Chem. Soc.
Faraday Trans. 1, 1987, 83(8), 2407.
37. Savastano, C. A.; Osseo-Asare, K.; Pérez de Ortiz, E. S. Separation Processes in
Hydrometallurgy; Davis, G. A. Ed., SCI; London, 1987; p. 89.
38. Kim, H. S.; Tondre, C. Sep. Sci. Technol., 1989, 24, 485.
39. Plucinski, P.; Nitsch, W. Ber. Bunsenges. Phys. Chem., 1989, 93, 994.
40. Plucinski, P.; Nitsch, W. J. Phys. Chem., 1992, 97, 8983.
41. Plucinski, P.; Nitsch, W. Langmuir, 1994, 10, 371.
678 Pérez de Ortiz and Stuckey
42. Bausch, T. E.; Plucinski, P. K.; Nitsch, W. J. Colloid Interface Sci., 1992,
150(1), 226.
43. Dekker, M.; van’t Riet, K.; Bijsterbosch, B. H.; Fijneman, P.; Hilhorst, R.
Chem. Eng. Sci., 1990, 45(9), 2949.
44. Brejza, E. V. Doctoral Thesis, University of London, 1994.
45. Poppenborg, L. H.; Brillis, A.; Stuckey, D. C. Sep. Sci. Tech., 2000, 35, 843–858.
46. Goklen, K. E.; Hatton, T. A. Proc. ISEC ’86, Munchen, 1986, 3, 587.
47. Rahaman, R. S.; Lee, J. Y.; Cabral, J. M. S.; Hatton, T. A. Biotechnol. Prog.,
1988, 4(4), 218.
48. Krei, G. A.; Hustedt, H. Chem. Eng. Sci., 1992, 47(1), 99.
49. Jarudilokkul, S.; Paulsen, E.; Stuckey, D. C. Biosep., 2000, 9, 81–91.
50. Laane, C.; Dekker, M. Proc. 6th Int. Symp. on Surfactants in Solution; Mittal,
E. L., Ed., Plenum Press: New York, 9, 1989; 9p.
51. Giovenco, S.; Verheggen, F.; Laane, C. Enzyme Microb. Technol., 1987, 9, 470.
52. Jarudilokkul, S.; Poppenborg, L. H.; Valetti, F.; Gilardi, G.; Stuckey, D. C.
Biotech. Techniques, 1999, 13, 159–163.
53. Lye, G. Y. Ph. D Thesis, University of Reading, 1993.
54. Jarudilokkul, S.; Paulsen, E.; Stuckey, D. C. Biotech. Bioeng. 2000, 69, 618–
626.
55. Jarudilokkul, S.; Paulsen, E.; Stuckey, D. C. Biotech Progress, 2000, 16, 1071–
1078.
56. Jarudilokkul, S.; Stuckey, D. C. Sep. Sci. Technol., 2001, 36, 657–670.
57. Sebba, F. Foams and Biliquid Foams-Aphrons; John Wiley: New York, 1987.
58. Sebba, F. J. Colloid Interface Sci., 1972, 40, 468.
59. Sebba, F. Colloid Polymer Sci., 1979, 257, 392.
60. Lye, G. J.; Stuckey, D. C. Separations for Biotechnology III; Pyle, D. L. Ed.,
SCI: London, 1994; 280–286.
61. Wallis, D. A.; Michelsen, D. L.; Sebba, F.; Carpenter, J. K.; Houle, D. Biotech.
Bioeng. Symp. 1985, 15, 399.
62. Lye, G. J.; Poutiainen, L. V.; Stuckey, D. C. Biotechnology ’94, 2nd Int. Sym-
posium on Environmental Biotechnology; I. Chem. E., II, 1994; 25–27.
63. Matsushita, K.; Mollah, A. H.; Stuckey, D. C.; del Cerro, C.; Bailey, A. I.
Colloids Surfaces, 1992, 69, 65–72.
64. Chaphalkar, P. G.; Valsaraj, K. T.; Roy, D. Sep. Sci. Technol., 1993, 28, 1287.
65. Lye, G. J.; Stuckey, D. C. Colloids Surfaces, A: Physiochem. Eng. Aspects,
1998, 131(1–3), 113–130.
66. Scarpello, J. T.; Stuckey, D. C. J. Chem. Tech. Biotech., 1999, 74, 409–416.
67. Save, S. V.; Pangarkar, V. G. Chem. Eng. Comm., 1994, 127, 35.
68. Stuckey, D. C.; Matsushita, K.; Mollah, A. H.; Bailey, A. I. Third Int. Conf. on
Effective Membrane Processes: New Perspectives; University of Bath, U.K.,
1993.
69. Lye, G. J.; Stuckey, D. C. Chem. Eng. Sci., 2001, 56, 97–108.
70. Lye, G. J.; Pavlou, O. P.; Rosjidi, M.; Stuckey, D. C. Biotech Bioeng., 1996, 51,
69–78.
16
Computational Chemistry in Modeling
Solvent Extraction of Metal Ions
JERZY NARBUTT Institute of Nuclear Chemistry and Technology,
Warsaw, Poland
MARIAN CZERWIŃSKI Zawiercie University of Administration and
Management, Zawiercie, Poland
16.1 INTRODUCTION
Great progress in computational techniques and methods in the last decade
of the 20th century has made it possible to perform quantum mechanical
(QM) calculations on large molecular systems composed of several dozens
or even hundreds of atoms. General ab initio schemes based on Hartree-
Fock-Roothaan (HFR) methods [1], formerly used for small molecules
composed of light elements, now are becoming applicable for metal com-
plexes, including even heavy metals. Calculations based on the density-
functional theory [2] have appeared particularly fruitful in this respect.
These methods have also been applied in solvent extraction chemistry.
Reviews have been presented, e.g., [3,4], on the computational methods-
from purely empirical molecular mechanics (MM) approaches to advanced
ab initio calculations rendered accessible in the form of commercial software
packages: Cerius2 [5], Gaussian 98 [6], ADF [7], Spartan [8], HyperChem [9],
etc. These methods and programs have been widely used for structural
optimization of metal complexes and calculation of their energy of forma-
tion, problems of primary importance in solvent extraction.
Numerical methods have been used in solvent extraction chemistry for
treating experimental data for many years. As shown in Chapter 4 (see
section 4.14.3), modeling of extraction processes in terms of a set of assumed
equilibria with adjustable (best-fit) coefficients was widely used to identify
and characterize the major species formed in these processes. The im-
provement of this approach, directed toward studying systems of greater
diversity and complexity, including corrections for thermodynamic activity
679
680 Narbutt and Czerwiński
methods, based upon classical physics, use simple expressions and param-
eters describing interatomic forces in the molecule and also between mole-
cules. To find the optimized structure of a molecule, the total strain energy
accompanying the formation of the molecule must be minimized. This total
strain energy includes the sum of the energies of deformation of all bond
lengths and bond angles; the torsional energies related to barriers to rotation
of the molecule fragments about bonds; and the energies related to non-
bonded interactions between the molecule fragments, i.e., terms describing
van der Waals (Lennard-Jones potential) and long-range electrostatic
(Coulomb potential) interactions [Eq. (16.1)]:
X X X
E¼ Kr (r req )2 þ K ( eq )2 þ Vn [1 þ cos (n)]
bonds angles dihedrals
" #
X Aij Bij qi qj
þ 6þ (16.1)
nonbonded
R12
ij Rij "Rij
The initial set of parameters for a selected force field, including the ideal
bond lengths and bond angles, the related force constants, the effective
barriers to rotation, point charges and van der Waals parameters of the
atoms, and effective dielectric constant of the medium, are obtained from
experimental data and/or QM calculations, and their performance is tested
by comparing the calculated and experimental structures and energies. If
necessary, the procedure is repeated to minimize the errors in the calculated
output [13]. In some MM models a separate term is added, describing
hydrogen-bonding potentials; however, the hydrogen bond energies calcu-
lated in such a way are usually approximate.
Optimizing the geometry of metal complexes in the gas phase neglects
environmental effects (solvent, concentration), important in the analysis of
both structure and stability of the complexes because of their interactions
with neighboring molecules. To study complex formation in solution, some
MM models use the parameterization of the entire force field, based on
experimental data. In this case, the environment is implicitly included in the
model and the environmental effects are isotropic [14]. Another way to
account for the environment is to calculate through-space attractive and
repulsive intermolecular interactions, by the use of the last right-hand-side
term of Eq. (16.1), parameterized for the whole system, and, if related, also
the term describing hydrogen bonding.
MM methods, originally developed for organic compounds, have been
modified to describe metal complexes [13–17] where polarization interac-
tions must be considered and where a variety of geometrical configuration is
observed. This structural diversity reflecting the existence of multiple local
energy minima is due to ligand-dependent effects observed in complexes of
682 Narbutt and Czerwiński
certain transition metal (TM) ions: Jahn-Teller effect, spin crossover [18,19],
and stereochemical activity of valence s-electron pairs in some p-block heavy
metal ions in lower oxidation states [20]. Various types of potential energy
equations, and methods developing the parameters, are used to determine
force fields for coordination compounds. It is generally assumed and con-
firmed for a variety of coordinated ligands that the parameters for metal
complexes are transferable from the organic force fields [13]. Modification
and extension of MM by the effect of ligand field stabilization energy on the
strain energy in complexes of certain TM ions (the LFMM model) makes it
possible automatically to take into account Jahn-Teller distortions, and to
reproduce with a reasonable accuracy both the structures and the associated
d-d transition energies [18,19]. The not very high but often sufficient accu-
racy of MM calculations is recompensed by lesser requirements for com-
putational resources, shorter calculation time, and a possibility to deal with
larger systems.
satisfy the coordination and geometric requirements of the ion, the geom-
etry of ligands (e.g., bite size), and orientation of the donor atoms [28].
More advanced semiempirical molecular orbital methods have also
been used in this respect in modeling, e.g., the structure of a diphosphonium
extractant in the gas phase, and then the percentage extraction of zinc ion-
pair complexes was correlated with the calculated energy of association of
the ion pairs [29]. Semiempirical SCF calculations, used to study structure,
conformational changes and hydration of hydroxyoximes as extractants of
copper, appeared helpful in interpreting their interfacial activity and the rate
of extraction [30]. Similar (PM3, ZINDO) methods were also used to model
the structure of some commercial extractants (pyridine dicarboxylates,
pyridyloctanoates, b-diketones, hydroxyoximes), as well as the effects of
their hydration and association with modifiers (alcohols, b-diketones) on
their thermodynamic and interfacial activity [31–33]. In addition, the
structure of copper complexes with these extractants was calculated [32].
Semiempirical and ab initio calculations were used to interpret experi-
mental data and to find structure-complexation relationships in the extrac-
tion of transition metals and lanthanides [34–38]. The calculations contribute
to understanding the factors that affect the strength of metal–ligand bond-
ing, and to designing new ligands. For example, the extraction ability and
selectivity of some b-diketonate ligands toward lanthanides is shown to be a
function of ligand structure (bite size) and the energy of intramolecular
hydrogen bonds [37,38]. Semiempirical PM3-tm and DFT B3LYP calcula-
tions in the gas phase made it possible to determine the experimentally
unavailable molecular structure of some heterometallic TM complexes
extracted by trioctylphosphine oxide into hexane [39].
Semiempirical QM calculations of the geometry, charge distribution,
bond orders, dipole moments and HOMO energy for a series of monoamide
extractants in the gas phase, carried out at the Hartree-Fock level, were
supplemented by MM calculations on uranyl complexes with these ligands
[40,41]. The observed qualitative relationships between the calculated values
and the experimental distribution ratios of the metal have emphasized
the importance of electron density on the coordinating atoms or groups,
and steric effects in the ligands. Ligand basicity is a factor that affects the
complex stability. Semiempirical calculations appeared insufficiently accu-
rate to calculate basicities of several malonamides, the extractants of tri-
valent lanthanides and actinides; therefore advanced ab initio methods
were used, which brought about a good agreement between the calculated
(relative) values and the experimental data [42]. Conformational ab initio
analysis of the protonated terpyridine extractant used for lanthanide/
actinide separation (ion-pair extraction from strongly acidic solution) has
confirmed stabilization of the trans,trans form of the terpyridine cation in
Modeling Solvent Extraction of Metal Ions 685
Fig. 16.1 Complex of Cs+ Pic with calix-[4]-bis-crown-6 at the chloroform (left) and
water (right) interface. A snapshot at 500 ps of MD. (A. Varnek and G. Wipff, unpub-
lished).
with the stability constant, bn, of the complex, defined by Eq. (3.5). The
concentration of ‘‘free’’ ligand in the anionic form, [L], depends on the
distribution of the extractant HL between the phases and its dissocia-
tion in the aqueous phase:
HL $ HLorg (16.4)
HL $ Hþ þ L (16.5)
with the partition constant (KD,HL) and the dissociation constant (Ka) of
the extractant, defined by Eqs. (4.17) and (4.18), respectively.
2. Partition, i.e., transfer of the complex from the aqueous to the organic
phase, accompanied [77] by at least partial dehydration
MLn ðH2 OÞkm $ MLn;org þ ðk mÞH2 O (16.6)
with the equilibrium constant equal to the partition constant, KDM, of
the complex, also called the distribution constant, and defined by Eq.
(4.1).
Taking into account all these relationships, one can express the
extraction constant [the equilibrium constant in Eq. (16.2)] as:
Kex ¼ n KDM ðKa Þn ðKD;HL Þn (16.7)
When extraction of coordinatively unsaturated chelates proceeds in the
presence of a coextractant (synergist), i.e., a lipophilic neutral electron
donor, B, which additionally coordinates the metal ion, a third step
must be added:
3. Adduct formation in the organic phase:
MLn;org þ pBorg $ MLn Bp;org (16.8)
with the equilibrium constant (the stability constant, bn,p, of the adduct)
given by:
½MLn Bp org
n;p ¼ (16.9)
½MLn org ½Bporg
3
Fig. 16.2 3D contours of bonding molecular orbitals in the tris(2,3-butanedionato)
yttrium(III)-TMPO adduct-a model of tris(tropolonato)yttrium(III)-TOPO: (a) MO of s
symmetry, Es = 0.4006 a.u.; (b) MO of p symmetry, Ep = 0.2874 a.u. (From Ref. 98;
with permission from Wiley-VCH Verlag Weinheim, Germany.)
702 Narbutt and Czerwiński
phase is the neutral chelate, MLn, the bn values influence neither the dis-
tribution ratios, nor the SF value, which then becomes dependent only on
the KDM ratio [77].
Of course, one has to recognize that for practical separation a high SF
value is the prerequisite but not the sufficient condition. To separate selec-
tively two metal ions, we must ensure their distribution ratios in the system to
be respectively equally greater and less than unity. All these considerations
show how many factors one must control to assure high selectivity of solvent
extraction separations of metal ions. In the case of coordinatively unsatu-
rated metal chelates, the increasing concentration of a synergist in the
extraction system decreases, as a rule, the selectivity. For example, scandium
and yttrium tris-tropolonates undergo partition with KDM = 0.381 and
0.0020, respectively, in the system toluene/102 mol dm3 tropolone/water
( pH = 5 ± 1) at 298 K [98]; therefore SFSc/Y = 190. Taking into account the
stability constants [Eq. (16.9)] of their TOPO adducts [logb3,1 = 4.27 for
Sc(trop)3; logb3,1 = 4.99 and logb3,2 = 7.39 for Y(trop)3] we arrive at the
following calculated values for DSc, DY and SFSc/Y: 1.09, 0.022 and 50 at 104
mol dm3 TOPO; 7.48, 0.246 and 30 at 103 mol dm3 TOPO; 71.3, 6.87 and
10.4 at 102 mol dm3 TOPO, respectively. In spite of the decrease in SFSc/Y
with increasing TOPO concentration, the optimum separation can be
achieved at ca. 103 mol dm3 TOPO. The effects become greater for ions
that show greater differences in b3,1, e.g., Sc3+ and In3+ (logb3,1 = 4.27 and
0.97 [98]). However, the unexpectedly increased selectivity in the synergistic
Am3+–Eu3+ system with CYANEX 301 and some O-bearing weak-base
coextractants [45] shows that still other factors, for example different partic-
ipation of the lanthanide and actinide f-orbitals in bonding, must be taken
into account when considering the selectivity of separation.
There is, therefore, a good deal of work for theoretical chemists,
dedicated to merging the calculations of the separate steps of extraction of
metal ions into a whole. Only such a united approach will allow us to
analyze correctly all the factors that affect the thermodynamics of extrac-
tion. The greatest challenge stems from the need to evaluate solvent effects
by the use of more accurate, explicit solvation models. That will require
quantum mechanical calculations on extremely large systems consisting of
many hundred molecules, thousands of atoms.
calculations and the experiment. It has been argued before that in the case of
coordinatively unsaturated lanthanide complexes, the contribution from the
partition step increases the differences in the Kex values as well as the
selectivity of lanthanide separation.
The necessity of considering the whole extraction process has been
pointed out, however, calculations dealing with one particular extraction
steps can also be important. For example, the authors’ DFT calculations on
adduct formation in extraction systems [98] well correlate with experimental
data on the synergistic enhancement of partition constants of metal(III)
tropolonates by trioctylphosphine oxide. In spite of simplifications (in con-
trast to the experimental conditions, the calculations were performed with
models for tropolonate and TOPO, in the gas phase, and for 0 K), the com-
puted order of the energies of adduct formation (Y > Sc > Tl > In; no Ga
adduct) agreed with the experimental order of the adduct stability constants.
The calculated energies of adduct formation for tris-tropolonates of scan-
dium and yttrium are significantly less negative than the experimental free
energy changes of adduct formation (which also include the entropy term),
but the differences between the respective values, calculated DEadd(Y)
DEadd(Sc) = 5.5 kJ mol1 and experimental DGadd(Y) DGadd(Sc) = 4.2 kJ
mol1 [98] are comparable. Another work [94] points to a reasonable
agreement between the relative values of standard enthalpy and free energy
changes of transfer (heptane ? water) of tris-acetylacetonates of scandium
and cobalt(III); the experimental [92] and the calculated (DFT, continuous
solvation model; see section 16.2.5) for model compounds.
16.4.6.2 Artificial Neural Networks
Another, very different calculation method, which belongs to the field of
artificial intelligence, is till now only occasionally used in solvent extraction.
It consists in the operation of artificial neural networks (NN), which are
subject to appropriate procedures of construction and learning. NN com-
puting techniques arose from attempts to model the functioning of the
human brain and have evolved into a powerful tool for solving even the
most complex, nonlinear problems, which are difficult to handle by standard
modeling. This nonparametric method offers universal approximators able
to represent any, even the most complex, implicit functional relationship.
The calculations are based on the analysis of large sets of (empirical) data,
part of which is used for training the system, the rest for testing. The
widespread back propagation networks consist of large numbers of com-
putational units, ‘‘neurons,’’ connected by means of artificial synapses
characterized by matrices of adjustable weights. These connections and the
transfer functions of the neurons form distributed representations between
the input and output data. The required representation is selected during the
706 Narbutt and Czerwiński
ACKNOWLEDGMENTS
The authors would like to thank Alexandre Varnek for fruitful discussion,
and also Mariusz Bogacki, Yizhak Marcus, Bruce Moyer, Claude Musikas,
Jan Rydberg, and Slawomir Siekierski for their valuable comments on the
draft of this chapter at various stages of writing. The work was supported in
parts by the Institute of Nuclear Chemistry and Technology and by the
Polish Committee for Scientific Research (KBN) under grant number 4
T09A 110 23.
REFERENCES
1. Roothaan, C. C. J. Rev. Mod. Phys. 1951, 23, 69.
2. Parr, R. G.; Yang, W. Density-Functional Theory of Atoms and Molecules;
Oxford University Press: Oxford, 1989.
3. Deeth, R. J. Struct. Bonding 1995, 82, 1.
4. Rambusch, T.; Hollmann-Gloe, K.; Gloe, K. J. Prakt. Chem. 1999, 341, 202.
5. Cerius2 Version 3.5; Molecular Simulations Inc., San Diego, CA.
6. GAUSSIAN 98; Revision A.5, Gaussian Inc., Pittsburgh, PA, 1998.
7. ADF Version 2.3; Scientific Computing and Modeling, Vrije Universiteit,
Amsterdam.
8. Spartan Version 5.0.3; Wavefunction Inc., Irvine, CA.
9. HyperChem Version 5.02; Hypercube Inc., Gainesville, Florida.
10. Baes, C. F. Jr., Solv. Extr. Ion Exch. 2001, 19, 193.
11. Moyer, B. A.; Baes, C. F. Jr.; Case, F. I.; Driver, J. L. Solv. Extr. Ion Exch.
2001, 19, 757.
12. Varnek, A.; Wipff, G.; Solov’ev, V. P. Solv. Extr. Ion Exch. 2001, 19, 791.
13. Hay, B. P. Coord. Chem. Rev. 1993, 126, 177.
14. Boyens, J. C. A.; Comba, P. Coord. Chem. Rev. 2001, 212, 3.
15. Brubaker, G. R.; Johnson, D. W. Coord. Chem. Rev. 1984, 53, 1.
16. Hancock, R. D. Progr. Inorg. Chem. 1989, 37, 187.
17. Comba, P. Coord. Chem. Rev. 1993, 123, 1.
18. Deeth, R. J. Coord. Chem. Rev. 2001, 212, 11.
19. Deeth, R. J.; Foulis, D. L. Phys. Chem. Chem. Phys. 2002, 4, 4292.
20. Hancock, R. D.; Martell, A. E. Chem. Rev. 1989, 89, 1875.
Modeling Solvent Extraction of Metal Ions 711
21. Pople, J. A.; Santry, D. P.; Segal, G. A. J. Chem. Phys. 1965, 43, 129.
22. Hohenberg, P.; Kohn, W. Phys. Rev. B 1964, 136, 864.
23. Warshel, A.; Levitt, M. J. Mol. Biol. 1976, 103, 227.
24. Weiner, S. J.; Kollman, P. A.; Case, D. A.; Singh, U. C.; Ghio, C.; Alangona,
G.; Profeta, S. Jr.; Weiner, P. J. Am. Chem. Soc. 1984, 106, 765.
25. Hay, B. P. Metal Ion Separation and Preconcentration: Progress and Oppor-
tunities; Bond, A. H. Dietz, M. L. Rogers, R. D. Eds.; ACS Symposium Series
716; American Chemical Society: Washington, D.C., 1999; p. 102.
26. Sachleben, R. A.; Moyer, B. A. Metal Ion Separation and Preconcentration:
Progress and Opportunities; Bond, A. H. Dietz, M. L. Rogers, R. D. Eds.;
ACS Symposium Series 716; American Chemical Society: Washington, D.C.,
1999, p. 114.
27a. Sella, C.; Bauer, D. Solv. Extr. Ion Exch. 1992, 10, 579.
27b. Sella, C.; Bauer, D. Solv. Extr. Ion Exch. 1993, 11, 395.
28. Bond, A. H.; Chiarizia, R.; Huber, V. J.; Dietz, M. L.; Herlinger, A. W.; Hay,
B. P. Anal. Chem. 1999, 71, 2757.
29. Hay, B. P.; Hancock, R. D. Coord. Chem. Rev. 2001, 212, 61.
30. Kopczyński, T.; Łozyński, M.; Prochaska, K.; Burdzy, A.; Cierpiszewski, R.;
Szymanowski, J. Solv. Extr. Ion Exch. 1994, 12, 701.
31. Bogacki, M. B.; Jakubiak, A.; Prochaska, K.; Szymanowski, J. Colloids
Surfaces A 1996, 110, 263.
32. Bogacki, M. B.; Jakubiak, A.; Cote, G.; Szymanowski, J. Ind. Eng. Chem.
Res. 1997, 36, 838.
33. Kyuchoukov, G.; Bogacki, M. B.; Szymanowski, J. Ind. Eng. Chem. Res.
1998, 37, 4084.
34. Krueger, T.; Gloe, K.; Stephan, H.; Habermann, B.; Hollmann, K.; Weber, E.
J. Mol. Model. 1996, 2, 386.
35. Stephan, H.; Gloe, K.; Krüger, T.; Chartroux, C.; Neumann, R.; Weber, E.;
Möckel, A.; Woller, N.; Subklew, G.; Schwuger, M. J. Solv. Extr. Res. De-
velop. Japan. 1996, 3, 43.
36. Comba, P.; Gloe, K.; Inoue, K.; Krüger, T.; Stephan, H.; Yoshizuka, K.
Inorg. Chem. 1998, 37, 3310.
37. Le, Q. T. H.; Umetani, S.; Suzuki, M.; Matsui, M. J. Chem. Soc. Dalton
Trans. 1997, 643.
38. Umetani, S.; Kawase, Y.; Le, Q. T. H.; Matsui, M. Inorg. Chim. Acta. 1998,
267, 201.
39. Torgov, V.; Erenburg, S.; Bausk, N.; Stoyanov, E.; Kalchenko, V.; Varnek,
A.; Wipff, G. J. Mol. Struct., 2002, 611, 131.
40. Rabbe, C.; Madic, C.; Godard, A. Solv. Extr. Ion Exch. 1998, 16, 1091.
41. Rabbe, C.; Sella, C.; Madic, C.; Godard, A. Solv. Extr. Ion Exch. 1999, 17, 87.
42. Spjuth, L.; Liljenzin, J. O.; Hudson, M. J.; Drew, M. G. B.; Iveson, P. B.;
Madic, C. Solv. Extr. Ion Exch. 2000, 18, 1.
43. Drew, M. G. B.; Hudson, M. J.; Iveson, P. B.; Russel, M. L.; Liljenzin, J. O.;
Skalberg, M.; Spjuth, L.; Madic, C. J. Chem. Soc. Dalton Trans. 2002, 2973.
44. Coupez, B.; Boehme, C.; Wipff, G. Phys. Chem. Chem. Phys. 2002, 4, 5716.
712 Narbutt and Czerwiński
45. Ionova, G.; Ionov, S.; Rabbe, C.; Hill, C.; Madic, C.; Guillaumont, R.;
Krupa, J. C. Solv. Extr. Ion Exch. 2001, 19, 391.
46. Pyykkö, P. Chem. Rev. 1988, 88, 563.
47. Schreckenbach, G.; Hay, P. J.; Martin, R. L. J. Comput. Chem. 1999, 20, 70.
48. Schautz, F.; Flad, H.-J.; Dolg, M. Theor. Chim. Acta 1999, 99, 231.
49. Skylaris, C.-K.; Gagliardi, L.; Handy, N. C.; Ioannou, A. G.; Spencer, S.;
Willets, A.; Simper, A. M. Chem. Phys. Lett. 1998, 296, 445.
50. Willets, A.; Gagliardi, L.; Ioannou, A. G.; Skylaris, C.-K.; Simper, A. M.;
Spencer, S.; Handy, N. C. Int. Rev. Phys. Chem. 2000, 19, 327.
51. Gagliardi, L.; Handy, N. C.; Skylaris, C.-K.; Willets, A. Chem. Phys. 2000,
252, 47.
52. Spencer, S.; Gagliardi, L.; Handy, N. C.; Ioannou, A. G.; Skylaris, C.-K.;
Willets, A.; Simper, A. M. J. Phys. Chem. A 1999, 103, 1831.
53. Pershina, V.; Fricke, B.; Kratz, J. V.; Ionova, G. V. Radiochim. Acta 1994, 64,
37.
54. Pershina, V.; Fricke, B. J. Phys. Chem. 1996, 100, 8746.
55. Pershina, V. Chem. Rev. 1996, 96, 1977.
56. Cramer, C. J.; Truhlar, D. G. Chem. Rev. 1999, 99, 2161.
57. Gao, J. L. Acc. Chem. Res. 1996, 29, 289.
58. Kollman, P. Chem. Rev., 1993, 93, 2395.
59. Leach, A. R. Molecular Modeling: Principles and Applications, 2nd Ed.;
Pearson Education Limited, 2001.
60. Allen, M. P.; Tildesley, D. J. Computer Simulations of Liquids; Clarendon:
Oxford, 1987.
61. Case, D. A.; Pearlman, D. A.; Caldwell, J. C.; Cheatham, T. E. III; Ross, W. S.;
Simmerling, C. L.; Daren, T. A.; Mertz, K. M.; Stanton, R. V.; Cheng, A. L.;
Vincent, J. J.; Crowley, M.; Ferguson, D. M.; Radmer, R. J.; Seibel, G. L.;
Singh, U. C.; Weiner, P. K.; Kollman, P. A. AMBER5; University of Cali-
fornia: San Francisco, 1997.
62. Hummer, G.; Pratt, L. R.; Garcia, A. E. J. Phys. Chem. A 1998, 102, 7885.
63. Garde, S.; Hummer, G.; Paulaitis, M. E. J. Chem. Phys. 1998, 108, 1552.
64. Hummer, G.; Garde, S.; Garcia, A. E.; Paulaitis, M. E.; Pratt, L. R. J. Phys.
Chem. B 1998, 102, 10469.
65. Hummer, G.; Garde, S.; Garcia, A. E.; Pratt, L. R. Chem. Phys. 2000, 258,
349.
66. Gomez, M. A.; Pratt, L. R.; Hummer, G.; Garde, S. J. Phys. Chem. B 1999,
103, 3520.
67. Behler, J.; Price, D. W.; Drew, M. G. B. Phys. Chem. 2001, 3, 588.
68a. Benjamin, I. Chem. Rev. 1996, 96, 1449.
68b. Benjamin, I. Solvent Extraction for the 21st Century, Proc. ISEC’99; Cox, M.
Hildalgo, M. Valiente, M. Eds.; Soc. Chem. Ind.: London, 2001, 1, 3.
69. Varnek, A.; Wipff, G. J. Mol. Struct. (Theochem) 1996, 363, 67.
70a. Varnek, A.; Wipff, G. J. Comput. Chem. 1996, 17, 1520.
70b. Varnek, A.; Wipff, G. Solv. Extr. Ion Exch. 1999, 17, 1493.
71. Varnek, A.; Troxler, L.; Wipff, G. Chem. Eur. J. 1997, 3, 552.
Modeling Solvent Extraction of Metal Ions 713
72. Nazarenko, A. Y.; Baulin, V. E.; Lamb, J. D.; Volkova, T. A.; Varnek, A. A.;
Wipff, G. Solv. Extr. Ion Exch. 1999, 17, 495.
73. Szymanowski, J. Solv. Extr. Ion Exch. 2000, 18, 729.
74. Felix, V.; Matthews, S. E.; Beer, P. D.; Drew, M. G. B. Phys. Chem. Chem.
Phys. 2002, 4, 3849.
75. Baaden, M.; Berny, F.; Madic, C.; Wipff, G. J. Phys. Chem. A 2000, 104, 7659.
76. Siekierski, S. J. Radioanal. Chem. 1976, 31, 335.
77. Narbutt, J. J. Radioanal. Nucl. Chem. 1992, 163, 59.
78. Reiss, H.; Frisch, H. L.; Lebowitz, J. L. J. Chem. Phys. 1959, 31, 369.
79. Kalra, A.; Tugcu, N.; Cramer, S. M.; Garde, S. J. Phys. Chem. B 2001, 105,
6380.
80. Nemethy, G.; Scheraga, H. A. J. Chem. Phys. 1962, 36, 3382.
81. Madan, B.; Sharp, K. J. Phys. Chem. B 1997, 101, 11237.
82. Gniazdowska, E.; Narbutt, J. J. Mol. Liquids 2000, 84, 273.
83. Sithoff, D.; Ben-Tal, N.; Honig, B. J. Phys. Chem. 1996, 100, 2744.
84. Li, J.; Hawkins, G. D.; Cramer, C. J.; Truhlar, D. G. Chem. Phys. Lett. 1998,
288, 293.
85. Best, S. A.; Merz, K. M. Jr.; Reynolds, C. H. J. Phys. Chem. B 1999, 103, 714.
86. Osakai, T.; Ebina, K. J. Electroanal. Chem. 1996, 412, 1.
87. Parker, A. J.; Alexander, R. J. Am. Chem. Soc. 1968, 90, 3313.
88. Marcus, Y. Ion Properties; Dekker: New York, 1997.
89. Narbutt, J. Polish J. Chem. 1993, 67, 293.
90. Polyakov, V. R.; Czerwiński, M. Inorg. Chem. 2001, 40, 4798.
91a. Narbutt, J. J. Inorg. Nucl. Chem. 1981, 43, 3343.
91b. Narbutt, J. J. Phys. Chem 1991, 95, 3432.
92. Narbutt, J.; Bartoś, B.; Siekierski, S. Solv. Extr. Ion Exch. 1994, 12, 1001.
93. Narbutt, J.; Fuks, L. Radiochim. Acta, 1997, 78, 27.
94. Czerwiński, M.; Narbutt, J. Phys. Chem. Chem. Phys. submitted.
95. Foresman, J. B.; Keith, T. A.; Wiberg, K. B.; Snoonian, J.; Frisch, M. J. J.
Phys. Chem. 1996, 100, 16098.
96. Katsuta, S.; Imura, H.; Suzuki, N. Bull. Chem. Soc. Jpn 1991, 64, 2470.
97. Katsuta, S.; Yanagihara, H. Solv. Extr. Ion Exch. 1997, 15, 577.
98. Narbutt, J.; Czerwiński, M.; Krejzler, J. Eur. J. Inorg. Chem. 2001, 3187.
99. Narbutt, J.; Krejzler, J. Inorg. Chim. Acta 1999, 286, 175.
100. Albinsson, Y.; Mahmood, A.; Majdan, M. Rydberg, J. Radiochim. Acta 1989,
48, 49.
101. Albinsson, Y. Acta Chem. Scand. 1989, 43, 919.
102. Habenschuss, A.; Spedding, F. H. J. Chem. Phys. 1980, 73, 442.
103. Suzuki, N.; Nakamura, S. Inorg. Chim. Acta 1985, 110, 243.
104. Kohonen, T. Self-Organizing Maps; Springer Verlag: Berlin, 1995.
105. Hayki, S. Neural Networks: A Comprehensive Foundation; Prentice Hall:
Englewood Cliffs, NJ, 1994.
106. Zell, A.; Mamier, G.; Vogt, M.; Mache, N.; Hübner, R.; Döring, S.; Herr-
mann, K.-U.; Soyez, T.; Schmalzl, M.; Sommer, T.; Hatzigeorgiou, A.; Pos-
selt, D.; Schreiner, T.; Kett, B.; Clemente, G. SNNS-Stuttgart Neural
714 Narbutt and Czerwiński
Length l meter m
Mass m kilogram kg
Time t second s
Electric current I ampere A
Thermodynamic T Kelvin K
temperature
715
716 Appendix
Physical state:
Gaseous (g)
Solid (s)
Liquid (l), not recommended in this text
Aqueous [i.e., dissolved in water (aq)]
Organic [i.e., dissolved in organic solvent (org); not recommended (s)]
Water (w), to specifically refer to water molecules
D. FORMULA STRUCTURES
Structural formulas of a number of commonly used extractants, and some
typical chelate compounds.
5. b-Diketones:
(a) ketoform; (b) enol form; (c) enolate ion;
(d) acetylacetone, 2,4-pentanedione, HAA: R = R0 = CH3;
(e) trifluoroacetylacetone, HTFA: R = CH3, R0 = CF3;
(f) benzoylacetone, HBA: R = CH3, R0 = C6H5;
(g) thenoyltrifluoroacetone, HTTA: R = formula (g), R0 = CF3;
(h) bis-acetylacetonatocopper(II) complex; and
(i) dibenzoylmethane: R = R0 = C6H5.
10. N-Nitrosophenylhydroxylamine
9. N-Phenylbenzo-hydroxamic acid (ammonium salt is cupferron)
REFERENCES
1. Högfeldt, E. Stability Constants of Metal-Ion Complexes. Part A: Inorganic
ligands; IUPAC Chemical Data Series No 22; Pergamon Press: New York, 1982.
2. Freiser, H.; Nancollas, G. H. Compendium of Analytical Nomenclature. Definitive
Rules 1987; IUPAC; Blackwell Scientific Publications (www.IUPAC.com/
publications): Oxford, 1987.
3. McNaught, A. D.; Wilkinson, A. IUPAC Compendium of Chemical Terminology,
2nd Ed.; Blackwell Science, 1997.
4. IUPAC. Quantities, Units and Symbols in Physical Chemistry, 2nd Ed.; Blackwell
Science, 1993.
Index of Compounds
The key for the abbreviations contained within the index can be found on page 739.
725
726 Index of Compounds
[Americium] Bismuth
sepn from Ln, 547ff, flowsheet, 543 analysis w diphenyldithiocarbazone,
hardness, 99 568, dithiocarbamate, 573,
ionic radius, 45 thioxine, 573
oxid states, 511 extn by D2EHPA, 664
partitioning, 539 (microemulsion)
solub param w decanoic acid, 173, hardness, 99
terpyridine, 173, 182–5 ionic radius, 45
struct w triazine and pyridine adducts, stereochem, 93
182–5 Boron
Antimony boron trifluoride/HF extractant, 431
acceptor acid as SbCl5, 101 Bromine
analysis w dithiocarbamate, 573, basicity, 137
thioxine, 573 elemental, 101
extn by dibutylcarbitol/Cl, 486, extrn, 124
MIBK/Cl, 487 hardness, 99
hardness, 99 ionic radius, 45
stereochem, 93 trioctylamine salts, 159
SbCl3 distribution, 124
Arsenic Cadmium
analysis tetrachloromethane, 573, analysis, 572, w 8-(benzene-
thioxine, 573 sulfonamido)quinoline, 575,
extn by dibutylcarbitol, 486, MIBK, diethyldithiocarbamate, 572,
487, TBP, 632, phosphine oxide, D2EHPA, 581 (SLM),
632 diphenyldithiocarbazone, 568
extn from copper electrolyte, 632 cloud point extn, 586
stereochem, 93 compl w amm, 85, 96, 106, Cl, 89,
AsCl3 distrib, 124 en, 106
extrn by alkyl phosphoric acids,
Barium 179, D2EHPA from
analysis w crown ethers, 572 accumulator waste, 636ff
extrn by crown ether in SREX hardness, 99
process, 538 ionic radius, 45
hardness, 99 stereochem, 93
ionic radius, 45 Calcium
stab const TTA/TBP, 148 analysis, 572, w diethyldithiocarbamate,
stereochem, 93 572, phosphates/
water exch rate, 218 phosphonates, 590, (ISE),
Beryllium w thioxine, 572
analysis w b-diketones, 572, compl w F, 90, TTA/TBP, 148, TBP,
triphenylmethane dyes, 568 572
compl w oxalate, 165 in crud, 313, 468
extrn by TTA/MIBK, 165 extrn by D2EHPA, 474
stereochem, 93 hardness, 99
water exch rate, 218 ionic radius, 45
Index of Compounds 727
[Calcium] [Chromium]
stereochem, 93 hardness, 99
water exch rate, 218 ionic radius, 45
Californium stereochem, 93
compl w F, 104 water exch rate, 218
extrn by TTA/MIBK, 544 Cobalt
oxid states, 511 analysis w diethyldithiocarbamate,
Calomel 573, diphenyldithiocarbazone,
complexation, 68 568, 1-nitroso-2-naphthol, 573
Carbon compl w AA, 172, Alamine-336, 6,
activated, 312, 319 amm, 95, BFA, 172, BZA, 172,
Carbon dioxide Cl, 91, 461, en, 95, F, 90, FTA,
supercritical fluid extn, 444ff 172, PTA, 172, SCN, 9, 462, TFA,
solubilities in, 448 172, TOA/Cl, 156, TTA, 172
Cerium extrn by Acorga P50, 463, amine, 5,
extrn by TLA/NO3, 161, TBP/NO3, 138, 461ff, 625ff, carboxylic acid,
525, in Thorex process, 531 463, 465, CYANEX-272, 466,
hardness, 99 468, b-diketones, 464, D2EHPA,
ionic radius, 45, 494 280ff, 286, 289ff, 305, 322, 465ff,
Cesium hydroxyoximes, 464ff, LIX84I,
analysis, 572, w nitromethane, 572 467, LIX63, 465, Kelex100, 308,
extn w crown ethers (modeling), 463, PC88A, 466, TBP, 290,
690, fr radioactive waste, 541ff 465, 468, TOPO, 465,
in nitrobenzene, 77 Versatic acid, 304, 465, 468
hardness, 98, 99 extn fr scrap alloy, 625
ionic radius, 45 hardness, 99
nuclear waste management, 528 hydration, 88, 91
partitioning, 545ff ionic radius, 45
precip by phosphotungstate, 545 oxidn of diluents by, 311
water exch rate, 218 redox, 458, 464, 468
Chlorine stereochem, 93, 458
basicity, 137 synergistic extn, 465
basicity of ClO4, 137 water exch rate, 218
hardness, 99 Cobalt/Nickel
ionic radius, 45 Caron process, 467
TOA salts, 159 extrn, 458ff, by Kelex/LIX, 308,
Chromium extrn proc fr laterite ores, 466ff,
analysis w Aliquat336/D2EHPA flowsheets: Bulong, 469, Cawse,
(SLM), 582, 8-hydroxyquinoline, 470, Murrin Murrin, 471
573, MIBK, 573, nitromethane, sepn, 471ff
573 Copper
extrn by Alamine-336, 6, 627, analysis, 572, w diethyldithiocarbamate,
D2EHPA NHþ 4 , 618 572, D2EHPA, 581, (SLM),
extn from electroplating baths, 617, diphenyldithiocarbazone, 568,
surface finishing wastewater, 626 dithizone, 572
728 Index of Compounds
[Copper] Erbium
cloud point extrn, 586 cloud point extrn, 586
compl w amine, 96, amm, 95, en, ionic radius, 494
95, dithizone, 235, Europium
hydroxyoxime, 235 adducts w bromodecanoic acid,
dump leach, 329 pyridine, terphenyl, terpyridine,
extrn w AA, 177, hydroxyaryloxime, triazine, 182ff, AA, hexone,
11, IPT, 144–5, TTA, 144–5, 188, ITP, 156, quinoline, TBP,
b-diketones, 478ff, LIX54, 479, TOPO, TTA, 149, TBP/BFA,
amine/Cl, 463, carboxylic acid, /FTA, /BZA, /TFA,
473, D2EHPA, 474, hydroxy- 149
oxime, 474ff, Kelex-100, 476ff, compl w acetate, 86, b-diketones,
LIX, 474ff, pyridine-carboxylic 149, F, 90, alkylphosphoric acids,
acid ester (Acorga CLX50), 479, 179
ACORGA, 475, 8-quinolinol extrn by D2EHPA, 500, CYANEX-
reagents, 476ff 301 (modeling), 685, 702
extrn from ammoniacal etch soln, ionic radius, 494
636, brass mill flue dust, 618
extrn costs (LIX64N), 327 Fluorine
extrn eff of org solv, 77, 141–3, 146, basicity, 137
174, of temp, 177 hardness, 99
extrn plants, 477ff ionic radius, 45
hardness, 98, 99
ionic radius, 45 Gadolinium
kinetics/hydroxyoxime, 474 cloud point extrn, 586
ore, 11 ionic radius, 494
solub param, 174 Gallium
stereochem, 93 analysis, 572, w diphenyldithiocarba-
struct, 169, 721 zone, 568, di-isopropyl ether,
water exch rate, 218 572, 8-hydroxyquinoline, 572,
Curium thioxine, 57
compl w Cl, 104 cloud point extrn, 586
extrn, 531ff, by CMPO, 532ff, TTA/ compl w F, 90, OH, 166,
MIBK, 544 Cl in diethyl ether,
extn flowsheet, 543, PUREX, 520, 110
DIAMEX, 537 extrn w AA, 166
oxid states, 511 hardness, 99
partitioning, 539 ionic radius, 45
Cyanide stereochem, 93
recovery using phosphine oxides, water exch rate, 218
632 Germanium
analysis, w dibutyl ether, 572
Dysprosium distribution of GeCl4,
compl w F, 90 124
ionic radius, 494 stereochem, 93
Index of Compounds 729
Gold [Indium]
analysis, 572, w diphenyldipyridyl- stereochem, 93
methane, 572, diphenyldithio- water exch rate, 218
carbazone, 568, dithizone, 572, Iodine
polyoxyethylene nonyl phenyl in nitrobenzene, 77
ether (PONPE), 584 basicity, 137
cloud point extrn, 584, 586 elemental, 101, extrn, 124
compl, 219, w Cl, 485 hardness, 99
extrn, 485ff, by amide, 486, ionic radius, 45
amine/Cl (TOA), 485, /CN, 487, Iridium
/S2O3, 488, CYANEX-471X/Cl, analysis w diphenyldithiocarbazone, 568
485, butylcarbitol/Cl, 485, comp w Cl, 481
2-ethylhexanol/Cl, 485, guanidine extrn, 493
(LIX79)/CN, 487, MIBK/Cl, 485, extrn flowsheet INCO, 482, Lonhro,
8-quinolinol (Kelex 100, LIX26), 483, MRR, 484
486, TBP/S2O3, 488 redox, 489
extn flowsheet INCO, 482, Lonhro, stereochem, 93
483, MRR, 484 Iron
hardness, 99 analysis w AA, 572, diethyldithio-
solub of Cl compl, 9 carbamate, 573,
stereochem, 93 triphenylmethane dyes, 568
cloud point extrn, 586
Hafnium compl w amm, en, 95, TOA/Cl,
analysis w TTA, 573 157, F, 90, hydroxyoxime, 235,
extrn by MIBK, 322, TBP/NO3, 317, hydroxamic acid, 235
329 extrn by trialkylamm/Cl, 236, amine/
hardness, 99 Cl, 463, calixarenes, 519,
ionic radius, 45 dibutylcarbitol/Cl, 486,
stereochem, 93 D2EHPA, 474, MIBK/Cl, 487,
Hydrogen ion TBP/NO3,Cl, 512
hardness, 99 extrn w Cu, 473ff
ionic radius, 45 in crud, 313, 317
HCl in ethylether, 136 hardness, 99
HClO4/TBP, 136 ionic radius, 45
HF extn by TBP, 628, CYANEX-923, oxidn of CYANEX-302, 311
628, fr pickling acid baths, 628 stereochem, 93
sulfamate as reductant, 520
Indium water exch rate, 218
analysis, 572 w diphenyldithiocarba-
zone, 568, 8-hydroxyquinoline, Lanthanides
572, b-mercaptoquinoline, 572, analysis, 572, in chromatogr sepn, 59,
thioxine, 572 D2EHPA, 572
compl w F, 90, F, Cl, Br, I, 100 charge density, 168, 178
hardness, 99 compl, 179, w b-diketones, 100, OH
ionic radius, 45 150, TTA/TBP, 118
730 Index of Compounds
[Lanthanides] Magnesium
compl distr const, adducts w AA, 153, analysis, 572, w 8-hydroxyquinoline,
155, alkylphosphoric acids, 235 572
contraction, 169, 494 compl w F, 90
coord number, 100, 118, 170, 179 in crud, 313, 468
in crud, 317, 319 extrn by D2EHPA, 474
electron shells, 178 hardness, 99
extn, 493ff, w AA, 210, 215, 225, 701, ionic radius, 45
alkylphosphonic acid ester, 501, stereochem, 93
(modeling), 706, amine, 497, water exch rate, 218
b-diketones (modeling), 698, Manganese
carboxylic acid, 497, 500, analysis w tetraethylene-
CMPO, 532, CYANEX-272, dithiocarbamate, 573,
501, D2EHPA, 329, 498, 500ff, 8-hydroxyquinoline, 573
TBP/NO3, 496, tropolone, 702 compl w amm, en, 95
extrn proc DAPEX, 512, 514 extrn by Alamine-336 6, amine/Cl,
hardness, 99 463, CYANEX272/TBP, 468
ionic radius, 45, 168, 494 hardness, 99
partitioning sep from An, 539 ionic radius, 45
modeling, 691, 697, 698, 701ff, 704, 706 stereochem, 93
scale-up, 333 water exch rate, 218
stereochem, 93 Mercury
water exch rate, 218 analysis, 572, w diphenyldithiocarba-
Lanthanum zone, 568, dithizone, 572
compl w F, 90 compl w Cl, 9, 10, F, Cl, Br, I, 100,
extrn w alkyl phosphoric acids, 179, 125
D2EHPA, 218 extrn by alkyl phosphoric acids,
self-adducts w TTA, ITP, 156 179, monothiophosphoric acid,
Lead 641
analysis w diethyldithiocarbamate, hardness, 99
572, D2EHPA (SLM), 581, ionic radius, 45
diphenyldithiocarbazone, 568, stereochem, 93
dithizone, 573, thioxine, 573 HgCl2 distribution, 124
compl w F, 90 water exch rate, 218
hardness, 99 Molybdenum
ionic radius, 45 analysis w N-benzoyl-N-phenyl-
stereochem, 93 hydroxylamine, 573,
water exch rate, 218 8-hydroxyquinoline, 573
Lithium in crud, 321
analysis, 572, w dipivaloylmethane, extrn by Alamine-336, 6, w U
572, TOPO (ISE), 590 (DAPEX process), 512,
hardness, 99 514, from spent catalyst,
in org solv, 77 641
ionic radius, 45 hardness, 99
water exch rate, 218 stereochem, 93
Index of Compounds 731
Neodymium Niobium
compl w F, 90 analysis w tetraphenylarsonium
extrn by TBP/NO3, 496 chloride, 573, 8-hydroxyquino-
ionic radius, 494 line, 573
Neptunium stereochem, 93
compl w F, 90, SO4, 104, AA, 146, 169 Nitrate, NO3/Nitric acid
extn, 531ff, by CMPO, 532ff, analysis w tributyl octadecylphos-
diamides, 534ff, TBP/NO3, 525ff, phonium nitrate (ISE), 590
in partitioning, 539 compl w trioctylamine salts, 159
oxidn states, 511 basicity, 137
Nickel extrn by TBP/HNO3, 138, 628,
analysis w diethyldithiocarbamate, CYANEX-923/HNO3, 628
573, dimethylglyoxime, 573, extrn from acid pickling baths,
diphenyldithiocarbazone, 568, 628
TTA, 573
cloud point extn, 586 Osmium
compl, 219, w amm, en, 92, 95, Cl, 92, extn, 481, 485
94, 463, dithizone, 235, TOA/Cl, flowsheets INCO, 482, Lonhro, 483,
156, CN, 92 MRR, 484
extrn by Alamine-336, 6, 138, 463 hardness, 99
extrn by carboxylic acids, 463, 468: stereochem, 93
Versatic acid, 322, 468, OsO4 compl w AA, 124
naphthenic acid, 322
extrn by hydroxyoximes: ACORGA- Palladium
P50, 463, LIX84I, 467, 468, analysis w diethyldithiocarbamate,
LIX84, 623ff, 638ff 574, diphenyldithiocarbazone,
extrn by organophosphorus acids: 568, diphenyldithiourea, 574,
D2EHPA, 279ff, 289, 465, 613, 8-hydroxyquinoline, 574
615, 641, CYANEX272, 466, 471, cloud point extrn w Triton X-100,
PC88A, 466, monothiophos- 585
phoric acid, 641 compl, 219, w Cl, 480ff,
extn proc, 469–71, fr accumulator extrn, 488ff, by amine/Cl, 491,
waste, 638ff, cadmium aminoacid, 490, CYANEX-
electrolytes, 615, fly ash and 471X/Cl, 490, di-n-octylsulphide/
soot, 621, laterites, 466ff, scrap Cl, 488, hydroxyoxime,
alloy, 625, spent catalyst, 488ff, quinolinol (LIX26),
641 491
extrn rate, 463 extrn flowsheet INCO, 482, Lonhro,
hardness, 99 483, MRR, 484, fr nucl waste
ionic radius, 45 (OMEGA proj), 545
redox, 458 hardness, 99
scrubbing, 286ff stereochem, 93
sepn from Co, 458ff Phosphate PO4/Phosphoric acid
stereochem, 93, 94, 458 analysis w butanol, 573
water exch rate, 218 in crud, 317
732 Index of Compounds
[Scandium] Technetium
ionic radius, 45 analysis w tetraphenylarsonium
stereochem, 93 chloride, 573
water exch rate, 218 extrn by CMPO, 532, diamides, 535,
Selenium in PUREX proc, 527, SREX
analysis w diethyldithiocarbamate, proc, 538
573, triphenylmethane dyes, partitioning, 527, 539
568 stereochem, 93
Silica Tellurium
in crud, 313ff, 317, 319, 468 analysis w MIBK, 573
Silver extrn by dibutylcarbitol/Cl, 486,
analysis w diethyldithiocarbamate, 572, MIBK/Cl, 487
diphenyldithiocarbazone, 568, stereochem, 93
dithizone, 572 Thallium
cloud point extrn, 586 analysis w diethyldithiocarbamate,
compl w Cl, 480ff 572, 8-hydroxyquinoline, 572
extrn by CYANEX-301, 488, hardness, 98, 99
CYANEX-471X, 488 ionic radius, 45
hardness, 99 stereochem, 93
ionic radius, 45 Thorium
stereochem, 93 analysis w 8-hydroxyquinoline, 572,
Sodium TOPO, 572
in nitrobenzene, 77 extn by AA, 186ff, amines, 518,
extrn by crown ethers (modeling), TBP/NO3,Cl, 518, 528, TTA,
690 118
hardness, 99 extrn from ores, 518, fuels (THOREX
ionic radius, 45 process), 528ff, reproc, 552
water exch rate, 218 hardness, 99
Strontium ionic radius, 45
analysis w polyethylene glycol, 572 Thulium
compl w F, 90, by TTA/TBP, 148, compl w F, 90
TBP/NO3, 525 Tin
extrn w crown ether (SREX process), analysis w diisopropyl ether/SCN,
538, from nuclear waste, 537ff, 572
541–2 extrn by dibutylcarbitol/Cl, 486,
hardness, 99 TOA/Cl, 485
ionic radius, 45 hardness, 99
stereochem, 93 stereochem, 93
water exch rate, 218 Titanium
Sulfate analysis w monolaurylphosphoric
trioctylamine salts, 159 acid, 573, triphenylmethane dyes,
568
Tantalum hardness, 99
analysis w MIBK, 573 stereochem, 93
stereochem, 93 water exch rate, 218
734 Index of Compounds
Tungsten [Uranium]
analysis w MIBK, 573 oxidn states, 511
stereochem, 93 production, 11, 511
redox reactn, 260
Uranium sepn fr Pu, 520
activity coeff in ClO4, 253–275, Cl, struct w dibromosalicylate, 92,
259 water exch rate, 218
adduct compl, chelation, 114, 155
analysis w diethyldithiocarbamate, Vanadium
572, 8-hydroxyquinoline, 572, analysis w cupferron, 573,
quaternary amine, 572, TTA, 572 tetraphenylarsonium
cloud point extn, 586 chloride, 573, TTA, 573
compl w carbonate, 272, F, 90, 104, extrn by D2EHPA/TBP, 622ff,
Cl, SO4, 104, NO3, 10, AA, 189ff, monothiophosphoric acid, 641
selfadd, 156, HFA, 101, TTA, extrn fr fly ash and soot, 621, spent
116, selfadd, 156, D2EHPA, 723, catalyst, 641
NO3/TBP, inner/outer sphere, hardness, 99
180, struct, 181 stereochem, 93
composn of irrad U, 518ff water exch rate, 218
coord, 180–1
in crud, 311, 314, 316ff Xenon
dissolution, 519, 522 elemental, 101
extrn by alkylpyrophosphoric acids, extrn, 124
516, amide/NO3, 189, 322, 325,
calixarenes, 517–8, diamides, Ytterbium
178, 528, D2EHPA/TOPO, compl w F, 90
515ff, DOBA, DOiBA, 182, ionic radius, 494
organophosphate/NO3, 182, Yttrium
salicylic acid, 92, TBP, TiBP, 182, analysis w diantipyrylmethane,
TBP/NO3, 100, 113ff, 322, 513, 572
515, TBPO/NO3, 116, TLA/SO4, extrn by NO3/TBP, 496, 497
157, TTA/TBP, 116, 185, TTA/ modeling, 699
TOPO, 116, 185, NO3+TBP, hardness, 99
236, fr phosphoric acids, 319, ionic radius, 45, 494
515, fr sea-water, 517 stereochem, 93
extrn plants, 524 water exch rate, 218
extrn proc AMEX, 512, PUREX,
518ff, DAPEX, 512, Thorex, Zinc
528ff analysis w 8-(benzene-sulfonamido)-
hardness, 99 quinoline, 575,
hydration, 114 diethyldithiocarbamate, 572,
ionic radius, 45 diphenyldithiocarbazone, 568,
modeling, 690 dithizone, 572,
nucl fuels, 507 thioxine, 572
ore leaching, 511ff cloud point extrn, 586
Index of Compounds 735
Dithizone, 140, 568, 572ff, 599 LIX63, 460, 465, 474, 489
Dodecane, 427 LIX64N, 460, 474
Dyes, 568 LIX79, 488, 487
LIX84, 460, 476, 623, 638
Enzyme extrn, 442ff, 669ff LIX84I, 460, 467, 468, 475
Ethanol extrn, 438
Ethane, supercrit, 446 b-Mercaptoquinoline, 572
Ethers, 462 Mesityl oxide, 572
Ethyl acetate/acetic acid/water system, Methanol, 570
426 Methyl isobutyl ketone, (MIBK/
Ethylenediamine (en), 96, 101 hexone), 485, 487, 508, 569,
Ethylenediamine tetraacetate, (EDTA), 573ff, 723
96, 217 Methyl carbonate, 427
Ethylene carbonate, 101 4-Methyl-2-pentanone (see Methyl
2-Ethylhexanol, 281, 310, 485 isobutyl ketone)
2-Ethylhexylphosphonic acid 2- N-Methylpyrrolidone, 427
ethylhexylester (PC88A), 459, MIBK (see Methyl isobutyl ketone)
466, 501, 592 Modifiers, 310
Monolaurylphosphoric acid, 573
Fatty acids, 125–6, 593 (LLPC) Monothiophosphoric acid (see also
N-Formylmorpholine, 427 CYANEX-302), 641
The Key for the abbreviations contained within the index can be found on page 739.
741
742 Subject Index
Taft-Kamlet param, 64
Teflon phase sepn, 244 van der Waal forces, 221
Temperature eff, 176 volume, 67
Terdentate, 96 Vapor pressure relations, 49ff
Ternary systems, 70ff Virial coeff, 254, 267
Tetrahedral compl, 93, 458 Viscosity, 29, 329
Thermodynamic funct/const, 41, 46,
84–85 Water activity, 258, Ch.6
of chelation, 105, Water exchange, 214, 216,
of complexn, 102ff rate, 216
of extrn syst, 113ff Water solubilization capacity,
of interaction, 83 micelles, 662
Third phase, 288 Weber number, 381
Thorex proc (Th), 528ff Weighting, 192
Throughput (of phases), 387ff Winsor microemulsions, 661
колхоз8/1/06