You are on page 1of 22

This article was downloaded by: [USC University of Southern California]

On: 28 August 2013, At: 01:42


Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office: Mortimer House,
37-41 Mortimer Street, London W1T 3JH, UK

Solvent Extraction and Ion Exchange


Publication details, including instructions for authors and subscription information:
http://www.tandfonline.com/loi/lsei20

THERMODYNAMICS OF SOLVENT EXTRACTION


a a
G. R. Choppin & A. Morgenstern
a
Florida State University, Tallahassee, FL-32306-4390, USA
Published online: 10 May 2007.

To cite this article: G. R. Choppin & A. Morgenstern (2000) THERMODYNAMICS OF SOLVENT EXTRACTION, Solvent Extraction
and Ion Exchange, 18:6, 1029-1049, DOI: 10.1080/07366290008934721

To link to this article: http://dx.doi.org/10.1080/07366290008934721

PLEASE SCROLL DOWN FOR ARTICLE

Taylor & Francis makes every effort to ensure the accuracy of all the information (the “Content”) contained
in the publications on our platform. However, Taylor & Francis, our agents, and our licensors make no
representations or warranties whatsoever as to the accuracy, completeness, or suitability for any purpose of the
Content. Any opinions and views expressed in this publication are the opinions and views of the authors, and
are not the views of or endorsed by Taylor & Francis. The accuracy of the Content should not be relied upon and
should be independently verified with primary sources of information. Taylor and Francis shall not be liable for
any losses, actions, claims, proceedings, demands, costs, expenses, damages, and other liabilities whatsoever
or howsoever caused arising directly or indirectly in connection with, in relation to or arising out of the use of
the Content.

This article may be used for research, teaching, and private study purposes. Any substantial or systematic
reproduction, redistribution, reselling, loan, sub-licensing, systematic supply, or distribution in any
form to anyone is expressly forbidden. Terms & Conditions of access and use can be found at http://
www.tandfonline.com/page/terms-and-conditions
SOLVENT EXTRACTION AND IONEXCHANGE, 18(6), 1029-1049(2000)

THERMODYNAMICS OF SOLVENT EXTRACTION


Downloaded by [USC University of Southern California] at 01:42 28 August 2013

G. R. Choppin and A. Morgenstern


Florida State University, Tallahassee, FL-32306-4390, USA

ABSTRACT
The fundamental thermodynamic characteristics of aqueous electrolyte solutions
and organic solutions which affect solvent extraction are summarized and the influence of
metal complexation and hydration on the distribution ratios is discussed. The
thermodynamics of extraction systems, including synergistic systems, is reviewed. The
influence of structural aspects of the complexing extractant agents on these
thermodynamic parameters is also reviewed.

THERMODYNAMICS OF AOUEOUS ELECTROLYTE SOLUTIONS AND


ORGANIC SOLUTIONS

In solvent extraction systems, the interaction of the extracted solute with both
aqueous and organic solvent molecules plays a significant role in the distribution of the
solute between the phases. Thus, an understanding of the physico-chemical properties of
aqueous electrolyte and of organic solutions as they determine the role of the interactions
with the solute is necessary for successful design of solvent extraction systems. An
extensive review of the thermodynamics of aqueous and organic solutions is beyond the
scope of this paper and can be found in textbooks on physical chemistry and solvent

1029

Copyright IJ:> 2000 by Marcel Dekker, Inc. www.dekker.com


1030 CHOPPINAND MORGENSTERN

extraction [1-4). Here we restrict the discussion to the fundamental concepts of molecular
interactions in aqueous and organic solutions.
Commonly in solvent extraction systems, one of the phases between which the
solute distributes is an aqueous solution that contains one or more electrolytes The ions
of the solute are likely to have associated waters of hydration, in the aqueous phases and
Downloaded by [USC University of Southern California] at 01:42 28 August 2013

associated solvent molecules in the non-aqueous phase. At sufficiently high


concentrations, the ions can interact with one another, repulsively if of the same charge
sign, attractively if of opposite sign. A fundamental principle of electrolyte solutions is
that of the electroneutrality of the solution. Another useful concept, related to the
activities of electrolytes, is that of the ionic strength, I, of the solution where
I = 1/2 ~ CiZ? (the summation extends over the concentration and charge of all cations
and anions present in the solution). Debye and Huckel (5) used electrostatic and statistical
mechanical theories to obtain an equation for calculation of the mean ionic activity
coefficient of a dilute electrolyte solution. When two or more electrolytes are present in
the same solution, and one is at a significantly higher concentration, the activity
coefficient of the secondary electrolytes are a function of the high concentration
component. This is the basis of the ionic medium method in which an electrolyte (e.g.
sodium perchlorate) is present at a fixed, high concentration (e.g., ;>: 1 mol L") which
allows variation of lower concentrations of reactive electrolytes within certain limits
without significant change in their activity coefficients [1).
Solutions in organic solvents or in mixed aqueous-organic solvents may behave
similarly to purely aqueous solutions. In particular, if the relative permittivity (E) of the
solvent is greater than about 40, electrolytes at lower concentrations are, more or less,
completely dissociated into ions. However, for solvents of E < 10, ionic dissociation is
insignificant and the behavior is likely to differ significantly from aqueous systems. It
must also be noted that mixed solutions of aqueous-organic nature usually have major
disruption of the three-dimensional, cooperative hydrogen-bonded network that
characterizes the water structures in aqueous solutions. Many "anomalous" properties of
aqueous solutions that depend on this structured nature of water are absent in organic
solvents.
The presence of an organic solvent, even at low concentrations, in a solvent mixture
affects the activity coefficient of an electrolyte relative to its value in water.
THERMODYNAMICS OF SOLVENT EXTRACTION 1031

Accordingly, in mixed aqueous-organic solvents there is a primary medium effect on the


activity coefficient which reflects the interactions of the ions with their mixed-solvent
surroundings as it differs from their interactions with a purely aqueous environment.
Anion - cation attraction to form "ion-pairs" occurs readily in organic and mixed
aqueous-organic solvents. It also occurs in aqueous solutions, notably with higher
valence type electrolytes at higher concentrations. Ion pairing, generally, is weaker than
Downloaded by [USC University of Southern California] at 01:42 28 August 2013

normal complex formation involving covalent bonding between metal cations and anionic
ligands. A useful concept to describe the association of a cation and an anion to form an
ion pair is Bjerrum's theory [I]. Electrostatic interactions also must include ion-dipole
and dipole-dipole bonding. Dipole-dipole interactions between polar solute particles in
organic solvents may be either repulsive "head-head" interactions or attractive "head-tail
interactions. Such attractive dipole interactions can lead to solute aggregation, of dimers,
(in which the "head" of each of the two partners is near the "tail" of the other), and of
chainlike and cyclic aggregates (oligomers). Tertiary amine salts in hydrocarbon solvents
are typical examples of such aggregated solutes. Hydrogen bonding between solute
particles also leads to aggregation. Typical of such solutes are carboxylic acids and acidic
phosphate esters and cyclic dimers. Noncyclic aggregates can form which result in an
increase in the viscosity of the solution with increasing concentrations. Another kind of
solute-solute interaction, donor-acceptor adduct formation, forms I: I species between
molecules of two different kinds of solute. Adduct formation results when one partner
(the donor) has a pair of unbonded electrons, (e.g. the nitrogen atom in trioctylamine),
and the acceptor has an empty orbital that can accept the pair of electrons. These effects
play varying roles in the extraction of the solute between the aqueous and the organic
solvent phases

METAL ION COMPLEXATION AND HYDRATION

The complexation of metal ions can have an important influence on the relative
affinity of different metals for the solvent phases and can provide a sufficient difference
in extractability to allow separation of the metals. In this section, the thermodynamics of
metal complexation relevant to solvent extraction are described briefly with attention to
1032 CHOPPIN ANDMORGENSTERN

the parameters influencing complexation and to models of complex formation. For more
extensive discussions, references 1- 4 are recommended.
For any metal-ligand system, the equilibrium constant (i.e., the stability constant)
is a quantitative measure of the metal ion complexation, and is expressed as the stepwise
reaction constants K, for the reaction MLn.I+L=MLn, or as the overall constant, Bn, for
,
Downloaded by [USC University of Southern California] at 01:42 28 August 2013

the net reaction M+nL=ML, (~, = n K,). These overall constants are also written with
..I

several subscripts; e.g., ~pq, where p=number of metal ions, q=number of protons (or, if
negative, to the number of OH) and r=number of ligands. The ratio of complexed and
free species is expressed by the equation:
[ML,l = f3 . [L)' (I)
[M] n

Rigorously, a thermodynamic stability constant is defined in terms of standard


state conditions where the constant is defined by the activities of the different species.
However, the conditions of measurements of ions and complexes provide their
concentrations are related by the activity coefficients. Activity coefficients at ionic
strengths below ca. 2 mol kg" can generally be modeled with acceptable accuracy using
the SIT approach [6], while systems at higher ionic strengths are better treated by the
Pitzer model [7].

FACTORS IN STABILITY CONSTANTS

Many factors, among them statistical, electronic, geometric (bonding and steric),
chelation, and the nature of the metal and ligands are reviewed for their effects on the
stability constrants.

Statistical Effects
In complexation, ligands displace hydrate waters, although not necessarily on a
I: I basis. Charge, steric, etc. effects determine the number of displaced waters and of
bonded ligands. For example, Co" is an octahedral hexahydrate, Co(H 20)62+, but with
chloride forms complexes with 1-4 Cl' anions (CoCl;' is tetrahedral). Trivalent actinides

have large hydration numbers - usually 8 or 9 - but may form complexes with N<8 with
THERMODYNAMICS OF SOLVENT EXTRACTION 1033

bulky anionic ligands. In his treatment of the statistical effect, Bjerrum assumed that the
total coordination, N, remained constant as hydration was replaced by complexation. If a
ligand with a single binding site approaches the metal, its probabilitiy of interacting is
proportional to the number of available bonding sites on the metal, N. The probability of
dissociating the complex is proportional to the number of ligands present; i.e., I for ML.
Therefore, statistically the stepwise stability constant, K, should be proportional to the
Downloaded by [USC University of Southern California] at 01:42 28 August 2013

ratio of the probabilities ofML formation and ofML dissociation; i.e., K, oe Nil [2].
Similarly, for the ML, formation by ML + L, only N-I sites are avialable on the
M since one site is already occupied but the probability of dissociation of L from ML, is
proportional to 2, i.e., K, ee (N-l )/2. In general, for MLn.1+L=ML n
Kn~ [N-Cn-I)]/[n] (2)
and the ratio of successive stability constants would be (assuming other factors play no
role):
Kn+1 =:: n N-n
(3)
K, N-(n+l) n+1
Thus, for a hypothetical systems in which ~,=IOO and N=6, we can calculate K,0"42,
(~,=4.2 x \03) while K3=21, C~3=8.8 x \04), etc. The experimental (8) values reported for
cadmium complexation by chloride are ~1=\OO, ~,=400, ~3=50. Obviously, factors other
than the statistical effect reduce K" K3, etc., in the Cd + Cl complexation system.
Nevertheless, such statistical calculations can be used for an upper limit on the Kn+,lK n
ratio and the difference in the calculated values and the experimental values can offer a
measure of the ligand and steric effects which lower the constants.

Electrostatic Effect
The Born equation, which describes the electrostatic interactions between an
anion and a cation in a solvent, has the form:
Ll.G.1=-A'Iz,z./Er (4)
where A'l is a constant, z, and z, are the cationic and anionic charges, E is the dielectric
constant and r is the distance between the charge centers (i.e., the sum of the radii of the
cation and anionic bonding group). Ionic radii for a number of cations and anions are
listed in [8]. Although Equation (4) has the theoretical form to calculate Ll.G,\, the proper
values of E are uncertain and empirical values of E or of E/r are obtained from experi-
mental values of Ll.G'1 for use in calculation of related systems.
1034 CHOPPIN AND MORGENSTERN

cE.. 4.00
Cl
Downloaded by [USC University of Southern California] at 01:42 28 August 2013

.2

2.00

0.00 ""_-'-..L>*'*,,-_'----_.-"';-------'_----,."""--'
0.00

Figure 1. Correlation of log 13 of fluoride complexation vs. ~r, of the metal.

Equation (4) indicates that for complexation systems with ligands in which the
bonding is strongly electrostatic and steric effects are very similar, the stability constant
should be related to the charge of the cation divided by r. For a particular ligand (constant
anionic radius), log 13 could be expected to correlate with ~r, (r,= cationic radius).
Figure I shows such a correlation for a number of metals complexcd with the fluoride
anion; the linearity of this correlation confirms the dominance of electrostatic bonding in
these I: I flouride complexes.
Another useful correlation can be derived from Eq. (4). Proton association with
ligands is electrostatic so the equation should be applicable to proton association.
Moreover, if both HL and ML are ionic, log 131 should correlate with log K. (or, with
pKa), assuming no structural changes in the metal complex. Figure 2 shows the excellent
correlation found for log 131 for samarium (1II) with pKaof a series of monocarboxylate
ligands. Such correlations can be very useful in estimating the role of unknown stability
constants of new complexing/extractant ligands being considered for use in solvent
extraction separations.
THERMODYNAMICSOF SOLVENT EXTRACTION 1035

di 2 .O
Downloaded by [USC University of Southern California] at 01:42 28 August 2013

Dl
..2

pKa

Figure 2. Relationship between the stability constant, ~IO" for formation of SmL2+ and
the acid constant, pK", of HL: (I) propionic; (2) acetic acid; (3) iodoacetic; (4)
chloroacetic acid; (5) benzoic acid, (6) 4-flurobenzoic acid; (7) 3-flurobenzoic acid; (8)
3-nitrobenzoic acid.

Geometric Effects
The geometry of the complex can playa significant role in the strength of many
complexes. The relative sizes of the cations and anions is very important in determining
the geometrical pattern in ionic crystals and, for such cation-anion pairs, steric effects can
be expected to be found in their complexes. In Table 1, the coordination number and the
geometric pattern are listed for various ratios of the radius of the metal cation to that of
the anionic ligand. Co'+ and H,O have a radius ratio of about 0.5 and, as predicted
octahedral Co(H,O)6'+ is the hydrated species; Co'+ and Cl' have a radius radio close to
0.3 and, as predicted, the CoC!.- complex has a tetrahedral structure.
Factors important in the geometry of metal complexes are: (a) arrangement of the
ligands about the metal to minimize electrostatic repulsions (predominant in ionic
complexes) and (b) overlap of the metal and ligand orbitals (important in covalent
compounds). The first requirement (a) favors a tetrahedral configuration for CN=4 as the
1036 CHOPPIN AND MORGENSTERN

ligands are father apart than in the square planar geometry. However, if overlap of
orbitals is a stronger requirement, and a d orbital can be included in the hybridization, the
dsp2square planar geometry gives the more stable complex. In the octahedral complexes
of CN=6, secondary structural effects can be observed that can be attributed to
differences in ligand field effects related to the electron distribution among the metal d
Downloaded by [USC University of Southern California] at 01:42 28 August 2013

orbitals.

The Chelate Effect


This effect involves binding of the metal to more than a single donor site of the
ligand. Ethylenediamine, H2N-C2H.-NH2, is a bidentate ligand which binds through the
two nitrogen atoms. Ethylenediaminetetraacetate, EDT A, is a hexadentate ligand
[(OOC·CH2h NC2H.N(CH2·COO)2]'", binding through the 2 nitrogens and an oxygen of
each of the 4 carboxylate groups. Chelates are commonly stronger. than analogous non-
chelate complexes. Ammonia and ethylenediamine both bind via nitrogen atoms and log
~2 for M(NH3h 2+ can be compared with log ~, for M(en)2+ while log ~. for M(NH3)P
and log ~2 for M(enh 2+ can be similarly compared to ascertain the stabilizing effect of
chelate formation. Cu(en) 2+ is more stable than CU(NH 3)22+ by about 2 units of log ~,
Cu(enh 2+ is almost 6 units more stable than Cu(NH 3).'+.
Chelate complexes are used in many solvent extraction systems. In such systems,
the chelating ligands are organic compounds which provide solubility in the organic
phase. The metal binds to the polar charged donor sites, leaving an outer organic structure
about the metal which favors solubility in an organic solvent, and, hence, extraction from
the aqueous phase.

MODELS OF COMPLEX FORMAnON

A major advance in developing a theory of how and why metal ions form
complexes was made by N. V. Sidgwick in 1927 who proposed that the number of
ligands that bond to a metal ion was determined by the number of electron pairs (one per
bound ligand donor site) accepted by the metal ion to achieve a stable inert gas electronic
configuration. The metal is the electron pair "acceptor" and the ligand, the electron pair
THERMODYNAMICS OF SOLVENT EXTRACTION 1037

Table I
Cation/Anion Radius Ratio Values for Cation Coordination Numbers

Coordination number Geometric pattern Radius ratio (rc/ra)

2 Linear SO.15
Downloaded by [USC University of Southern California] at 01:42 28 August 2013

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

"donor". A few years earlier, G. N. Lewis had introduced a generalized acid-base theory
in which a base was a substance which furnished a pair of electrons for a chemical bond
and an acid was a substance accepted such the electron pair. In this model, complexation
is a class of acid-base reactions. This relationship was developed further in the next
model described.

Hard/Soft Acids and Bases (HSAB)


The strength of interaction of metal and a ligand is predicted qualitatively by the
hard-soft, acid-base principle [9]. In general, hard acids are cations which favor ionic
bonding and their log ~n values correlate linearly with the pK. of the ligand acid. By
contrast, soft acid cations favor covalent bonds and their log ~ values correlate with the
redox potential, EO, or the ionization potential of the ligand. Ligands which are hard bases
tend to have higher pK, values in their acid form while ligands which are soft bases have
large EO or IP values. The important principle of the HSAB model is that hard acids react
strongly with hard bases and soft acids with soft bases. There are exceptions to this
principle due to factors more important in the interaction than the inherent acidity and/or
1038 CHOPPIN AND MORGENSTERN

basicity. Nevertheless, the simple HSAB principle has proven a very useful model for a
large variety of complexation reactions.
The bond in soft acid-soft base complexes results from sharing an electron pair,
and soft species generally have large polarizabilities. In hard (metal) acids, the energy
difference between the acceptor and donor orbital levels is so large that they do not share
the electron pair, and the bonding is strongly electrostatic.
Downloaded by [USC University of Southern California] at 01:42 28 August 2013

The characteristic features of these species are as follows:


I. hard species are difficult to oxidize (bases) or reduce (acids), and have low polariz-
abilities, small radii, higher oxidation states (acids), high pK. (bases), more positive
(for the acids) or more negative (for the bases) electronegativities, and high charge
densities at acceptor (acid) or donor (base) sites;
ii. soft species are relatively easy to oxidize (bases), or reduce (acids), and have high
polarizability, large radii, small differences in electronegativities between the acceptor
and donor atoms, low charge densities at acceptor and donor sites, and also often have
low-lying empty orbitals (bases), and 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. Thus, Cu+ and rr' are soft acids; Cu" is borderline and
Tl 3+ is hard. Even though Cs" has a large radius and low charge density, its low
ionization potential makes it a hard acid. This illustrates that the properties listed above
may not be possessed by all hard (or soft) species, but indicate the characteristics useful
to predict the acid-base nature of cations and anions.

Qualitative Use of Acid-Base Model


A number of species are listed in the hard, soft, or borderline categories in Table 2
which can be used to predict the strength of complexation. For example, Pu4+ is a hard
acid, F-, a hard base and r, a soft base. This leads to the prediction that log ~1(PuF3'J
would be larger than log ~1(PuI3+); the experimental values are 6.8 and <1.0,
respectively. By contrast, since Cd2+ is a soft acid, log ~1(CdF+) could be expected to be
smaller than log ~1(CdI). The respective values are 0.46 and 1.89 (8). For the borderline
metals such as Fe'+, ce", Ni'+, Cu", and Zn" the complexing trends are less easily
predicted. For F-, a hard base, the order oflog ~I is:
Cu> Zn- Fe> Mn> Ni >Co
TIlERMODYNAMICS OF SOLVENT EXTRACTION 1039

Table 2
List of Some Hard and Soft Acids and Bases

A. Acids

1. Hard. +1 ions H, Li to Cs
+2 ions Mg to Ba, Fe, Co, Mn
Downloaded by [USC University of Southern California] at 01:42 28 August 2013

+3 ions Fe, Cr, Ga, In, Sc, Y, Ln, An


+4 ions Ti, z-, Hf, Ln, An
-yl ions VO'+, MoO" AnO" Mn(VIl)O,"

ii. Borderline. +2 ions Fe, Co, Ni, Cu, In, Sn, Pb


+3 ions Sb, Bi, Rh, Ir, Ru, Os

iii. Soft. Neutral BH)


+1 ions Cu, Ag, Au, Hg, CH)Hg,
+2 ions Cd, Hg, Pd, Pt

B. Bases

i. Hard. Neutral H,O, ROH, NH), RNH" N,H., R,O, R)PO, (RO»)PO
-1 ions OH, RO, RCO" NO), CIO., F, CI
-2 ions 0, R(CO,)" CO), SO.
-3 ions PO.

ii. Borderline. Neutral C6H,NH" C,H,N


-I ions N), NO" Br
-2 ions SO)

111. Soft. Neutral C,H" C6Ho, CO, R)P, (RO»)P, R)As, R,S
-1 ions H, CN, SCN, RS, I
-2 ions S,O)

whereas for SCN', a soft base, it is:


Cu» Ni > Co > Fe > In> Mn
In solvent extraction, the HSAB principle can be used to indicate ligands which
react strongly with metal ions to form extractable complexes. For example, the actinide
elements would be predicted to complex strongly with the ~-diketonates, R)C-CO-CH,·
CO-CR) (where R ; H or an organic group) as the bonding is through the oxygens (hard
base) of the enolate anion. The formation of a chelate structure by the metal-enolate
1040 CHOPPIN AND MORGENSTERN

complex results in a relatively strong complex which combines with the hydrophobic
nature of the R groups on the ~.diketonates to produce high solubility in organic solvents.

Coordination Numbers
The solvent extraction of the ~-diketonate system can be used to illustrate the
Downloaded by [USC University of Southern California] at 01:42 28 August 2013

importance of the high coordination numbers of the actinides and lanthanides which
allows coordination of neutral hydrophobic adducts. The ~-diketonate ligands balance
the cation ionic charge while occupying 2n coordination sites of the metal cations. Since
the trivalent f-element cations have total coordination numbers of 8 or 9, the metal in the
neutral MLl has residual hydration. Alkyl phosphates, (RO)l PO, can coordinate through
an oxygen donor by replacing these water molecules to form MLm·Sn (n=1 to 3). This
MLm·S n species is more hydrophobic than the more hydrated MLm·(H20)n and, thus,
more soluble in an organic solvent. In the Purex process for processing irradiated nuclear
fuel, uranium and plutonium are extracted from nitric acid solution into a kerosene
solution of tributyl phosphate, TBP, as U02(NOl)dTBP)2 and Pu(NOlk(TBPh. U in
UO/+ usually has a maximum coordination number of 6 while that of Pu'+ can be 8 to
10. The addition of two TBP adduct molecules in each molecule (indicating that NOl· is
bidentate) causes the compound to be soluble (and, hence, extractable) in the kerosene
solvent.

THERMODYNAMICS OF COMPLEXATION

The standard free energy of a complexation reaction, is defined by:


LlGo" = -RT In ~n (5)
where ~n is the stability constant for the complexation. The enthalpy of complexation
LlH n can be measured directly by reacting the metal and ligand in a calorimeter or,
indirectly, by measuring log ~n at different temperatures and applying the equation:
d In ~n /dT = -LlHn/R (6)
The temperature variation method is used often in solvent extraction studies. It
can give reliable values of the enthalpy over the temperature range for which the graph of
ln~n vs I IT is linear.
THERMODYNAMICS OF SOLVENT EXTRACTION 1041

Enthalpy-Entropy Compensation
Interaction of hard cations and hard (0, N donor) ligands to form complexes are
often characterized by positive values of both the enthalpy and entropy changes. In such
cases, if T~So >~Ho, (~Go = ~Ho - T~So), log ~o is positive. These "entropy driven"
reactions are due to a decrease in the hydration of the ions, which increases the
Downloaded by [USC University of Southern California] at 01:42 28 August 2013

randomness of the system, resulting in a positive entropy contribution. Such dehydration


results in an endothermic enthalpy contribution (~Hh > 0) since the metal-water bonds of
the hydrated species must be broken. The interaction between the cation and the ligand
provides a negative enthalpy contribution (~H, < 0) due to formation of the cation-anion
bonds. This bonding decreases the randomness of the system, resulting in negative
entropy contribution (~S, < 0). The observed overall enthalpy reflects the sum of the
contributions of dehydration and cation-ligand combination. Positive values of ~Ho and
~So imply that the dehydration is more significant in these terms than the combination
step.
Many hard-hard complexation systems, have a linear correlation between the
experimental ~H and ~S values (Figure 3), which has been termed the compensation
effect. In the compensation effect, the positive values of ~Ho (=~H,+~Hh) and ~So (=~S,

+~Sh) mean I~Hhl>I~HoI and I~Shl>I~S, Iand ~Go = ~G,. This suggest the following:

a. the free energy change of the total complexation reaction, ~Go, is related principally to
the combination subreaction;
b. the enthalpy and entropy changes of the total complexation reaction, ~Ho and ~ So,

reflect, primarily, the dehydration subreaction.


These trends are important in solvent extraction systems as they provide insight into the
aqueous phase complexation and, also, have significance for the organic phase reactions.
In organic solvents, solvation generally is weaker than for aqueous solutions. As a
consequence, the desolvation analogous to step I would result in small values of
~H(solv) and ~S(solv). Therefore, ~H, ;:: ~Hsolv and ~S, ;:: ~Ssolv which ineans ~H is
more often negative while ~So may be positive or negative and relatively small.

Thermodynamics of Chelation
The positive entropy change observed in many complexation reactions has been
related to the release of a larger number of water molecules than the number of bound
1042 CHOPPIN AND MORGENSTERN

40 3/

~
30
1 2' ~ CfF+2
CmF+2
AmF+2
;:' 20 691 4 U0 2 F+ 1
10 8 5 PuCI+2
E 14 6 U02 C1+1
• 10 11
Downloaded by [USC University of Southern California] at 01:42 28 August 2013

7 PUN0 3+3
~
~ 16 8 NpS04+2
:I: 0 9 AmS04+1
<I 10 CmS04+ 1
5 11 U02S04
-10 US04+2
12
13 AmSCN+2
-20 14 Am (Glycinate)+2
15 Am (Acetate)+2
16 Am (Glycolate)+2

Figure 3. Correlation of ~H and ~s of formation of a series of hard-hard, J: I complexes


at different ionic strengths.

ligands. As a result the total degrees of freedom of the system are increased by
complexation and results in a positive value of the entropy change. In many systems, a
similar explanation can be given for the enhanced stability of ehelates. Consider the
reaction of Cd(II) with ammonia and with ethylenediamine to form Cd(NH 3)l+ and
Cd(en),'+ whose structures are given in Figure 4. Table 3 gives the values of log K, ~H
and T~S for the reactions:
Cd(NH3),2+ + en ; Cd(en)2+ + 2NH3 (14a)
Cd(NH3)l+ + Zen > Cd(en)/+ + 4NH3 (I4b)
The enthalpy value of Eq.(14a) is very small as might be expected if two Cd-N bonds
in Cd(NH3),'+ are replaced by two Cd-N bonds in Cd(en)'+. The favorable equilibrium
constants for the two reactions are due to the positive entropy change. Note that in the
reaction of Eq. (l4a), two reactant molecules form three product molecules so chelation
increases the net disorder (i.e., increased the degrees of freedom) of the system which
THERMODYNAMICS OF SOLVENT EXTRACTION 1043

H2 H2
H2C-N, /N-CH2
I Cd I
H2C_N/ 'N-C H2
H2 H2
Cd (NH2C2H4NH2)22+

Figure 4. Structures ofCd(NHJ)l+ and Cd(en)/+.


Downloaded by [USC University of Southern California] at 01:42 28 August 2013

Table 3
Thermodynamic Parameters of Reaction of Cadmiurn(ll)-Ammonia Complex with
Ethylenediamine

Complex n LogK: Mi" (klmor l ) ss: (J·m·K· 1


)

Z
Cd(ent 0.9 +0.4 5.4

Cd(en):z 2 2.2 -3.4 + 15

• K = [Cd(en);z][NH,]z,
n [Cd(NH,);~][en]'

contributes a positive 6.S change). Such chelation effects playa role in many extraction
reactions in which the extractant ligand forms chelate bonds to the metal.

THERMODYNAMICS OF SYNERGISM

Synergism is an important factor in the degree of extraction and of separation


factors in many extraction systems (10). The major factor in synergism is an increase in
hydrophobic character of the extracted metal complex upon addition of the adduct. Three
mechanisms have been proposed to explain the synergism for metal + chelant + adduct.
The first involves an opening of one or more of the chelate rings and occupation by the
1044 CHOPPIN AND MORGENSTERN

adduct molecule(s) of the vacated metal coordination site(s). In a second mechanism, the
metal ion is not coordinately saturated by the ligand and, hence, retains residual water(s)
in the coordination sphere which can be replaced by adduct molecules. The third
mechanism involves expansion of the coordination sphere of the metal ion to allow
bonding of adduct molecules. From the extraction constants, it is rarely possible to
choose between these alternative mechanisms but enthalpy and entropy data of the
Downloaded by [USC University of Southern California] at 01:42 28 August 2013

synergistic reactions can be used to provide more definitive arguments.


The overall extraction reaction for metal extraction by HITA and TBP is written
as:
Mn\aq) + n HITA(org) + pTBP(org) = M(ITA)n (TBP)p(org) + pH+(aq) (15)
The extraction equation for TTA is,
Mn+ + n HITA(org) = M(ITAln(org) + nll''(aq) (16)
The adduct formation reaction in the organic phase (the "synergistic reaction") is
obtained by subtracting Eq. (16) from Eq. (15):
M(TTAln(org) + pTBP(org) = M(ITAln(TBP)p(org) (17)
Direct calorimetric measurements of the reaction to form U02·(ITA)n·TBP(org), gave
log K = 5.10, llH = -9.3 kl-mol', TIlS = 20.0kJ·mor'·K"'. The formation of
Th(ITA).(org) + TBP(org) = Th(ITA).-TBP(org) gave values oflog K =4.94, llH=

-14.4 kl-mol", TllS= 13.7 krmol'. It was ascertained that both U02(ITAh and
Th(TTA). have two molecules of hydrate water when extracted in the benzene and these
are released when TBP is added. Since two reactant molecules (e.g., U02(ITAh'2H20
and TBP) formed three product molecules (e.g., U02(ITAh·TBP and 2H20), llS is
positive. TBP is more basic than H20 and forms stronger adduct bonds resulting in, the
enthalpy being exothermic. Hence, both the enthalpy and entropy changes favor the
reaction, resulting in relatively large values oflog K.
These equations do not provide complete definition of the reactions which may be
of significance in a particular system. For example, HITA can exist as a keto, an enol,
and a keto-hydrate species. The metal combines with the enol form which usually is
dominant in "organic" solvents (e.g. K=[HTTAle/[HTTAlk -6 in wet benzene). The

kinetics of the keto - enol reaction are not fast, but, apparently, are catalyzed by the
presence of reagents such as TBP or TOPO. Such reagents react with the enol form in
drier solvents but cannot compete with water in wetter ones. HTTATBP and TBP'H20
THERMODYNAMICS OF SOLVENT EXTRACTION 1045

species also form readily. However, for extractions into the same solvent (e.g., benzene),
these effects and need not be considered in a simple analysis of comparative extractions.
For trivalent lanthanides and actinides, the data suggest a reaction in which
addition of TBP displaces some or all of the hydrate molecules;
An(TIA)3(H,O)3 ~An(TIA)3(TBP)(H,O)1-2~An(TIA)3(TBP)l-3 (I 8)
Downloaded by [USC University of Southern California] at 01:42 28 August 2013

This scheme of steps reflects the ability of the trivalent actinides and lanthanides to vary
their coordination number in different species.
Th(TIA). can be dissolved in dry benzene with no residual hydration. Values for
the formation of Th(TTA)•. TBP from Th(TT A). + TBP in this dry system are log K =
5.46, 6W = -39.2 kJ mol", T6S o = -8.0 kJ mol" x'. The negative entropy and enthalpy
values reflect a decrease in the net degrees of freedom (2 reactant molecules combine to
form I product molecule).

CORRELAnON BETWEEN &. AND EXTRACTANT STRUCTURE

In the hard-soft model of complexation reactions, hard-hard and soft-soft


combinations lead to larger equilibrium constants than for hard-soft interactions. Since
the lanthanides belong to the hard acids and oxygen to the hard bases, the HSAB theory
helps to explain the data in Table 4 which shows K.x for the extraction reaction
M"+ (aq) + n HR(org) ~ MR,,(org) + n H+(aq) (19)
for M2+ and M 3+ cations. The increasing charge density from LaJ+ to Lu3+ (due to
decreasing ionic radii) results in increasing stability constants for formation of LnR 3.
Since K,x is proportional to the stability constants, K,x should also increase. Further,
since the "hardness" of the extractant increases from RPS,H to RPO,H (soft S being
replaced by hard 0) 133 (and, hence, K.x) also increases in this order. These trends may be
compared with the data for three soft metals, with the same extractants in Table 4. For
these soft-hard systems the extraction constant decreases with increase in charge density;
and, also with increasing hardness of the ligand which is in agreement with the HSAB
theory.
The complexation of UO/+ with diamide extractants can be used to show the
effect of ring size in chelation. TBOA (C.H9)zNCO" TBMA [(C.H9),NCO,],CH"
TBSA [(C.H9)2NC02],(CH2)z form complexes with UO;' of 5,6 and 7-membered rings,
t046 CHOPPINAND MORGENSTERN

Table 4
Comparison of the Extraction Constants K" for Metal Cations and Sulfur or Oxygen Dialkyl
Phosphoric Acids

Zn2+ Cd 2+ Hg2+ La'+ Eu'+ LuJ+


Downloaded by [USC University of Southern California] at 01:42 28 August 2013

Charge density Z; /r, 5.4 4.1 3.6 7.83 8.74 9.68

log K" for R,'POOH' -1.20 -1.80 -2.20 -2.52 -0.44 2.9

log K" for R"POSHb 0.70 3.70 5.40 -4.78 -4.23 0.34

log K" for R2'PSSH' 2.40 3.49 4.40 -8.28

R' ~ C4H,O; R" ~ C4H,CH(C2HS)CH2O


a. R ~ R' for M2+, ~ R" for MJ+
b. R ~ R' for M2+, ~ R" for M'+
c. R~ R" for M2+ and M'+

respectively. The data {(log 1(", ~ 0.005 (TBOA), 10.6 (TBMA), 9.3 (TBSA)) show that
the best extraction (i.e., largest 1(", value) is obtained for the 6-membered ring.
Steric hindrance plays an important role in the interaction with the metal. The
extraction constants log I("x for U022+ with the organophosphorous extractants, TBP and
the more branched TiBP are 28 and 26, respectively. The relatively small difference
indicates that the branching in TiBP has only a small effect because of the free rotation of
the substituents around the phosphorous atom. For the amide, DOBA, and its branched
isomer, DOiBA,log K,,~ 5.75 and 0.55, respectively. In these ligands, the nature ofR,
R' and R" in R"R'NCOR is important due to the molecular rigidity of the amide group.

SYSTEMATICS OF ADDUCT FORMATION CONSTANTS

Many common neutral extractants are able to replace solvated water at the metal
complex, e.g.:
THERMODYNAMICS OF SOLVENT EXTRACTION 1047

Table 5
Effect of Charge Density on the Adduct Formation Constants

M ZJr, Ligand Adduct Solvent log K. d1 log K. d2

Ca(II) 3.4 ITA TBP CCl, 4.11 8.22


Downloaded by [USC University of Southern California] at 01:42 28 August 2013

Sr(II) 3.0 ITA TBP CCI, 3.76 7.52

Ba(II) 2.6 ITA TBP CCI, 2.62 5.84

Eu(III) 8.74 ITA TBP CCI, 5.36 8.96

MAz(H20)w (org) + bB(org) ~ MAzBb(org) + H20(org) K.d (20)


The interaction of these adduct molecules with metallic ions depends strongly on the
organic function in which the donor atom, resides (i.e., on the charge density of the
donor). For oxygen donors, a sequence can be established based on the order in which
they are able to displace each other in the complex:
RCHO < R2CO < ROH < H20 = (RO)lPO < R"R'NCOR = (RO)2RPO < RlPO (21)
A large adduct formation constant means greater extraction which is referred to
commonly as a synergistic effect. Table 5 lists formation constants, (K.dx for the alkaline
earth complexes M(TTA)z(TBP)x in carbon tetrachloride [II]. In the alkaline earth
metals, when x=2, the TBP molecules bond perpendicular to the plane of the two ITA
rings producing an octahedral complex. Since the coordination radius increases in the
order Ca, Sr, Ba (1.00, 1.18 and 1.35 A, respectively), the charge density decreases, in
that order, which leads to weaker bonding as the sequence progresses from Ca to Ba. The
consequence is a decreased synergistic effect for the extraction.
Table 6 lists constants for the formation to Eu(ITA)lB x complexes in chloroform,
or carbon tetrachloride, where B is a series of different donor molecules, arranged in
order of increasing K.d values. In Eu(ITA)l the ITA molecules occupy only 6 of the 9
coordination positions available; the 3 empty positions have been shown to be occupied
1048 CHOPPIN AND MORGENSTERN

Table 6
Adduct constants for formation of Eu(TTA)lBb

Adduct Chloroform Carbon tetrachloride

log KOJ!! !l!lLK.d2 log K,dl log K,d2


Downloaded by [USC University of Southern California] at 01:42 28 August 2013

HTTA(self-adduct) 0.56 >0.5

Hcxone 1.16 1.52 1.71 2.34

Quinoline 3.29 3.48 5.16

TBP 3.63 5.40 5.36 8.96

Tapa 5.40 7.60 7.49 12.26

by three water molecules. The order of increasing K.dl and K.d2, is the same as that of
increasing basicity of the donor molecules. Thus, the order of magnitude of the
formation constant for adducts whose relative position in the basicity sequence or their
relative donor strength can be predicted, provided steric hindrance does not interfere. In
this system CHCIl solvates the Eu complex to some extent, while CC4 is inert. This has
two effects, the K DC value increases due to the solvation by CHCb (not shown in the
Table), while the adduct formation constants K.d decreases as the solvation hinders the
adduct formation. The more inert solvent CCI. causes an opposite effect, a lower Koc
and a larger K ad•

SUMMARY

The important role of thermodynamics in complex formation, ionic medium


effects, hydration, solvation, Lewis acid-base interactions, and chelation, has been
discussed in this paper. An understanding of the factors are of value in assessing solvent
the design of new, improved extraction systems for metal ion separation and purification.
THERMODYNAMICS OF SOLVENT EXTRACTION 1049

Preparation of this paper was assisted by a contract with the USDOE-OBES


Division of Chemical Sciences.

REFERENCES
Downloaded by [USC University of Southern California] at 01:42 28 August 2013

1. H.S. Harned and R.A. Robinson. "Multicornponent Electrolyte Solutions." Pergamon


Press, Oxford, 1968.

2. M.M. Jones, "Elementary Coordination Chemistry", Prentice-Hall, Englewood


Cliffs, N. J., 1964.

3. Y. Marcus, " Ion Solvation", John Wiley and Sons, N.Y., 1985.

4. 1. Rydberg, C. Musikas, and G.R. Choppin, Marcel Dekker, "Principles and Practices
of Solvent Extraction", eds. N.Y., 1992.

5. 1.N. Levine, "Physical Chemistry" 3'd Ed., McGraw Hill Co., New York, 1998.

6. Grenthe, et aI., "Chemical Thermodynamics of Uranium" Nucl. Ener. Agency


Tech. Data Base, OECD, Paris, 1991.

7. K.S. Pitzer, J. Phys. Chern., 7.L 268 (1973); K.S. Pitzer and J.J. Kim, J. Am. Chern.
Soc.,2Q, 5701 (1974).

8. R.D. Shannon, Acta Crystallogr., A32, 751 (1976).

9. R. G. Pearson, Ed., "Hard and Soft Bases", Dowden, Huchinson and Ross, East
Strouburg, Pa., (1973).
10. G.R. Choppin, " Complexation of Metal Ions", Ch.3. in ref. 4.
11.B. Allard, G. Choppin, C. Musikas and 1. Rydberg, "Systematic of Solvent
Extraction:, Ch. 6 in ref 4.

Received by Editor
June 2, 2000

You might also like