Sei sulla pagina 1di 24

Chem. Rev. 1988. 88.

1475-1498 1475

Ionic Radii in Aqueous Solutions

YIZHAK MARCUS
oeparlment of Inwganlc and Analytical Chemishy. The Hebrew University of Jerusalem, 91904 Jerusalem, Israel

Received February 1. 1988 (Revised Manuscript Recehred April 20. 1988)

Contenfs
1. Introduction 1475
11. Some Uses of Ionic Radii in Solution 1476
A. Thermodynamic Quantiiies 1476
E. Kinetic and Transport Quantities 1477
C. Spectroscopic Quantities 1477
I l l . The Radii of Ions 1477
A. The Concept of Ionic Radii 1477
E. The Pair Distribution and Correlation 1479
Functions
I V . Methods for the Determination of Interparticle 1479
Distances
A. Diffraction Methods 1479 Yirhak Marcus was born in Koiberg. Germany, In 1931 and im-
1. X-ray Diffraction 1479 migrated to Israel in 1936. He obtained his MSc. and Ph.D.
2. Neutron Diffraction 1481 degrees in 1952 and 1956 from T h e Hebrew University of Jeru-
3. EXAFS Measurements 1482
salem. He retumed alter postdoctcrai research in Sweden and lhe
U.S. to t h e Soreq Nuclear Research Center in Israel in 1958 to
4. Other Diffraction Methods 1482 work on the actinides and on ion exchange and solvent extraction
E. Computer Simulation Methods 1482 separations. On this appointment in 1965 as Professor at The
1. The Molecular Dynamics Method 1482 Hebrew University. he served until 1973 as Head of the Inorganic
2. The Monte Carlo Method 1483 and Analyscal Chemishy Department. His research interests range
from molten salts. through aqueous, nonaqueous, and mixed
V Ionic Radii of Individual Ions 1484 electrolyte SoIUtions, lo mixtures of organic liquids. He has pub-
A. Cations 1485 lished some 160 research papers and thee books ~1 bn exchange
1. Univalent Cations 1485 and solvent extraction (with A. S. Kerbs). on liqui&state chemisby,
2. Divalent Cations 1487 and on ion solvation. He has served as a visiting professor in
3. Trivalent Cations 1490
uniersities in Germany. Japan, and the U.S. He is holding a Senior
Alexander von Humboldt Fellowship (1971-1976. 1988). and has
4. Tetravalent Cations 1491 Served on IUPAC bodies in various capacities since 1963.
B. Anions 1491
C. Ionic Radii from lnteroarticle Distances 1492
VI. Ionic Radii in Nonaqueous Solvents 1493 versely, it has been argued by Fiashin and Honig2 that
VII. Concluding Remarks 1495 (for the alkali-metal cations) the covalent radii are the
V I I I . References 1496 relevant ones. Referring hack to the usually employed
crystal ionic radii, only if their validity for expressing
I . Introducflon the radii of the ions in solution is demonstrated for a
wide variety of ions on the basis of experimental evi-
In many studies of electrolyte solutions the concept dence can the implicit assumption of the validity of the
of "ionic radius" plays an important, sometimes even use of crystal ionic radii in lieu of the solution ionic radii
a dominant, role. Most properties of such solutions, be maintained.
including thermodynamic, transport, spectroscopic, and In fact, several authors have compared experimental
other properties, depend in one manner or another on evidence that they had obtained on a particular elec-
the ability of solvent or other solute particles to ap- trolyte solution concerning the distance between the
proach a given ion and on their distance-dependent center of an ion and that of the oxygen atom of the
interactions with it. Most authors employ for this nearest water molecules with the sum of the crystal
purpose, without thinking it necessary to justify this, ionic radius of the ion and the 'radius" of a water
the "crystal ionic radii" of ions. These quantities may, molecule. They then stated that the agreement they
indeed, characterize the ions in crystalline ionic solids ohserved between these two distances validated their
and commonly are considered to be inherent properties results. Attempts at a generalization of this finding
of them (but see section IILA). have recently been made by the present author?,' This
It is, however, not self-evident that these quantities review substantiates this contention further, by an ex-
are valid as descriptors of the ions in general, in gaseous haustive examination of the relevant literature.
as well as in liquid phases, and in particular in elec- Studies are deemed to be relevant if they report nu-
trolyte solutions. For instance, it has been argued by merical values for the distance between a specified ion
Stokes' that for ions in the gaseous phase the radii that and the nearest atom belonging to a surrounding solvent
are relevant for their hydration energies are their van molecule in a solution. Many experimental studies on
der Waals, rather than their crystal ionic, radii. Con- the structure of electrolyte solutions do not answer this
0009-2665/88/0788-1475$06.50/0 0 1988 American Chemical Society
1476 Chemical Reviews, 1988, VOI. 88,NO. 8 Marcus

criterion: they may report distances inside complexes saturation takes place, and the solvent is translationally
in solution (central atom-ligand distances), distances immobilized. Only beyond this first shell does eq 1
involving unspecified hydrolysis products, the number hold, with the factor (1 - l / e ) modified appropriately
of nearest solvent molecules without giving a distance, with derivatives of the relative permittivity with respect
etc. Such studies, although of great interest in them- to the temperature and the pressure to accommodate
selves, are not considered in this review. The literature all the derived thermodynamic functions. Many other
has been examined till the end of 1986. models have been suggested, and undoubtedly many
more will yet be, but all must take into account the size
II. Some Uses of Ionic Radii in Solution of the ion in the solution. Many of them, including that
of M a r ~ u sconsider
,~ also the work required to produce
Many phenomena depend on the ability of other a spherical cavity in the solvent for the accommodation
particles to approach an ion in solution. These include of the ion in it, AsolvGocav.The radius of this cavity is
ion-solvent, ion-ion, and ion-nonelectrolyte interac- given directly by Rion,so that this further contribution
tions, giving rise to thermodynamic, kinetic, spectro- (and its appropriate derivatives with respect to the
scopic, and other properties of systems that involve ions temperature and the pressure) to the thermodynamic
in solution. In many cases is it possible to relate the quantities of solvation is also dependent on the radius
phenomena to the radius of the “bare” ion in the solu- of the ion in solution.
tion, i.e., to the radius of the ion without any of its Yet other ion-solvent interactions that give rise to the
solvation shells. In other cases it is the solvated ion, partial molar thermodynamic quantities of electrolytes
with at least its first solvation shell, that is the unit that in solution are related to the sizes of the ions in solution.
is considered, and then the sum of the radius of the bare For instance, the partial molar heat capacity at constant
ion and the diameter of the solvent molecule is the pressure and at infinite dilution6 has been related,
relevant quantity. In general, therefore, the radius of similarly to the corresponding quantity of solvation, to
the ion itself, without its solvation shells in the solution Rion. Many treatments of the partial molar volume7and
is a highly useful and often employed quantity. Fol- entropy8 of electrolytes at infinite dilution have also
lowing is a brief account of some of the instances where been related to Rion.
the radii of bare ions, Rion,in solution can be employed. Ion-Ion interactions at low concentrations but be-
yond the applicability of the DebyeHuckel limiting law
A. Thermodynamic Quantities are governed by the extended Debye-Huckel equation,
The thermodynamic quantities of solvation are which for the activity coefficient has the familiar form
among those that depend strongly on the size of the ion
in solution. Many theoretical approaches, ranging from
log y+ = -Apl/’/(l + Bapl/’) (2)
the simplest consideration of the electrostatic contri- where A and B are parameters that depend on the
bution to the standard molar Gibbs free energy of temperature and the relative permittivity of the solvent,
solvation, .$olvGoel, in terms of the primitive model and p is the ionic strength of the electrolyte solution, and
the Born equation to more sophisticated considerations a is a parameter describing the mean diameter of the
involving the particulate nature of the solvent, dielectric ions or the mean distance of nearest approach of the
saturation, etc., depend on the radius of the ion. In the ions to each other. This quantity, again, may be in-
primitive model, the ion is considered as a sphere with terpreted as the sum of the radii Rion of the ions of
a point charge in its center and the solvent as a di- opposite charge. It is conceded that in many cases this
electric continuum. According to the Born equation sum produces a value of a that does not give a good fit
of the activity coefficients calculated from eq 2 with the
theoretical parameters A and B with the experimental
where NAvis Avogadro’s constant, e the unit charge: eo values, but then there are many objections to the va-
the permittivity of vacuum, t the relative permittivity lidity of eq 2 itself.
of the solvent (its dielectric constant), z the charge of When ion pairing becomes important, then again the
the ion, and Rion its radius. A calculation with eq 1 mean distance of approach of the ions to each other is
overestimates the absolute value of AsolvGoelseverely, a crucial parameter, in both the Bjerrum and the Fuoss
and various modifications have been suggested. The theories. In the Bjerrum theory the expression for the
main merit of the primitive model and the Born equa- association constant is
tion is the indication of how the electrostatic contri- b
butions to derivative thermodynamic functions, such K,,,/(L mol-’) = 4 0 O O r N ~ , u ~t-4
1 exp(t) dt (3)
as the standard molar enthalpy, entropy, heat capacity,
compressibility, etc., of solvation should depend on the where u = Iz+z-le2/4rtotkBT, t is an auxiliary variable,
properties of the solvent. The model and equation kB is the Boltzmann constant, and b = u / a is the limit
specify that these quantities should depend on the ap- of integration. As before, a is the distance of nearest
propriate derivatives of the relative permittivity of the approach of the oppositely charged ions, which should
solvent with respect to the temperature and the pres- equal the sum of their radii in solution, Rion. The
sure. Bjerrum theory specifies ions that are within distances
A recent model that considers all these thermody- a Ir 5 u / 2 of each other to be associated, whether in
namic quantities of solvation but takes into account the contact or separated by some solvent. In the Fuoss
dielectric saturation near the ion is due to M a r ~ u s A.~ theory only contact ion pairs are considered, and the
first solvation shell is defined to extend from Rionto Rion expression for the association constant is
+ AR,where AR itself is a specified function of (the K,,,/(L mol-l) = 4 0 0 0 r N ~ &exp(b)/3
~ (4)
reciprocal of) Rion. In this first shell complete dielectric
Ionic Radii in Aqueous Solutions Chemical Reviews, 1988, Vol. 88, No. 8 1477

It contains the factor exp(b)/3, where at high b values strated by many instances, of which the following are
the Bjerrum theory has approximately exp(b)/b. In a few examples.
both theories, however, b, i.e., a quantity proportional The chemical shift of the NMR signal of the water
to the reciprocal of a, plays a central role. protons, 6, depends in the presence of (diamagnetic)
ions on their concentration and their natures. The
6. Kinetic and Transport Quantities measurable limiting chemical shift per unit concentra-
tion at infinite dilution (dS/dc),,o for an electrolyte
The rate of the solvent exchange between the solva- must be divided into the (additive) ionic contributions
tion shells of an ion that undergoes mainly electrostatic for their interpretation in terms of the interactions that
interactions with them and the bulk of the solvent is take place in the solution. This can be made on the
governed by the electric field of the ion. That is, the basis of the relative sizes of the ions in the solution.12
ratio of the charge of the ion to its radius determines The partially compensating negative shift due to the
this rate: the larger the field strength, the slower the electrostatic field of the ion and a positive shift due to
exchange. Even for transition-metal cations the field the breaking of the hydrogen bonds in the solvent by
strength is important in this respect, although the lig- it can then be evaluated for each ion. The former effect
and field stabilization plays a role, too.g is thus proportional to the reciprocal of the ionic radius,
The rate of complex formation can often be described and the latter effect to the volume of the ion, i.e., to the
as the product of an outer-sphere complex formation third power of its radius.13 The NMR line widths, from
equilibrium constant and the rate constant for the re- which the rates of relaxation of the solvent molecules
placement of a solvent molecule in the inner sphere by in the solvation can be evaluated, are less sensitive to
the entering ligand. The former quantity, the outer- the sizes of the ions.
sphere equilibrium constant, is calculated by eq 4 for The ultraviolet charge transfer to solvent spectra of
a charged ligand or an expression analogous to it, with anions in various solvents have been interpreted in
b replaced by -U/kBT for an uncharged one, where U terms of the "confined model", according to which the
is the interaction energy between the metal ion and the anion is located in a square potential well of a definite
ligand.g The rate of formation of the outer-sphere ion radius.14 This radius depends on the radius of the anion
pair is given, according to Eigen,lo by in the solution, with added terms for the penetration
k = 4sNA&(D+D-)/(exp(au) - 1) (5) distance of the electron into the solvent and the free
space between the ion and the solvent. The latter
where u = ~z+z-~(e2/4~q,ckBZ') as before and Di is the quantity is temperature dependent and can be elimi-
diffusion coefficient of the designated ion. The rate nated by extrapolation to absolute zero, whereas the
constant for the dissociation of the outer-sphere ion pair former added term is solvent dependent but ion inde-
is given,by a similar expression, with -au in the expo- pendent. Thus the knowledge of the size of the anion
nent in the denominator and an additional factor of permits the estimation of the location and shape of the
( 3 / 4 ~ N ~ , ) ain- ~
the numerator. The distance of ap- charge transfer to solvent band due to it.
proach of ligand to cation, a, is thus again involved, Infrared spectroscopy and Raman spectroscopy yield
hence also Rion. evidence for ion-solvent vibrations, the frequencies and
The rate of movement of ions in solution, be it self- force constants of which depend appreciably on the
diffusion or mobility in an electric field, also depends electrical field strength of the ion, and hence on its
on the size of the ion. According to the Nernst-Hartley radius on the solution. For instance, the frequencies
relation, the limiting (infinite dilution) value of the of the symmetric stretching vibration (cation-solvent)
(self-) diffusion coefficient of an ion, Doion,is R T / P of divalent metal ions in liquid ammonia decrease mo-
times its limiting equivalent conductivity, Xoion;hence notonously with increasing ionic radii.15 A similar trend
both phenomena may be discussed in terms of the lim- is observed for these frequencies in solutions of salts
iting ionic mobilities uoion= Xoion/F,where F is the of the alkali-metal ions in dimethyl sulfoxide.16 The
Faraday constant. The mobilities are related to prop- corresponding force constants in tetrahydrofuran and
erties of the solvent and of the ion by the Stokes law dimethyl sulfoxide diminish as the ions become bigger.
expression The presence of ions also affects the frequencies of the
internal vibrations of the solvents, causing positive
frequency shifts. For instance, the C-C and the C=N
where vo is the viscosity of the solvent and Rst is the vibration frequencies in a~etonitrile'~ are shifted to
"Stokes radius" of the ion. This is not necessarily a higher frequencies, again in the order expected from the
valid measure of the size of the moving ion in the so- ionic radii (and charges) for the alkali metal and alka-
lution, in particular not for ions with a "bare-ion" radius line earth metal cations. Since for these kinds of ions
< 0.25 nm. However, a plot of the Stokes radii of ions the interactions between the ions and the solvents are
against their bare-ion radii (actually, their crystal ionic of an electrostatic nature mainly: the nearer the solvent
radii) permitted Nightingalell to arrive at a set of radii molecule can approach the ion, the stronger is its in-
for the moving ions in solution (their so-called solvated teraction with it, and hence this ordering of the fre-
radii). It is thus seen that the bare-ion radii Rionplay quencies, force constants, and frequency shifts.
an important role even in the determination of the
"solvated-ion" radii from experimental mobility data. I I I . The Radii of Ions
C. Spectroscopic Quantities A. The Concept of Ionic Radii
The significance of ionic radii in spectroscopic in- The concept of the size of microscopic particles, such
vestigations of solutions of electrolytes can be demon- as molecules, ions, or atoms, is intuitively understood,
1478 Chemical Reviews, 1988, Vol. 88, No. 8 Marcus

once the notion that matter is microscopically struc- the additivity is preserved. Gourary and Adrian2zhave
tured is accepted. However, it is not simple to define tabulated values of the radii of the alkali metal and the
this size precisely. The wave functions that describe halide ions only.
the electron density around the nucleus of an isolated The radii of polyatomic ions are more difficult to
atom approach zero asymptotically as the distance from define. An operational way to their definition is via
the nucleus increases. Hence, an isolated atom extends their use for the description of thermodynamic prop-
indefinitely into space, and some arbitrary decision (a erties of salts containing such ions; see section 11. Thus,
minimal electron density still considered to “belong” to thermochemical radii have been defined by Kapustin-
the atom) must be made about its size. For atoms in skiiZ3and by Yatsimirskiiz4and have found application
their isolated state, the van der Waals radius is an ac- beyond their defining equations. A recent listing of
ceptable descriptor of their sizes. The collision diam- thermochemical radii for many ions not included in the
eters of some atoms can be measured in the gas phase original listings of such radii was presented by Jenkins
by various method, and they relate well to the sums of and T h a k ~ r .Marcus
~~ and Loewenschussz6have re-
the van der Waals radii of the colliding atoms. Mole- cently suggested that for polyatomic ions of a given
cules, being in general nonspherical, cannot simply be structure (e.g., tetrahedral oxyanions or octahedral
assigned radii, but mean radii can be assigned to them cyano complexes) a constant ratio exists between the
from their experimentally determined “collision thermochemical radii and the characteristic intraionic
diameters”. distances (the distance between the nucleus of the
Ions are generally treated differently, since they are central atom and the nuclei of the oxygen atoms or
generally studied in condensed matter and are not carbon atoms of the cyano groups in the above exam-
usually encountered in the gas phase, i.e., in relatively ples). For many additional ions, therefore, thermo-
isolated states where individual collisions can be noted. chemical radii can be assigned, if it is accepted that this
(This is not strictly true, since many modern tech- constancy prevails for them, too. Such methods work
niques, ranging from mass spectrometry to ion syn- best for ions that have approximate spherical symmetry,
chrotron resonance, produce gaseous ions that can be e.g., regular tetrahedral or octahedral ions, but are not
studied.) Ions are generally found in solid crystalline expected to work well otherwise, Le., for grossly asym-
ionic compounds, i.e., salts. Most of the information metric ions such as acetate. Ions such as nitrate or
on the sizes of ions comes traditionally from the study chlorate are problematic, since molecules or other ions
of such solids, following the observation that the in- may approach them to different proximities, depending
terionic distances dionPion in their crystals are additive on their direction of approach. Such ions may be as-
in terms of the contributions of the cation and the an- signed mean radii, with limited applicability.
ion. These distances are obtained with very high ac- It is not immediately obvious that the radii of ions
curacy and precision (f0.000, nm or even better) by that were established in crystalline compounds, whether
modern X-ray diffraction techniques. on the basis of interionic distances dioWionor on the basis
Goldschmidt18 and Paulinglg provided in the 1920s of thermochemical data, are also applicable to ions in
sets of radii for monoatomic ions, based on measured liquid solutions. One step toward the resolution of this
dion-ionvalues in crystalline salts and on theoretical problem is the examination of ionic radii or dion-ion
considerations of how these distances should be ap- values in ionic liquids, i.e., molten salts. It must be
portioned among the ions. Such radii are constant noted that due to the thermal expansion, dion-ionis
characteristics of ions in various crystal lattices, pro- temperature dependent in both the crystals and melts.
vided that the coordination number is the same as that Experimental values of dion-ionof most of the molten
in the crystals employed for their determination. A alkali-metal halides near the melting point were given
comprehensive list of ionic radii, valid for crystals, and by Levy and Danfordz7(see also Marcusz8). They are
mainly for coordination number 6, based on the Pauling 1-10% shorter than in the corresponding solid salts
theoretical considerations, was more recently presented near the melting point. However, the “coordination
by Shannon and Prewitt.zOThese radii reproduce the numbers”, i.e., the number of nearest neighbors, in the
measured in most fluoride and oxide lattices, melts are considerably smaller than in the crystals: they
with the specified radii of the anions: R ( P ) = 0.133 nm are 3.5-5.6 rather than 6. The fewer the neighbors, the
and R(O2-) = 0.140 nm. They are in substantial shorter is on the whole the distance. (This shortening
agreement with the original set of Pauling radiilg and is countered by a lengthening of the distances to the
with their revisions by Ahrens21 (within f0.005 nm). next nearest neighbors, which have the same sign of the
Although these Pauling-type ionic radii for coordi- charge, and hence are repulsed.) There seems, there-
nation number 6 are widely used, they are not univer- fore, to be a close connection between the coordination
sally accepted, since the theoretical basis used by number of an ion and its size not only in crystalline
Pauling for the apportioning of dion-ion between the solidsz0but also in liquid condensed phases, and pre-
cation and anion can be challenged. For instance, sumably also in aqueous solutions.
Gourary and Adrianzzconsidered the electron density The following questions then arise: Are there any
as a function of the distance from a given ion in a so- experimental techniques that allow the determination
dium chloride lattice. They preferred to take the extent of accurate distances between the centers of ions and
of the space around the sodium nucleus that is to be their immediately surrounding solvent molecules in
assigned to the sodium ion as that which is limited by solutions? If so, are there theoretical means for the
the minimum in the electron density curve. The extraction of ionic radii from such measured distances?
Gourary and Adrian radii of cations are then consid- The answer to the first question is definitely: yes!, as
erably larger than the Pauling radii, and those of anions is shown in section IV. Judgment concerning the sec-
are considerably smaller (e.g., R(F-) = 0.116 nm), since ond question is deferred till the end of this review. It
Ionlc Radii in Aqueous Solutions Chemical Reviews, 1988, Voi. 88, No. 8 1470

is necessary first to see how ionic radii might be ob- The r value of the first peak in the dependence of gi.(r)
tained from the experimental results and their struc- on r gives the shortest interatomic (internuclear) dis-
tural interpretation. tance dij between the two particles i and j . The region
of space around the particle i from the value of r where
B. The Pair Dlstrlbution and Correlation gij(r) becomes positive just before its first peak to the
Functlons value of r that corresponds to the first minimum (valley)
in gij(r) is called the first solvation (hydration) shell of
The probability of the presence of a particle (mole- the ion i. The number of nearest j neighbors of the
cule, atom, ion) of a given kind i in an element of particle i is given by the integral of 4a?gi.(r) from the
volume of a homogeneous isotropic fluid (liquid) is given lower value of r to its upper value that dednes this first
by its number density, pi = NJV, where Niis the shell. Other definitions of this number, depending on
number of such particles in the total volume V of the the asymmetry of the first peak in gij(r), are also pos-
fluid. This probability is also called the singlet dis- sible; the differences are generally minor.
tribution function of i, symbolyzed by ni(l). Once dion-water is known from the pair correlation
The conditional probability of finding a particle of function gion-water(r), the radius of the ion, Rion,can be
kind j in an element of volume at a distance r away defined as
from a given particle i in such a fluid is correspondingly
called the pair distribution function. It is symbolyzed Rion = dion-water - Rwater (9)
by nii2)(r),provided that the particle i exerts a spher-
ically symmetrical force field. Otherwise nii2)depends where Rwateris an appropriate distance that character-
izes the “radius” of a water molecule.
not only on r but also on the mutual orientation of the
particles i and j . As r tends to increasingly large dis- One approximation of this distance is half of the
tances, the conditional probability of the presence of mean intermolecular distance in liquid water,
j in the element of volume at the distance r from the that may be taken as the mean “diameter” of the water
origin of coordinates becomes increasingly independent molecules. Several authors have determined the radial
of the presence of i there, and nij(2)(r)tends to p2 (if i distribution function of liquid water, using X-ray,
and j are identical in nature). A t short distances, electron, and neutron diffraction and Monte Carlo and
however, of the order of up to 5 molecular diameters, molecular dynamics computer simulation
particle i influences the probability of the presence of The resulting value of RWa,, is temperature dependent;
particle j , and at room temperature, where most of the determinations
of dion-water
have been made, the average of the reported
nij(2)(r)= gij(r)p2 (7) values is Rwater= 0.1420 f 0.0005nm. It can, however,
where gi.(r) is called the pair correlation function. (If be argued that the water molecules near an ion are
i f j then pipj replaces p2.) At large values of r, packed closer than they are in liquid water at room
therefore, gij(r) tends to unity. At very small values of temperature. If the intermolecular distance between
the water molecules in ice at the melting point is taken
r, where strong repulsion excludes the j particles from
the neighborhood of the i particle, gij(r) is zero. A t to be a better estimate, the result is RWa,, = 0.1383 *
intermediate values of r, gij(r) oscillates between max- 0.0002 nm.33 This point is commented on further in
ima (peaks) and minima (valleys). section V.C.
The number of particles j in a spherical shell of
thickness dr at a distance r from a particle i is the I V. Methods for the Determination of
(differential) radial distribution function Interparticle Distances
dnij(r, dr) = pjgij(r)(4?rr2)dr (8) A. Diffraction Methods
The integral of this radial distribution function up to When a monochromatic beam of radiation hits a
a distance r’ gives the number of j particles that sur- liquid sample, some of it is transmitted unchanged,
round the particle i up to this distance. some is absorbed, but also some is diffracted at various
Although liquids and liquid solutions are devoid of angles 6’ from the direction of the incident beam. The
long-range order, there exists short-range order in them. intensities that are diffracted (scattered) at various
The partial order persists over 3-5 molecular diameters angles provide information on the structure of the liq-
and is expressed by the deviation of the radial distri- uid. The radiation employed can be X-rays, neutrons,
bution function from the smooth parabolic function or electrons, and various methods revolve around the
nij(r) = pj(4?rr2)or equivalently by the deviation of the use of these kinds of radiation. Diffraction studies of
pair correlation function gij(r) from unity. As the dis- the structure of liquids were reviewed in Chemical
tance from a given particle increases, the order in the Reviews long ago by K r ~ h . ~ ~
liquid decays and is harder and harder to discern, until
complete randomness in the mutual arrangement of the 1. X-ray Diffraction
particles is reached. The above-mentioned deviations
at short distances are the expression of the “structure” In an X-ray diffraction experiment conducted on a
of the liquid or the solution. This structure, expressed liquid, the direct experimentally observed quantity is
as the g&) function, can be obtained from experiment the intensity I’(6’)of X-rays scattered at an angle 0 from
or computer simulation, as detailed in section IV. It the incident beam. The X-rays scattered at this angle
leads eventually to the interatomic distances that are are assigned a wavenumber k
the subject of this review. k = 47rX-I sin (6/2) (10)
In the present context we are interested in particles
i that are ions and particles j that are water molecules. where X is the wavelength of the (monochromatic)
1480 Chemlcal Reviews, 1988, Vol. 88,No. 8 Marcus

X-rays. The mean intensity of X-rays scattered from difficulty is the insensitivity off to light elements and,
a liquid sample over a unit spatial angle, dI/dQ, is given in particular, to hydrogen, so that the conformation of
by the water molecules near the ion cannot be determined.
u / d n = [ ~ ’ ( e ) / ~ (-e s(e)
)] A further difficulty is the appearance of spurious small
(11) peaks in the Fourier-transform-calculated function
where u(0) and s(8) are calibration parameters, which Gij(r)at low values of r that arise from the truncation
have to be determined experimentally or estimated of the S J k ) function at high values of k. Generally,
theoretically. The former depends on the sample ge- these spurious peaks are disregarded, since the main
ometry, the apparatus geometry, and the absorption in (first large) peak stands out sufficiently clearly.
the sample and involves polarization corrections. The The main difficulty with the X-ray diffraction me-
latter involves Compton and multiple scattering from thod for the determination of the ion-water interpar-
the sample. On the other hand, the mean intensity over ticle distance arises from eq 13. Since the solution
a unit angle is related to the properties of the liquid by contains both cations C and anions A, there are six
distances that contribute their partial structure factors
dI/dQ = N[Ccifi2(k) FT(k)]
i
+ (12) to FT: W-W, C-W, C-C, C-A, A-W, and A-A, where
W symbolizes the oxygen atom of the water molecules
where N is the number of scattering centers (atoms) in nearest to the ion. This is the least number of relevant
the sample, ci is the molar concentration (moles per unit distances that applies in solutions containing only
volume) of the centers (atoms) of type i, and f i ( k ) is monoatomic ions and is a result of the insensitivity of
their coherent scattering amplitude, an inherent prop- the X-ray scattering intensities to the positions of the
erty that depends on k, the wavenumber of the X-ray hydrogen atoms (these would have added four addi-
radiation, and is available in tables. The term that is tional distances, W-H, H-H, C-H, and A-H).
sensitive to the structure of the liquid (solution) is It is in general impossible to extract the above-men-
F,(k). tioned six distances individually from the X-ray scat-
The experimentally obtained FT(k)is related to the tering results on a solution of CA in water. This is due
structure by to the impossibility, in the case of X-ray diffraction
experiments on solutions, to vary f(k) of the atoms of
interest in a systematic manner. An approximation to
where the summation extends overall the kinds of such a systematic variation is the use of isomorphous
scattering atoms in the liquid sample (including the substitution. The atom of interest is substituted by
cases of j = i) and where the S i j ( k ) are the partial another one having similar properties (with the ex-
structure factors pertaining to the pairs of atoms i and pectation that it will have the same gion-water(r) and
j in the solution. These, in turn, are related to their pair dion-water) with the exception of having a different f ( k ) .
correlation functions, gij(r),as follows: This property depends strongly and monotonously on
the atomic number 2 of the scattering atom. This
method has only seldom been applied. Recourse is
Sij(k) = 1 + ( 4 * N / k n J m0g i j ( r ) sin (kr)r dr (14) generally made instead to the following procedure.
Even in the more concentrated solutions studied (e.g.,
where V is the volume that contains the N scattering those containing up to 3 mol dm-3 of CA in water) most
centers and r is a vector from particle i. A Fourier of the scattered X-ray intensity still generally arises
transformation can be applied to the partial structure from the W-W distances, since the water molecules
factor, S i j ( k ) ,in order to obtain the individual pair outnumber the ions by about 10 to 1 or more. In the
correlation function gij(r): often utilized first-neighbor model (FNM), the W-W
gij(r) = distances in the bulk of the water, outside the first
hydration shells of the ions, are assumed to be the same
1 + ( 2 . r r 2 ( N / n r ) - ’ 0S m ( S U ( k-) l ) k sin ( k r ) dk (15) in the electrolyte solution as in pure water. Their
contribution to the observed G ( r ) may therefore be
However, application of the Fourier transform to the subtracted out. The interionic C-C, C-A, and A-A
experimental FT(k)data results in an observed overall distances can be neglected in not too concentrated so-
G(r) curve, in which the individual gij(r) curves are lutions, since their contribution to FT depends on the
superimposed. The individual g&) curves must then square of the concentration. (Sometimes,however, C-A
be disentangled from the observed G(r). distances are important when ion pairs are formed in
Several difficulties are inherent in the use of X-ray the solution.)
scattering from electrolyte solutions for the purpose of There remain, therefore, only two distances that have
the determination of gian-water(r)and the ion-water to be determined, C-W and A-W, provided that the
distance. One difficulty is the dependence of the f FNM is applicable. The superposition of the gcw(r) and
values, the scattering amplitudes of atoms, on k, and, gAW(r)curves (after subtraction of the concentration-
in turn, on 8; see eq 10. A consequence of the k de- weighted gww(r) curve from the “observed G ( r ) )leads
pendence off is that the overall pair correlation func- to a remaining g(r) curve that exhibits the peaks at the
tion G ( r ) ,which is defined by characteristic distances C-W and A-W. These can
generally be distinguished, although a second peak in
G ( r ) = ( 2 r 2 ( N / V ) r ) - l L m F T ( k )sin
k ( k r ) dk (16) the gcw(r) curve, due to second-nearest-neighbor water
molecules of the cation C that are ignored in the FNM,
and results from the Fourier transform, is not a linear may obscure the first peak in the gAW(r)curve in the
combination of the individual gij(r)functions. Another combined g(r). It is also helpful to assume a geometry
Ionic Radii in Aqueous Solutions Chemical Reviews, 1988, Vol. 88, No. 8 1481

and a coordination number around the ions, since then A major difference between X-ray and neutron
W-W distances within the first coordination shell scattering from solutions arises from the different de-
around an ion, which would superimpose their contri- pendencies of the coherent scattering amplitudes f on
bution to g(r) on those of the other characteristic dis- the wavenumber k of the radiation and on the atomic
tances, are uniquely determined by the ion-W distance. mass (mass number, A ) and atomic number 2 of the
These W-W distances can be identified in the curve, scattering atoms. Whereas in X-ray diffraction f de-
and their contribution to g(r) can be subtracted out. creases monotonously with 12, it is independent of it for
In many cases, however, it was found that the FNM neutron scattering. On the other hand, whereas the f
cannot account adequately for the experimental results. for X-ray scattering increases monotonously with the
A second hydration shell, attached by hydrogen bonds atomic mass of the atoms, for neutron scattering it is
to the first one and held also by the intense electrical very sensitive to the combination of their A and 2. For
field of a small and highly charged (z > 2) cation, must certain combinations (i.e., the combination of the
be taken into account. One consequence of this sec- number of protons and neutrons in the nucleus of a
ond-neighbor model (SNM) is that the amount of water given isotope of an element) f is very large, while for
present as “bulk water” is diminished. More assump- other combinations (in another isotope of the element)
tions, interatomic distances, and geometries are thus it may be near zero, of even negative. Examples for this
introduced by the SNM, until the desired fit of the are the values (in pm) of f for ‘H, -0.372, and 2H
computed F T and the experimental one is accomplished. (deuterium), 0.667; for 35Cl,1.17, and 37Cl,0.29; or for
The price to pay for this better fit is a reduction in the 58Ni, 1.44, and 62Ni,-0.87. A consequence of the ap-
credibility of the results, due to the added fitting pa- preciable value off for hydrogen atoms is the sensitivity
rameters and assumptions. of the neutron diffraction results to the relative posi-
In spite of all these difficulties, the ion-W distances tions of the hydrogen atoms in the aqueous solution (so
can nowadays be determined with modern instrumen- that the four distances H-H, O-H, C-H, and A-H now
tation and computation methods to a remarkable degree contribute to F T ) . Therefore the average configuration
of accuracy in the simpler cases. Uncertainties of 0.0003 of the water molecules near the ions can, in principle,
nm are often given for the reported distances. This be studied by the neutron diffraction method. Since
uncertainty may be realistic for cations, provided that ‘H nuclei have an appreciable absorption cross section
the assumed geometry of the water molecules around for thermal neutrons, most scattering experiments on
the cation is correct, which seems to be the case for the aqueous solutions are conducted with heavy water, D20,
regular octahedral first coordination shell of transi- but it is assumed that this does not affect the ion-water
tion-metal ions, provided no inner-sphere complexing distances or the configuration of the water molecules
takes place. In other cases, a more realistic estimate around the ions.
of the uncertainty of the determination of the ion-W The necessity to disentangle 10 partial structure
distance would be 0.001 nm or worse. In the earlier factors Sij(k)from the experimental FT according to eq
determinations, up to the late 19509, the state of the 13 (or 10 individual gij(r) from G(r) according to eq 16)
art allowed only determinations to within 0.005 nm or makes the direct neutron diffraction method
worse. impractical-this in spite of the fact tl at G(r)is a linear
combination of the gij(r),due to the independence of
2. Neutron Diffraction f for neutron scattering from k, contrary to the case of
X-ray scattering. A further important complication is
Neutron diffraction from electrolyte solutions is ap- the uncertainty due to the large and not fully known
plicable for the determination of the interatomic dis- Placzek corrections for inelastic neutron scattering.
tances similarly to X-ray diffraction. Whereas X-rays However, a further consequence of the sensitivity of
are scattered by the electronic shells of the atoms, the scattering amplitude f on the isotopic composition
neutrons are scattered by their nuclei. The overall is the possibility of using the first-order difference
approximate spherical symmetry of the electronic shells neutron scattering (FODNS) functions to obtain ion-
cause also the X-ray diffraction results to give the water coordination values (Soper et Enderby and
relative positions of the nuclei of the atoms, so that N e i l ~ o n ~ Suppose
~). that neutron diffraction experi-
there is no net difference between the methods arising ments are conducted on two solutions of an electrolyte
from this respect. Equations 10-16 continue to be valid CA in (heavy) water that are identical in all respects,
(the wavelength of the neutrons in eq 10 is obtained except in the isotopic composition of, say, the cation C.
from their kinetic energy), but there are some very im- Suppose also that the coherent scattering lengths of the
portant differences. One difference arises from the two compositions, f c and fb, differ appreciably. Then
factors that affect the calibration parameters a(0) and the difference between the corrected neutron scattering
s(0). The former depends, as in X-ray scattering, on the cross sections (intensities) between the two experiments
sample and instrumental geometries and absorption by will depend only on those terms in eq 12 that involve
the sample, but it also depends on absorption by the the cations:
sample container. The latter depends, as in X-ray
diffraction, on multiple scattering, but it also depends A d k ) = Ao(Sco(k) - 1) + AD&&) - 1) +
on scattering by the container, incoherent scattering, A c ( S c c ( k ) - 1) + A A ( S C A (-~1)
) (17)
and inelastic (Placzek) scattering. All these effects must
be corrected for, either experimentally or from theo- Here the species j in the partial structure factor Scj is
retical calculations. The Placzek corrections are, un- 0, D (for deuterium), the cation C, or the anion A, and
fortunately, not fully understood and not easy to apply, the Aj coefficients are products of the concentrations
but contribute heavily to the observed scattering in- of C and of j , the difference f c - f’c of the two isotopic
tensities. compositions of C, and f j . Note that the water-water
1482 Chemical Reviews, 1988, Vol. 88, No. 8 Marcus

correlations, which dominate in the X-ray diffraction a consideration of the entire body of data in the present
equation, are absent from eq 17. Furthermore, the review, has not been commented on so far. On the other
terms in Ac and A*, depending on the square of the hand, the application of EXAFS is still so recent and
concentration, are generally quite small compared with the amount of information gathered by it is still so scant
those in Ao and AD, which depend on its first power, that no great weight can as yet be placed on these re-
except in extremely concentrated solutions. Thus the sults. A deterrent to a wider application of this method
analogous equation to (9) where Ac(k) replaces FT now is the necessity of employing very intense X-ray beams,
becomes a simple linear combination of gco(r) and available in only a few cyclosynchrotron machines
gcD(r). If suitable isotopes exist for the anion, a similar throughout the world.
pair of experiments can be conducted on solution where
different isotopic compositions of the anions are em- 4. Other Diffraction Methods
ployed. A further advantage of the FODNS method is
Electron diffraction has been applied only very rarely
the cancellation of the major part of the Paczek cor-
to the study of solutions, and aqueous electrolyte so-
rections for inelastic scattering in the difference. This
lutions in particular. In one such study of aqueous zinc
reduces the uncertainty caused by them very consid-
erably. Similarly to the FODNS method, second-order bromide (Kalman et al.39), 68-keV (A = 0.005 nm)
electrons were employed. Corrections for scattering by
difference measurements that involve four isotopic
the apparatus and its geometry were applied, as were
compositions with sufficiently different f values can give
corrections for atomic self-scattering, multiple scat-
information on ion-ion distances, geometries, and in- tering, and inelastic scattering. The results were in good
teractions, but this subject is outside the scope of this agreement with a parallel X-ray diffraction study. No
review.
particular advantage of using electron diffraction was
Neutron diffraction results by the FODNS method
cited.
are limited to ions from elements that have suitable
pairs of isotopes, differing considerably in their neutron
B. Computer Simulation Methods
scattering cross sections. The results are rarely reported
to better than iO.001 nm (a study of aqueous dyspro- Simulation by computers is a powerful method for the
sium chloride (Annis et al.37) did report the Dy-0 “experimental” generation of fluids that yields at the
distance to ~0.000,nm). Contrary to X-ray diffraction same time many of their properties, such as thermo-
results, however, those from the FODNS method are dynamic or transport properties, as well as their
devoid of assumptions concerning the geometry of the structure, in terms of the pair correlation functions.
arrangement of the water molecules around the ion. The “experimental” generation of the fluid requires as
input specification the number of particles of each kind
3. EXAFS Measurements in the system, the volume of the system, the tempera-
ture or total energy of the system, the boundary con-
Extended X-ray absorption fine structure (EXAFS) ditions (generally a periodic boundary condition), and
has only recently started to be employed for the de- the forces that act between the particles. In general use
termination of ion-water distances in solutions, but it are the molecular dynamics method, which yields not
has proved to be a useful tool for probing the liquid only the equilibrium structure of the fluid but also its
structure near the ions. It is not capable, however, of dynamic behavior, and the Monte Carlo method, which
providing information on long-range order, but this is yields the structure and equilibrium properties only.
not a limitation in the present context. EXAFS shares The success of a computer simulation depends very
with the differential neutron scattering method (FOD- strongly on how realistic indeed are the interparticle
NS) the capability of giving direct information on the forces that are used. If a range of thermodynamic (or
distances involving a given atomic species (say, the transport) properties are predicted satisfactorily, then
cation C), contrary to the X-ray diffraction method. it is probable that the structure predicted is also ac-
EXAFS does so by the choice of the absorption edge ceptable. With present-day computers a simulation
of the X-rays that are employed and the element that involving 25-500 particles that yields acceptable results
provides these X-rays. The primary information of the is possible at costs in computer time that are within the
EXAFS method is x ( k ) data, where x is the absorbance capabilities of many institutions. This capability is
of the X-rays after pre-edge subtraction and the ap- probably going to become more widespread in the fu-
plication of appropriate corrections. Fourier transfor- ture. The application of computer simulation methods
mation of the x ( k ) data is then made, leading to a to the study of the structure of aqueous electrolyte
combined pair correlation function G(r). This must be solutions was reviewed recently by H e i n ~ i n g e r . ~ ~
deconvoluted into the partial gij(r) that constitute it,
as in the differential neutron scattering method. One 1. The Molecular Dynamics Method
advantage of the EXAFS method is that it is applicable
to much lower concentrations W O . 1 mol dm-3) than In the molecular dynamics method one ion (or, in
ordinary X-ray diffraction (Licheri et al.38). It is some cases, a cation and an anion) and a much larger
therefore freer from complications due to ion pairs and number of water molecules (25 to typically 215) are
the imposition of cation-anion distances on the desired placed inside a confined space of a size commonly de-
cation-water and anion-water distances. termined by the macroscopic density of water or an
The reported uncertainties of the ion-water inter- electrolyte solution of the given concentration at the
nuclear distances obtained by EXAFS are C0.003 nm, given temperature. They are placed there at a random
but these distances are generally and systematically configuration, and each is given a random velocity (in
1-270 shorter than those obtained from X-ray or neu- a random direction) at the start. The interparticle
tron diffraction. This observation, which resulted from (ion-water and water-water) forces are then “switched
Ionic Radii in Aqueous Solutions Chemical Reviews, 1988, Vol. 88, No. 8 1483

on”, and the dynamics of the system are followed as a Lennard-Jones potential for ion-water interactions is
function of time. The positions and the momenta of given by eq 18, and the appropriate parameters (width
all the particles are determined after each small incre- and depth of the potential well) are taken from models.
ment of time, At, by the application of the classical For cations, these are the isoelectronic noble gases, and
Newtonian laws of motion and the conservation of the a Berthelot-type mixing rule is adopted to obtain the
sum of the kinetic energies of all the particles and the ion-water interactions from the water-water and ion-
potential energies between each particle and its neigh- ion ones. This mixing rule specifies the arithmetic
bors (over their effective range). After a few thousands mean for aij and the geometric mean for cij. Other
of time intervals At the system reaches dynamic equi- combination rules (e.g., those of Kon$l) have also been
librium, due to momentum and energy interchanges on used. Otherwise, ab initio calculations have also been
collisions or near approaches of pairs of particles. Ac- employed. The Coulombic ion-water interactions are
cumulation of the positional configurations for a few added to the Lennard-Jones ones:
more thousands of intervals then presents a good sta- 4
tistical sample, which may be averaged in terms of the von-w(r,d)= -qzC(-1Y/di (20)
individual pair correlation functions, and in particular i=l
in terms of gion-watar(r). where z is the charge number of the ion.
In the past, when fast and highly powerful computers It must be emphasized that the $T2 model for water
were not yet available, the molecular dynamics method and the Lennard-Jones plus ion charge-assigned water
was applied mainly to systems with very simple inter- charge (q in the ST2 model) interactions for the ions
particle forces: hard disks (in two dimensions) and hard are not the only ones that have been investigated or
spheres (in three dimensions). In recent years the ca- necessarily the best ones to use. The results from the
pabilities of the computers increased so much that this molecular dynamics simulation are fairly sensitive to
limitation is no longer relevant, and more or less real- the input data and functional form of the interparticle
istic interparticle forces can be employed. forces employed, so that refinements are possible when
A commonly used model for the water molecules is more sophisticated models are used. Adjustments can,
the ST2 which pictures the water molecule in principle, be applied in order to fit as well as possible
has having a tetrahedral geometry, with the oxygen the thermodynamic and transport (e.g., diffusion)
atom in the center. The angles are the regular tetra- properties of the system, without trying to fit explicitly
hedral ones, and there are hydrogen atoms at two of the the interparticle distances found by diffraction mea-
apexes at a distance of 0.1000 nm from the center, each surements. Under such circumstances the results from
carrying a charge of q = +0.23 elementary charges. At the molecular dynamics simulation could be taken as
the other two apexes there are negative charges of q = independent measures of the ion-water distances. It
-0.23 elementary charges (representing the lone pairs must be remembered, however, that the potential for
of electrons of the oxygen atom) at distances of 0.0800 the ion-water interactions involves a distance parameter
nm from the center. This charge distribution, obtained (say, the effective width of the Lennard-Jones potential
from quantum chemical calculations, is assumed, al- well) taken from a model, so that it may be questioned
though the resultant dipole moment is somewhat higher if an altogether independent estimate of the prevailing
than that measured on gaseous water molecules. The ion-water distance can be obtained from the resulting
lengths of the 0-H bonds and the angle between them
are near those measured by electron diffraction for an gion-water(r) curve.
isolated water molecule, but not exactly the same. 2. The Monte Carlo Method
A Lennard-Jones type potential is centered on the
oxygen atom to account for dispersion forces and re- The Monte Carlo method of computer simulation of
pulsion at short distances. The Lennard-Jones poten- fluids is similar to the molecular dynamics method in
tials for both the water-water and the ion-water in- that an ion (or a pair of a cation and anion) and many
teractions are of the form water molecules (25 up to 512) are placed into a con-
fined space with a periodic boundary at some random
configuration. The density and total energy (corre-
sponding to a given temperature) are specified, and
where cij is the depth and aij is the effective width of hence also the volume of the system. The potential
the potential well. The Coulombic terms for the energy of the system is computed on the basis of as-
water-water interaction are sumed (normally painvise) interactions between all the
4 particles, employing specified interparticle forces. In
Ucww(r,d)= Sww(r)q2C (-~)~+j/dij (19) each following step one particle selected at random is
ij=1
moved (and reoriented, if not spherical) randomly to
where r and d are distances between oxygen centers and yield a new configuration, and the new potential energy
point charges, respectively, and Sm(r)is a “switching of the system is computed. If this energy is lower than
function” introduced32bto reduce excessive, unrealistic the previous one, then this step is accepted and the next
Coulombic forces. According to this model the hydro- step is attempted. If the new potential energy is higher
gen bonds formed in liquid water are linear, and the than the previous one, the new configuration may still
intermolecular distances agree with the experimental be accepted, depending on how much this energy is
ones measured on pure liquid water.32 higher, but otherwise the step is disregarded. In this
The ions are commonly described as spheres with manner the potential energy on the whole diminishes
point charges at their centers that undergo Lennard- in successive steps, approaching equilibrium asymp-
Jones type interactions in addition to charge-dipole totically. When equilibrium is practically achieved in
interactions with the neighboring water molecules. The the system, the potential energy fluctuates around its
1484 Chemical Reviews, 1988, Vol. 88,No. 8 Marcus

TABLE I. L i - 0 Internuclear Distances in Aqueous TABLE 11. Na-0 Internuclear Distances in Aqueous
Solutions Solutions
salt W/salt di0,/nm n method ref salt W/salt dion+./nm n method ref
Li+ 64 0.204 6.0 MD 47 Na+ 200 0.240 MC 44
LiCl 3-8 0.195-0.225 4 X 4Y Na+ 64 0.235 6.2 MD 47
LiCl 4.4-24.5 0.21 X 50 NaClO, 25 0.236 6.0 MD 51
LiI 25 0.213 6.1 MD 51 Nat 215 0.235 6.0 MC 51
Li+ 215 0.210 6.0 MC 51 Na+ 64, 125 0.229 6.0 MD 52
Li+ 64-125 0.198 5.3 MD 52 Na+ 215 0.237 6 MD 53
Li+ 215 0.213 6 MD 53 Na+ 215 0.235 6.0 MC 60
LiCl 4.0 0.200 MD 54b Na+ 64 0.23 MC 61
LiI 25 0.212 6.1 MD 55 NaCl 13.9-27.8 0.242 4 x 65
LiI 25 0.2157 6 MD 56 NaCl 25 0.23 7 MD 68
LiI 25 0.210 6 X 570 Na+ 10 0.23-0.24 5-6 theory 70
LiCl + CoC12 7.5 0.203 X 58 Na+ 200 0.233 4.3 MC 71
LiCl + CoClz 8-17 0.199-0.207 4 X 59" NaCl 54.3 0.250 8 N 72
Li+ 215 0.210 6 MC 60 NaCl 10.2 0.241 6 X 73
Li+ 64 0.19 MC 61 NaCl 100 0.23 7 MD 74
LiCl 4.0 0.211, 0.221 MD,X 62 NaCl 25 0.23 6 MD 75
LiCl 4.0 0.218, 0.222 4 X, MD 63 NaCl 25 0.230 6 MD 76
LiCl 5.6-15.6 0.195 5.3-3.3 N 64 NaI 7 0.24 4 x 77
LiCl 13.9-27.8 0.208-0.217 4 X 65 NaN03 6.1-9.3 0.246 6 X 78
LiCl 54.3 0.190 4 X 66 NaCl 100 0.23 6.1 MC 79
LiBr 8.4-25.0 0.216-0.225 4 X 67
LiI 25 0.206 7 MD 68 TABLE 111. K-0, Cs-0, and Ag-0 Internuclear Distances
LiI 25 0.213 6.1 MD 69 in Aqueous Solutions
Lit 10 0.19-0.20 4 theory 70
Li+ 200 0.195 4.0 MC 71 salt W/salt dion-w/nm n method ref
K+ 200 0.265 MC 44
OAssumed n. bThe Lit and C1- ions interact. K+ 64 0.286 7.3 MD 47
K+ 215 0.271 6.3 MC 51
equilibrium value. The positions of all the particles are K+ 64, 125 0.276 7.5 MD 52
then noted and averaged over several thousand addi- K+ 215 0.286 6.6 MD 53
tional configurations to yield the desired pair correlation K+ 215 0.270 6 MC 60
K+ 64 0.28 MC 61
function gion-water(r). KCl 13.9, 27.8 0.280 6 X 65
The same kinds of interparticle forces and models for K+ 10 0.285 5-7 theory 70
the ions and the water molecules as are used for the K+ 200 0.283 5.1 MC 71
molecular dynamics calculations can be used also for KCl 53.7 0.270 8 N 72
KOH 3.2 0.292 X 80"
the Monte Carlo ones. However, in the applications to KCl 12.5 0.26 N 81
the present problem ab initio calculations have been KF 13.3-27 0.295 X 82
used more often, with a Hartree-Fock type calculation KCl 0.27 X 83
with superimposed dispersion forces. About lo5 con- CsF 25 0.322 7.9 MD 51
figurations were collected in a typical calculation, but CsCl 13.9, 27.8 0.315 6, 8 X 65
problems due to a possible inadequacy of the periodic CsF 25 0.310 7 MD 68
boundary condition for the long-range ion-water in- CsF 25 0.322 7.9 MD 84
CsCl 53.1 0.295 8 N 85
teraction and a dependence on the number of particles CsC1 11.1-22.2 0.315 X 86
included in the system may not have been solved com- CsF 2.3-8.0 0.313 6 X 87
pletely satisfactorily. Again, thermodynamic quantities
are obtained from the calculation (but not dynamic AgClO4 10.4 0.2410 2.0 x 88
AgN03 14.2 0.2450 2.45 X 88
properties of the system), besides the structural infor- AgC104 10.6 0.241 3.9 N 8gb
mation. The validity of the latter can be judged from AgClO4 5-16.8 0.236-0.231 3-4 EXAFS 90
the agreement of the computed thermodynamic quan- AgClO4 3.3-16.3 0.243-0.238 4 X 91'
tities with experimentally determined ones. AgNO3 4.3-16.6 0.243-0.242 4 X 91'
The Monte Carlo method has been applied less fre- W-W and OH--W distances overlap the K+-W ones. Solution
quently than the molecular dynamics computer simu- contained HCIOa. Assumed n.
lation method for the purpose of the determination of
the structure of aqueous electrolyte solutions, specifi- TABLE IV. O(HJ)-O and N(H,)-0 Internuclear Distances
cally yielding ion-water distances. Both methods have in Aqueous Solutions
so far been confined to the study of solutions of the salt W/salt d2,,-wlnm n method ref
alkali-metal cations and halide anions, except for one HCl 3-31 0.252 N 44
study concerning the methylammonium and methyl- HC1 4-96 0.252 X 44
carboxylate (i.e., acetate) ions by the Monte Carlo DCl, DBr 54.9, 51.7 0.288 N 92"
H2SO4 30.5 0.275 X 93b
method.42 HC104 38.5 0.276 X 946
HC1 4.5 0.244 X 95
V. Ionic Radii of Individual Ions 0.290
HC10, 0.275 X 96
In Tables I-XI1 are presented the values of the in-
ternuclear distances dion-water(cf. eq 9) that have been NHdCl 25 0.305 8.1 MD 85
ND4Cl 10.0 20.260 110 N 97c
obtained by the methods outlined in section IV and
reported in the open literature till the end of 1986. The "Also n = 3 is compatible with data. bAssumed n.
tables are arranged according to groups of ions and Coordination of D20 is not well defined.
Ionic Radii in Aqueous Solutions Chemical Reviews, 1988, Vol. 88, No. 8 1485

TABLE V. Mg-0, Ca-0, Sr-0, and B a - 0 Internuclear


Distances in Aqueous Solutions
salt W/salt dio,,-w/nm n method ref
MdII) 0.2066 X 98
Mi(HiP04)2 34.0 0.211 6 X 99”
MeCL 50 0.200 6 MD 100
MgCl; 50 0.212 X 101
MgSO4 20.5 0.2094 6 X 102’
Mg(CHsCOz)z 0.2094 6 X 103’
MgCh 27.1-55.5 0.211 6 x 104
Mg(N0S)Z 10.8-24.8 0.211 6 X 105
I I
MgCl2 9.8-13.0 0.21 8.1-7.9 X 106
Mg(BF4)Z 14-37
25
0.215
0.2
6.4-6.2 X
6 X
107
108
1
0 ooo
4 25 100 2 25
1
MgClz
MgClz 25 0.2044 X 109 H,O /LI+

CaC1, 11.1-50.0 0.239-0.246 N 46b Figure 1. Li-0 internuclear distance, dionqab, (in nm), as a
Ca(I1) 0.2330 X 98 function of the ratio of water molecules to Li+ ions in solution.
CaCl, 10.6, 17.0 0.24 8.0, 8.2 x 106 Empty symbols, coordination number 6; filled symbols, coordi-
CaCl, 25 0.2428 X 109 nation number 4; circles, X-ray diffraction; triangles, neutron
CaC1, 50 0.239 9.2, 6.9 MD, X 110 diffraction; squares, molecular dynamics simulation; diamonds,
CaC1, 12.3-55.8 0.241-0.242 6 X 111 Monte Carlo simulations.
CaCl, 12.4 0.242 5.5 N 112
CaBr, 26.0-44.1 0.2444.240 6 X 113
SrClp 21.5 0.26 7.9 x 106 and values reported to only two significant figures are
SrCl, 26.5-34.6 0.264 8 X 114’ given a weight of 0.2, then the weighted average of the
BaClz 36 0.29 9.5 x 106 internuclear distances is 0.208 nm, with a standard
deviation from the mean of f0.006 nm. This value
’Assumed n. dinn-wis not monotonous with the concentration. should be applicable to dilute aqueous solutions of
lithium salts.
contain the following information. Under the heading Sodium. The internuclear distances Na-0 have also
“salt” is given the salt upon solutions of which the been determined in many studies, by both simulation
diffraction measurements or the ion upon which the and diffraction methods, and a consistent value of
computer simulation studies have been made. The 0.2356 f 0.0060 nm has been found; see Table 11. The
column “W/salt” presents the number of water mole- number of nearest neighbors is 6 in most of these
cules per unit formula of salt (or individual ion) present studies, although larger values (7 or 8) and a smaller
in the solution, or the range of such numbers. The one (4) have also been employed in the interpretation
value (or the range of values) of dion-water
in nanometers of the data. No effect of the salt concentration on the
is given in the next column, as found for the first peak distance can be discerned in the data.
in the pair correlation function gion-water(r). Under the Potassium. The internuclear distances K-0 have
heading n is given the coordination number found for been determined in several studies, mainly by simula-
the same peak, or assumed for it (mainly in the case of tion methods but also by diffraction methods. A con-
X-ray diffraction measurements). The next column sistent value of 0.27g8 f 0.0081 nm has been found; see
gives the method employed: X is X-ray diffraction, N Table 111. The number of nearest neighbors employed
is neutron diffraction, MD is molecular dynamics, and in the interpretation of the data is between 6 and 8. No
MC is Monte Carlo computer simulation, while the effect of the salt concentration on the distance can be
other entries are self-explanatory, Some specific com- discerned in the data.
ments for individual determinations are given at the Rubidium. In a single study by molecular dynamics
end of each table. General comments for each ion are computer simulation of a system consisting of 1Rb+ ion
given in the following. and 8 water molecules, a cluster of the rubidium ion
with 5 water molecules around it was found. The Rb-0
A. Cations distance was given as 0.15 f 0.01 nm + the radius of
i. Univalent Cations a water molecule.43
Cesium. Computer simulations by molecular dy-
Lithium. A large number of studies have been de- namics yielded Cs-0 internuclear distances of 0.31-0.32
voted to the structure of lithium salts in aqueous so- nm with 7-8 nearest neighbors, and an X-ray diffraction
lutions, unfortunately with conflicting results. A group study yielded similar results. A neutron diffraction
of these studies lead to Li-0 internuclear distances of study, on the other hand, yielded the shorter distance
0.19 nm and nearest-neighbor numbers of around 4,and of 0.295 nm (with 8 nearest neighbors). The average
the other leads to distances of 0.21 nm and around 6 of the data (disregarding the short value) is 0.3139 f
nearest neighbors. This grouping is independent of the 0.0076 nm; see Table 111.
method used to arrive at the results, but it is not com- Silver. Diffraction methods yielded consistent re-
pletely independent of the ratio of water molecules to sults, whether the number of nearest neighbors was
ions in the solution. A partly systematic dependence taken as 4, as 2, or some value in between. The Ag-0
can be discerned in Figure 1, where the distance (and internuclear distance is 0.2417 f 0.0021 nm; see Table
the number of nearest neighbors) seems to shrink as 111.
this ratio becomes smaller. The data are collected in Thallium(1). An X-ray diffraction study of con-
Table I. If the values for water/salt ratios >10 are given centrated (3.5 or 10.8 mol dm-3) aqueous thallium(1)
unit weight, those for ratios <10 are given half weight, formate showed the T1 atoms to be com-
1486 Chemical Reviews, 1988, Vol. 88, No. 8 Marcus

TABLE VI. Mn-0, Fe-0, Co-0, and Ni-0 Internuclear Distances in Aqueous Solutions
salt W/salt dion-wlnm n method ref
Mn(CH3COzh 0.2195 6 X 103'
MnSO, 20-100 0.22 6.0 X 115
Mn(I1) 500 0.217 6 X, EXAFS 116
MnS0, 13.2-26.8 0.2218-0.2195 5.25 X 117'ab
Mn(NO3h 11 0.2192 5.4 X lMaB
MnSO, 0.2196 6 X 119'
Mn(C104)2 21.2 0.220 6 X 120
MnC1, 9.7, 7.3 0.2184 4.65, 4.45 X 121b
MnBrz 7.7, 7.15 0.2181 5.0, 4.8 X 1216
Fe(I1) 500 0.210 6 X, EXAFS 116
Fe(C104)2 25.5 0.212 6 X 120
FeBrz 10.2-17.8 0.2122 5.1-5.5 X 1226
COCl, 14.8 0.21 6 X 27
coc1, 7.5 0.2099 5.5 X 58b*c
coc1, 8-1 7 0.2091-0.2085 5.6-6.2 X 59b7e
COCl, 0.214 5 X 95b
Co(I1) 0.2091 X 98
Co(CH3COz)z 0.2140 6 X 103'
Co(I1) 500 0.209 6 X, EXAFS 116'
CO(C104)2 21.0 0.208 6 X 120'
CoBr, 11.0-18.3 0.2101-0.2086 5.3-6.0 X 1226
coc1, 17.9 0.210 5.5 X 1236
CO(c104)z 0.209 6.5 X 124
CoBr, 17 0.211 5.9 X 125
CoSO, 25-100 0.215 6.15-6.32 X 126
Ni2+ 64 0.217 8.0 MD 47
Ni(I1) 0.2065 X 98
NiS0, 27.5 0.2063 6 X 102'
NiCl, 25 0.21 6 X 108'
Ni(I1) 500 0.206 6 X, EXAFS 116'
NiSO, 0.2068 6 X 119'
Ni(C10& 22.7 0.204 6 X 120'
Ni(C104)2 0.205 6 X 123'
NiCl, 17.9 0.207 5.5 X 124b
NiSO, 25-100 0.215 6.15-6.32 X 126
NiBr, 25.5 0.2065 5.1 X 127b
NiC1, 11.5 0.207 5.8 N 128d
NiCl, 11.3-580 0.207-0.210 5.8-6.8 N 129
Ni(H2P0A2 15.2 0.203 5.25 X 13CFb8
Ni(N0J2 13-110 0.206 6.5 EXAFS 131
NiC1, 12.6-27.4 0.205 6 X 132
NiCl, 12.6-645 0.205-0.210 5.8-6.8 N 133'
Ni(C104), 14.6 0.207 5.8 N 134
NiBrz 11.9-24.7 0.2079-0.2066 5.65 X 135b
NiBr, 9.6-26.5 0.204 5.4-5.7 X 136
NiS04 27 0.2059 X 137d
NiCl, 14-18 0.207 6 EXAFS 138
Ni(N03), 55 0.205 EXAFS 139
NiCl, 12.6 0.205 5.8 N 141
'Assumed n. Some ligand ions present in the inner coordination sphere beside the water molecules. LiCl also present in the solution.
No ligand ions found in the inner coordination sphere. e H3P04also present in the solution. fdi,,_w does not vary monotonously with the
concentration.

plexed by the formate (as a tetramer at the higher nearest neighbors at the very short distance of 0.244 nm
concentration) but not primarily hydrated. Therefore and a considerably farther neighbor at 0.290 nm re-
no T1-O(water) distances can be derived from this, the sulted from a recnt interpretation of X-ray diffraction
only study of aqueous Tl(1). data. This result would explain the discrepancies be-
Hydronium. According to X-ray diffraction results, tween the results of earlier work. In dilute solutions
the internuclear distance between the central oxygen the value 0,27& f 0.001, nm for the 0-0distance seems
atom of the H30+ grouping and the oxygen atom of a to be the best estimate.
nearest neighbor water molecule (one of assumed four) Ammonium. No agreement exists between the re-
is practically the same as between the oxygen atoms of sults of molecular dynamics computer simulation and
water molecules in bulk water. Neutron diffraction neutron diffraction results (Table IV). The number of
results, however, yield a higher as well as a lower value nearest neighbors given, 8 or <lo, respectively for the
(Table IV). The discrepancy cannot be ascribed to the two approaches, seems to be high, and the N-0 inter-
isotopic composition of the hydrogen atoms (deuterium nuclear distance obtained from neutron diffraction is
is used in the neutron diffraction studies), but the high a very low lower limit only (0.260 nm). The value from
concentrations employed in one study,44solutions with the computer simulation, on the other hand, seems to
down to 3 mol of water per mol of acid, could be re- be too high. No satisfactory estimate arises from these
sponsible for the low value. The presence of three data.
Ionic Radii in Aqueous Solutions Chemical Reviews, 1988, Voi. 88, No. 8 1487

TABLE VII. Cu-0, Zn-0, Cd-0, Hg-0, and Sn-0 TABLE VIII. A1-0, Y-0, Cr-0, Fe(II1)-0, Rh-0, In-0,
Internuclear Distances in Aqueous Solutions Tl(II1)-0, and Th-0 Internuclear Distances in Aqueous
salt W/salt di,+/nm n method ref Solutions
salt W/salt dio,,-w/nm n method ref
cuso, 48, 100 0.215 eq' 5.2 X 115
0.25 ax' AlCl, 23.8-54.0 0.188-0.190 6 X 166'
C~(C104)z 15.7 0.194 eq 4 eq X 120 Al(NO3)3 14.5 0.187 6 X 167'
0.238 ax 2 a i Al(NO3)3 100 0.190 6 X 168"
CUCl, 17.9 0.195 2.8 ea X 124'
CUCl2 11.6 0.205 eq 2.3 e4 N 142 Y(C104)~ 12.4 0.2365 8.0 X 169'
0.25 ax 2 or Ob ax Y(C104):, 18-55 0.2368 X 170
Cu(N03)S 23.5-34.7 0.2002 eq X 143' Cr(II1) 500 0.200 6 X,EXAFS 116
0.222 ax Cr(N03), 100 0.198 6 X 168'
cuc1, 12.9 0.205 2.3 eq N 144b CrC13 23.4 0.1995 6 X 171"~~
CuBr2 >lo0 0.193 7 EXAFS 145O Cr(S04)1,5 18.7 0.194 5.1 X 172'-
CuBr2 12-100 0.193 3.5 EXAFS 146O CrCl, 64.8 0.199 6 X 173'
CuBrp 10.8-53.1 0.197 eq 2.5-3.7 eq X 147c Cr(NO& 24.5-50.8 0.200 6 X 174'
0.250 ax 2 ax CrC13 223 0.190 6 X 175'
cuso4 75, 160 0.21 eq 4.1 eq X 148 CrCl, 17.9, 26.5 0.1982, 0.1971 6 X 176'
0.23 ax 2.2 ax
Cu(C104)2 11.4 0.194 eq 4 eq X 149 Fe(II1) 500 0.202 6 X, EXAFS 116
0.243 ax 2 a x Fe(N03), 6.6-32.5 0.203-0.205 6 X 178
CuBr2 >loo0 0.197 EXAFS 151 FeCl, 6.7-21.5 0.205-0.208 2.7-4.2 X 179'*'
Fe(S04)1,S 14.4-20.0 0.2017 6 X 180'
ZnBr2 8.1-15.9 0.221 2.4 X, EL 26' Fe(C104) 21.3-29.7 0.2008 6 X 18laSb
Zn(I1) 0.2093 X 98 FeCl, 0.207 4 X 182'
Z~(CHSCOZ)~ 0.2076 6 X 103d
ZnS04 31-100 0.215 6 X 115 Rh(C104)3 38.5 0.204 6.3 X 946
Zn(C104)2 19.4 0.208 6 X 120 Rh(ClOJ3 278 0.2062 6 X 96
ZnBrz 7-610 0.194 2.5-7 EXAFS 145' Rh(N03)a 61, 145 0.2023 X 183d
ZnS04 27.1 0.2107 6 X 152a*e
ZnS04 18.6 0.210 6 X 153d I ~ ( s o ~ ) 14.2
~,~ 0.2156 6 X 184'~'
Zn(N03h 9.0 0.217 6.6 X 154 In(C104)3 11.3 0.215 6 X 185
ZnCl, 5.4 0.205 X 15Ef Tl(ClOJ3 17.1-42.3 0.2236 5 X 187
ZnS04 25 0.2100 6 X 156d Tl(C104)~ 0.2227 6 X 188
Cd(I1) 0.2289
0.2287 6
X
X
98
103d
Th(NO3)I 57.0 0.255 5.5 x 18gC
Cd(CH&Oz)z Th(NO3)4 22.3 0.251 X 190
Cd(HZPO4)2 10.4 0.230 5.1 X 13Od9
Cd(C104)2 23.0 0.229 6 X 157d8 'Assumed a. bHC1, HC104, or HzSO4 also present in the solu-
CdSO4 17.6 0.229 5.2 X 157'4 tion.' cSome ligand ions also present in the inner coordination
CdSO4 25 0.231 6 X 158dh sphere. Dimeric species, with OH- and NO3- bridges.
Cd(NO& 9-54 0.2272 5.7 X 159
CdClz 35-63 0.237 4 X 160'
Cd(C104)2 26.6 0.231 6 X 161d TABLE IX. La-0, Ce-0, Pr-0, Nd-0, Sm-0, Eu-0, Gd-0,
CdBr2 13.2-662 0.218 3 EXAFS 162' Tb-0, Dy-0, Er-0, Tm-0, and Lu-0 Internuclear
Distances in Aqueous Solutions
Hg(BF4)z 27 0.233 6 X 163d salt W/salt di,,w/nm n method ref
Hg(C104)Z 19.7 0.234 6 X 164'
Hg(C10&2 19.7 0.241 6.0 X 165 La(ClO,L 12.0 0.2570 8.0 X 169'
20.8-31.9 0.248 8.0 X 190
Sn(C104)2 16.8 0.234 3.5 X 48i 14.6-27.8 0.258 9.1 X 191
0.221 3.8 EXAFS 20.9 0.248 8.0 X 192
Sn(C10,)2 12.0 0.233 2.4 X 15@ 25.4 0.2552 7.5 X 193
0.283 3.0 14.6 0.254 9.2 X 191
16.5 0.251 8.9 X 191
a Equatorial and axial water molecules in the Jahn-Teller dis-
17.5 0.248 8.5 N 194
torted octahedral inner coordination sphere. *If n = 0 then there 17.5 0.2479 8.4 N 195
are 4 C1- ions in the equatorial positions. 'Some ligand ions pres- 32.1-99.1 0.241 8.0 X 196
ent in the inner coordination sphere. dAssumed n. 'No ligand 15.2 0.2455 8.0 X 16ga
ions in the inner coordination sphere. 'Tetrahedral species ZnC12-
(H20),. gH,P04 or HC104also present in the solution. Data also
38.8 0.242 9.9 X 197
17.2 0.247 8.8 X 198
at 9 and 62 OC. 'Data also for hydrolyzed Hg(I1). 'Another model, 17.2 0.245 8.3 X 198
with additional 3.0 H20 at 0.29 nm also compatible with the data. 26.8 0.240 9.9 X 197
Both long and short distances present. 14.4-45.1 8.0 X
0.237 196, 199
13.7 0.2400 8.0 X 169'
Methylammonium. A single study by Monte Carlo 18-55 0.240 X 170
computer simulation, in a system with 216 water mol- 11.0 0.241 8.2 X 200
11.7 0.240 7.9 X 200
ecules and l CH3NH3*ion, led to 3.5 nearest neighbors, 21.0 0.2370 7.4 N 201b
on the average, with a C-O internuclear distance of 0.37 12.1 0.2360 8.0 X 169'
nm or a (C)H-0 distance of 0.185 nm.42 18-55 0.236 X 170
ErC13 10.8 0.237 8.2 X 200
2, Divalent Cations ErCl,, Er13 4.6-15 0.23 6.3-6.5 X 202
TmC13 10.6 0.236 8.1 X 200
Beryllium. A study by Yamaguchi et d.,&involving LUCl, 10.6 0.234 8.0 X 200
both molecular dynamics computer simulation and 'Assumed n. No C1- in inner coordination sphere.
X-ray diffraction, yielded conflicting results. The for-
mer method produced an octahedral arrangement of six ryllium, with a B e 0 internuclear distance of 0.175 nm,
water molecules in the first hydration shell of the be- as well as an indication of the presence of a second
1488 Chemical Reviews, 1988,Vol. 88, No. 8 Marcus

TABLE X. F-0, Br-0, and 1-0 Internuclear Distances in TABLE XI. C1-0 Internuclear Distances in Aqueous
Aqueous Solutions Solutions
salt W/salt dio,-w/nm n method ref salt Wlsalt ..,
di,-wlnm
.I"
n method ref
CsF 25 0.264 6.8 MD 51 Cl- 64 0.323 5.9 MD 47
F- 215 0.260 4.1 MC 51 LiCl 3-8.2 0.318-0.310 6 X 490
F- 64, 125 0.267 5.8 MD 52 c1- 64, 125 0.329 7.2 MD 52
F- 215 0.263 5 MD 53 Cl- 215 0.348 8.5 MD 53
F- 215 0.260 6.3 MC 60 LiCl 4.0 0.320 MD 54b
CsF 25 0.322 8 MD 68 CoCl, + LiCl 7.5 0.316 6 X 58'
F- 10 0.275 4-6 theory 70 CoCl, + LiCl 8-17 0.3156 6.0-6.4 X 59
KF 13.3, 27 0.262 X 82 c1- 215 0.325 8.4 MC 60
CsF 25 0.264 6.8 MD 84 LiCl 4.0 0.322 MD 62
NH,F 3.6 0.287 6 X 203 LiCl 4.0 0.319, 0.315 X,MD 63
LiCl 5.6-15.6 0.329-0.334 5.3-5.9 N 64
ZnBr, 8.1-15.9 0.345 4.2 X 26 LiCl 13.9 0.308 6 X 65
LiBr 25.0 0.329 6 X 67 NaC1, KC1 13.9-27.8 0.308-0.316 6 X 65
LiBr 5.6-22.2 0.343-0.337 7.2-8.9 X 86 LiCl 54.3 0.310 6 N 66
DBr 51.7 0.321 6 N 92' LiCl 25 0.27 7 MD 68
CaBr2 26.0-44.1 0.334-0.332 6 X 113 c1- 10 0.345 6-7 theory 70
NiBr, 25.5 0.312 5.15 X 127b NaCl, KCl 54.3 0.310 N 72
ZnBr, 15-150 0.340 4 EXAFS 145 NaCl 10.2 0.316 6 X 73
NHIBr 7.6 0.336 6 X 203 NaCl 100 0.27 8 MD 74
LiI 25 0.368 8.7 MD 55 NaCl 25 0.32 6 MD 75
LiI 25 0.370 MD 56 NaCl 25 0.330 8 MD 76
LiI 25.2 0.363 6.9 X 57 LiCl 8 0.324 X 80
LiI 25 0.30 7 MD 68 NHdCl 25 0.322 8.2 MD 85
NaI '7 0.360 6 X 77 LiCl 5.6-22.2 0.315-0.320 6.2-7.3 X 86
LiI 11.1-22.2 0.365-0.369 8.8-8.9 X 86 DC1 54.9 0.310 6 N 92
NaI 5.6 0.376 X 86 coc1, 0.3120 5.06 X 95
NHJ 8.2 0.361 6 X 203 RhC1, + HCl 0.3185 X 96
MgClz 50 0.318 7.0 MD 100
Also n = 4 and n = 8 are compatible with the data. Some of MgCl2 50 0.316 X 101
the Br- is in the inner sDhere of the Ni(I1). MgClz 27.1-55.5 0.314 6 X 104
MgCl2 25 0.3130 X 109
hydration shell, at a mean distance of 0.373 nm from CaCl, 50 0.319, 0.312 7.9, 8.2 MD, X 110
the beryllium atom. The other method indicated the CaCl, 12.3-55.8 0.315 6 X 111
CaC1, 12.4 0.325 5.8 N 112
presence of only four water molecules in the first hy- SrCl, 26.5-34.6 0.323 X 115
dration shell, at a Be-0 distance of 0.167 nm, but this coc1, 17.9 0.311 6 X 123c
at a lower number of moles of water per mole of salt NiCl, 17.9 0.306 6 X 124
(lo), compared with the ratio in the former study (50). NiC1, 11.5 0.320 5.7 N 127d
LiCl 3.4-14.0 0.334-0.325 5.9-4.4 N 129
The number of water molecules near the chloride anions NiC1, 12.6-27.4 0.313 6 X 132
at the higher concentration, 3.4,was also much lower NaCl 9.4 0.320 5.5 N 141
than in the more dilute solution (7), indicating the AlC1, 23.8-54.0 0.311-0.314 6 X 166
formation of contact ion pairs in the concentrated one. CrC1, 23.4 0.3135 X 171e
This could explain the smaller number of water mole- FeC1, 6.7-21.5 0.324-0.3156 6 X 179'
LiCl 3 0.320 4 MD 186
cules, and perhaps also the shorter distance. On the NHiCl 8.5 0.317 6 X 203
other hand, a coordination number of 6 for the beryl- LiCl 3.4 0.325 4 N 204b
lium in the dilute solution is surprising, 4 being the HCl 8-27 0.320-0.325 X 205
expected number. It is evident that this system merits HC1 4-96 0.313 4 X, N 206
NaCl, RbCl 10.4 0.320 5.5 N 207
further study before a firm conclusion can be reached. BaC1, 50.4 0.326 6.2 X 207
Magnesium. The results of several studies (Table KCl 14.9 0.316 208
V) by X-ray diffraction are consistent. With the as- NdC13 17.5 0.329 3.9 N 209'
sumption of a nearest-neighbor number of 6, the av- 'Assumed n. Direct interaction between Li+ and C1-. e Some of
erage Mg-0 internuclear distance is 0.20g0f 0.0041 nm. the C1- present in the inner coordination sphere of the metal ion.
A molecular dynamics computer simulation study dNone of the C1- is in the inner coordination sphere of the metal
yielded, with the same number of neighbors, the ion. e Some HC1 also present. f Some of the water shared between
somewhat lower value of 0.200 nm. the Nd3+ and C1- ions.
Calcium. Calcium salt solutions have been studied
by molecular dynamics computer simulation as well as an assumed 8 nearest water molecule neighbors of the
by neutron and X-ray diffraction methods, with gen- strontium ion. An earlier study reported 7.9 for the
erally consistent results concerning the Ca-0 internu- number of neighbors and estimated the internuclear
clear distance (Table V): 0.2422f o.0()52 nm. However, distance at 0.26 nm (Table V).
no agreement concerning the number of nearest Barium. An early and uncorroborated study re-
neighbors resulted, the reported values ranging from 5.5 ported 9.5 for the number of neighbors and estimated
to 10.0, possibly depending on the concentration the internuclear distance at 0.29 nm (Table V).
(Hewish et al.46).A recent molecular dynamics study, Manganese. Several X-ray diffraction studies of
however, reported a considerably larger Ca-0 distance, manganese salt solutions yielded consistent results. In
0.254 nm in a system with 64 water molecules per Ca2+ some of these studies, contact ion pairs with the anions
ion.47 were disregarded, and the number of nearest water
Strontium. An X-ray diffraction study reported the neighbors was assumed to equal the coordination num-
Sr-0 internuclear distance as 0.264 f 0.004 nm with ber of 6. In others, ion pairing was explicitly taken into
Ionic Radii in Aqueous Solutions Chemical Reviews, 1988, Vol. 88, No. 8 1489

TABLE XII. N(OS)-0, Cl(O,)-O, (H,)P(O,)-0, C(HSC02)-0,S(Od)-O, Se(04)-0, and Mo(0,)- or W ( 0 4 ) - 0 Internuclear
Distances in Aqueous Solutions
salt W/salt dion-wlnm n method ref
6.1, 9.3 0.318 3 or 6 X 78
14.2 0.317 4.3 X 88
4.3-16.6 0.313 1 X 91
11 0.344 6.3 X 1180
9.0 0.344 17.7 X 154
11-54 0.349 8.8 X 159
14.5 0.339 X 167
61.1, 144.9 0.350 8.2 X 183
25.4 0.340 7.0 X 193
6.4 0.265 ax 1.3 ax N 211
0.340 eq 2.4 eq
4.1 0.351 9 X 212
10.4 0.3570 25.6 X 88
16.3 0.352 1 X 91
278 0.376 12 X 96
12.0-15.2 0.380 8 X 16gbsm
8.9-20.1 0.380-0.367 X 181
17.1 0.37 4-5 N 211
34.0 0.387 8.8 X 99'
15.2 0.375 4.2 X 13OCsd
10.4 0.391 4.4 X 130Csd
6-18 0.306 MD 212e
7.7-25.0 0.360-0.373 4 or 8 X 213af
0.37 MC 29
0.36-0.37 4-6 X 1038
0.36-0.37 4-6 X 1038
0.36-0.37 4-6 X 1038
0.36-0.37 4-6 X 1038
0.35 4-6 X 1038
30.5 0.367 8 X 81b
20.5 0.370 7.7 X 102
27.5 0.381 8.1 X 102
13.2-26.8 0.3905 7.4 X 117h
0.382 8.0 X 119h
18, 27 0.380 8.2 X 137k
27.1 0.387 7.0 X 152'
18.6 0.383 8.2 X 153
25 0.387 6.6, 7.3 X 156
25 0.389 6.9 X 158hJ
18.7 0.381 7.8 X 172h
9.6-13.4 0.372-0.377 18-20 X 180
14.7 0.389 6.4 X 184h
11.7 0.379 7.6 X 214
24.5-46.4 0.395 8 X 169b.C.l
25 0.406 12 X 215
25 0.406 12 X 215
Some of the NO, is in the inner coordination sphere of the metal ion. bAssumed n. 'Some of the H2P04-is in inner coordination sphere
of the metal ion. dH3P04is also present in the solution. eSmall cluster. 'Two models are compatible with the data. #Some of the C H 3 C 0 ~
is in the inner coordination sphere of the metal ion. Some of the S042- is in the inner coordination sphere of the metal ion. 'None of the
SO?- is in the inner coordination sphere of the metal ion. 'Data also at 9 and 62 OC. kData also for solutions containing added Li2S04.
'Some Se02- is in the inner coordination sphere of the metal ion. mLn = La, Sm, Tb, Er, or Y.

account, as a result of the finding of the characteristic others, ion pairing was explicitly taken into account, as
cation-anion distances in the solution, so that the a result of the finding of the characteristic cation-anion
number of water molecules was lower. Still, the Mn-0 distances in the solution, so that the number of water
internuclear distances found in all the studies are nearly molecules was lower. Still, the Co-0 internuclear dis-
the same, 0.21g2f 0.0Ol3 nm, on the average; see Table tances found in all the studies are nearly the same,
VI. 0.2106 f 0.0022 nm, on the average; see Table VI.
Iron(I1). The X-ray studies of solutions of iron(I1) Nickel. Many studies have been concerned with
salts all yielded the Fe-0 internuclear distance near the nickel ions in aqueous solutions, comprising X-ray and
average of 0.211, f O.0Ol0 nm with an assumed 6 nearest neutron diffraction and EXAFS, as well as molecular
neighbors of either water molecules or bromide anions dynamics computer simulation. In the X-ray diffraction
and water molecules; see Table VI. studies, the number of nearest neighbors was assumed
Cobalt. Several X-ray diffraction studies of cobalt to be 6, whether they were only water molecules or both
salt solutions yielded consistent results. In some of these and anions, but the isotope-differential neutron
these studies, contact ion pairs with the anions were diffraction studies produced 5.8 nearest water molecules
disregarded, and the number of nearest water neighbors at the highest concentrations (about 13 mol of water per
was assumed to equal the coordination number of 6. In mol of salt), which increased to 6.8 nearest neighbors
1490 Chemical Reviews, 1988, Vol. 88, No. 8 Marcus

in the more dilute solutions (near 600 mol per mol). hydrolyzed Hg2+aqueous ion. For the latter, the Hg-0
The molecular dynamics study, however, yielded the distance is 0.242 nm, with 6.0 f 0.4 nearest water
value of 8 neighbors (in a system of 64 water molecules molecules; see Table VII.
and 1Ni2+ion), which is excessive relative to the known Tin(I1). A study of a 3.3 mol dm-3 aqueous solution
coordination number of 6. In all cases, however, the of tin(I1) perchlorate by X-ray diffraction and EXAFS
Ni-0 internuclear distance was near to the average of measurements was made by Yamaguchi et al.48 The
0.2061 f 0.0014 nm; see Table VI. X-ray measurements yielded 0.235 nm for the Sn-0
Copper(I1). It is impossible to assign a unique Cu-0 distance with 3.5 f 0.1 nearest water molecules, but a
internuclear distance to ions of Cu2+in aqueous solu- weaker interaction with a further (assumed) 3 water
tions because of the Jahn-Teller distortion of its octa- molecules at 0.29 nm was also indicated. A second
hedral first coordination shell. There are four shorter coordination shell of 7.2 water molecules was found at
equatorial distances and two longer axial ones. A fur- 0.44 nm, so that the 0.29-nm distance should be as-
ther complication is the propensity of copper(I1) ions cribed to water molecules in the first coordination shell.
to be complexed by halide and nitrate anions in the The EXAFS study yielded a shorter distance, 0.227 nm,
relatively concentrated solutions required for the dif- for 3.1 nearest water molecules and an even shorter one,
fraction studies. When the appropriate coordination 0.221 nm, for another 1.1water molecules, for an alto-
numbers of 4 near and 2 far water neighbors are as- gether occupation of about 4 water molecules in the first
sumed, then X-ray diffraction of copper(I1)perchlorate coordination shell. In another study, an average of 2.4
solutions indeed yields definite Cu-0 distances. These shorter internuclear distances of 0.233 nm and another
agree well with the averages obtained for these distances 3.0 longer distances of 0.283 nm is in agreement with
to the water oxygen atoms in chloride, bromide, and the above. The weighted mean distance of 0.262 is only
nitrate solutions, notwithstanding that some of the a rough indication of the mean distance of the water
other coordination sites are occupied by atoms from the molecules from the tin(I1) ion.
anions. These averages are 0.1968 f 0.0047nm for the
equatorial oxygen atoms and 0.240 f 0.010 nm for the 3. Trivalent Cations
axial ones. The results from neutron diffraction and
from EXAFS are on the high and low sides of the av- Aluminum. Three X-ray diffraction studies gave
erage of the X-ray diffraction results, but within the consistent results (see Table VIII): with assumed 6
standard deviation; see Table VII. nearest water molecule neighbors the Al-0 internuclear
distance was 0.18a7 f 0.001, nm.
Zinc. A number of X-ray diffraction studies yielded Yttrium. An X-ray diffraction study (Table VIII)
consistent values for the Zn-0 internuclear distance in on perchlorate solutions yielded with an assumed 8
aqueous zinc salt solutions, where the number of nearest nearest water neighbors a Y-O internuclear distance of
neighbors was assumed to be 6. In some cases (the 0.2365nm. In selenate solutions this distance was ap-
bromide and sulfate; see Table VII) inner-sphere com- parently somewhat shorter, 0.233 nm.
plexing takes place, but in the case of the sulfate this Lanthanides. X-ray diffraction studies of aqueous
does not affect appreciably the distance to the water solutions of salts of most of the lanthanides (except
molecules. The average distance is 0.209* i 0.0066 nm. promethium, holmium, and ytterbium) and neutron
Longer and shorter Zn-0 distances resulted from the diffraction studies of aqueous neodymium and dys-
interpretation of the data on zinc bromide solutions by prosium chlorides (Table IX) yielded M-0 distances
X-ray diffraction and EXAFS, respectively, when al- and in some cases numbers of nearest-neighbor water
lowance was made for the presence of bromide ions in molecules of M, where M = La, Ce, Pr, Nd, Sm, Eu, Gd,
the inner coordination shell. Tb, Dy, Er, Tm, and Lu. A downward trend of the
Cadmium. Several X-ray diffraction studies yielded number of water molecules in the first coordination
consistent values for the Cd-0 internuclear distance in shell, along with a shrinking of the M-0 distance can
aqueous cadmium salt solutions, where the number of be discerned in Table IX, but the data are not suffi-
nearest neighbors was assumed to be 6. In the case of ciently in agreement to indicate whether the decrease
the nitrate the number of nearest water neighbors found in the number of neighbors is gradual or abrupt, as has
was only 5.7, but no indication of the presence of the been claimed (see ref 202a).
nitrate anion in the inner coordination shell was given. Chromium. Several X-ray diffraction studies of
On the contrary, phosphate and sulfate ions were found various chromium(I1I) salt solutions gave consistent
in this shell in other studies (Table VII), but their results, with an assumed number of 6 nearest water
presence did not affect the Cd-0 distance appreciably. molecule neighbors, for the Cr-O internuclear distance
The average value of this distance is 0.2301f 0.002, nm. of 0.1969f 0.0032nm. An EXAFS study gave concor-
Only in the case of cadmium chloride solutions is there dant results; see Table VIII.
an indication of a decrease of the number of nearest Iron(II1). Several X-ray diffraction studies of var-
water molecules to 4, when 2 chloride anions occupy ious iron(II1) salt solutions gave consistent results for
sites in the inner shell, with a concomitant increase of the Fe-0 internuclear distance of 0.2031 f O.0Ol9 nm,
the Cd-0 distance. with in one case 5.8 + 0.2 and in others an assumed
Mercury(I1). Mercury(I1) was one of the earliest number of 6 nearest water molecule neighbors. An
ions for which a cation-water distance was estimated EXAFS study gave concordant results; see Table VIII.
from X-ray diffraction measurements on aqueous so- In chloride solutions some chlorine atoms occupied sites
lutions. This early value was later confirmed, but still in the first coordination shell, and then the Fe-0 dis-
later revised when it was realized that the relatively tance was somewhat longer (Table VIII).
short distance (0.233 nm) was characteristic for hy- Rhodium(II1). An X-ray diffraction study of
drolyzed mercury(I1) ions, rather than for the un- aqueous rhodium(II1) nitrate solutions showed the
Ionic Radii in Aqueous Solutions Chemical Reviews, 1988, Voi. 88, No. 8 1491

presence of dimeric species with OH and NO3 bridges ions are in the fiit coordination shells of the metal ions.
and Rh-0 distances of 0.202,, and 0.2026 nm to the Iodide. The internuclear distance 1-0 in aqueous
water molecules in the remaining sites of the first co- solutions was studied by both X-ray diffraction and
ordination shell. A similar distance, 0.204 nm, was molecular dynamics computer simulation, with gener-
found for the Rh-0 distance with 6.3 nearest neighbors ally consistent results, giving an average of 0.3647 z t
in acidified perchlorate solutions; see Table VIII. 0.0036 nm, with an assumed 6 or 7 nearest water mol-
Indium. Three X-ray diffraction studies on indium ecule neighbors in the X-ray diffraction studies and 8.7
sulfate and perchlorate solutions gave consistent results nearest neighbors in the computer simulation. One
of 0.2156 and 0.215 nm for the In-0 internuclear dis- outlying very short distance, 0.30 nm, obtained in a
tance to the (assumed) 6 nearest water molecules, al- molecular dynamics study (see Table X) is difficult to
though in the sulfate case there was an indication of the explain.
presence of sulfate in the inner coordination shell; see Nitrate. The nitrate anion, like the other oxyanions,
Table VIII. poses the difficulty of the distinction between the in-
Thallium(II1). Two X-ray diffraction studies of ternal internuclear distances and the external N-O one,
aqueous thallium(II1) perchlorate solutions yielded the relevant to the anion-to-water molecule distance, and
mean T1-0 internuclear distance 0.2231 f 0.0005 nm hence to the ionic radius in the solution. Furthermore,
with 5 f 1nearest-neighbor water molecules; see Table the nitrate anion is not spherically symmetrical, and
VIII. differences exist between the extension of the anion in
4. Tetravalent Cations the plane of its tree oxygen atoms (the equatorial ra-
dius) and perpendicular to it (the axial radius). Still,
Thorium. In two studies of aqueous thorium nitrate a fairly consistent picture arises from the various X-ray
solutions the main interest was the structure of the diffraction studies and the single neutron diffraction
hydrolyzed species, but the results for acidified solu- study; see Table XII. The equatorial N-O(water) dis-
tions, where no hydrolysis took place, were also re- tance is 0.3451 f 0.0043 nm, on the average, whereas the
ported. According to these studies, 3.5 nitrate anions axial water molecule can approach the nitrogen atom
are present as bidentate ligands in the inner coordina- to within 0.265 nm, so that the overall mean N-O-
tion sphere of the thorium, but 5.5 water molecules (water) distance is 0.3160 f 0.0022 nm. It is, thus, im-
complete the inner coordination shell of about 12. The possible to assign the nitrate anion a “fixed” radius in
Th-O(water) internuclear distance is 0.253 nm; see aqueous solutions.
Table VIII. Perchlorate. The perchlorate anion poses less dif-
ficulties than the nitrate anion, because it is more
6. Anions symmetrical, but only a few studies, involving diffrac-
Fluoride. The internuclear distance F-0 was stud- tion methods, report the C1-O(water) internuclear
ied practically only by computer simulation methods, distance. This is larger, 0.380 nm, in aqueous solutions
and only one X-ray diffraction study was reported; see of trivalent metal perchlorates and smaller (down to
Table X. This, indeed, gave a rather high value, 0.287 0.352 nm) in solutions of univalent ones or in solutions
nm, for this distance in a very concentrated aqueous containing both iron(II1) perchlorate and perchloric
ammonium fluoride solution. So did also one of the acid. Only an imprecise value of 0.37 f 0.01 nm can
molecular dynamics computer simulation studies (0.322 be given as an average from these studies; see Table
nm, with 8 nearest neighbors in a dilute solution), but XII.
the other computer simulation studies yielded the av- Phosphate. Three studies by X-ray diffraction on
erage value of 0.2631 f 0.0025 nm, with between 4.1 and aqueous divalent metal dihydrogen phosphate solutions
6.8 nearest water molecule neighbors of the fluoride yielded values for the P-O(water) distance ranging from
anion. 0.375 to 0.391 nm (see Table XII), but then 0.75-1.0
Chloride. A large number of studies of aqueous phosphate anions, on the average, are within the first
metal chloride solutions by diffraction and computer coordination sphere of the metal cation. An X-ray
simulation methods report also the C1-0 internuclear diffraction study of aqueous phosphoric acid yielded
distance. A few computer simulation studies were de- distances ranging from 0.360 to 0.373 nm, as its con-
voted specifically to the chloride anion in aqueous so- centration decreased, according to two models that had
lutions. Apart from a few exceptions, the reported 4 or 8 water molecules as the nearest neighbors (hy-
values are within the error limits of the average 0.3M7 drogen bonded to the phosphate group). A molecular
f 0.0067 nm, but the reported number of nearest water dynamics computer simulation study yielded a very low
molecule neighbors of the chloride anion range from value for the P-O(water) distance, 0.306 nm, in a small
somewhat less than 6 to 8.5, with particularly low values phosphate-water cluster, containing 6-18 water mole-
(see Table XI) in very concentrated solutions, where cules per P043-anion.
solvent-sharing ion pairs are formed. Acetate. The acetate anion poses special difficulties,
Bromide. The internuclear distance Br-0 was due to its nonsymmetrical nature, having a hydrophilic
studied mainly by X-ray diffraction, whereas a neutron end and a hydrophobic end. Two X-ray diffraction
diffraction study and an EXAFS study gave low and studies of aqueous divalent metal acetate solutions
high values relative to the X-ray results. In the latter yielded C(methy1)-O(water) distances of 0.35-0.37 nm,
the number of nearest water molecule neighbors to the in agreement with a molecular dynamics computer
bromide anion was generally assumed to be 6; see Table simulation study (Table XII). The latter study also
X. The average distance obtained is 0.337, f 0.0054nm, gave the O(carboxy1ate)-O(water) distance as 0.27 nm.
outlying values being obtained in the cases of concen- Sulfate. Many studies of aqueous divalent and
trated solutions of ZnBrz and NiBr2, where bromide trivalent metal sulfate solutions by X-ray diffraction,
1492 Chemical Reviews, 1988, Vol. 88, No. 8 Marcus

as well as a few studies of ammonium sulfate and sul- TABLE XIII. Mean Ion-Water Internuclear Distances
furic acid solutions, yielded rather consistent results and (from Tables I-XII), Calculated Ionic Radii in Solution
a S-O(water) internuclear distance of 0.3815 f 0.0071 (Illon,eq 9), and Pauling-Type Crystal Ionic Radii, RPion
nm. The number of nearest water molecule neighbors ion d;,-wlnm R;,lnm RP;,,lnm
ranged from 6.4 to 9.6, depending on the fraction of the Li+ 0.208 f 0.007 0.071 f 0.007 0.074
sulfate anion that was in the inner coordination sphere Na+ 0.2356 f 0.0060 0.097 f 0.006 0.102
of the metal cation. Eight hydrogen-bonded nearest K+ 0.2798 f 0.0081 0.141 f 0.008 0.138
Rb+ 0.289 0.150 0.149
neighbors were assumed for the sulfuric acid solutions; cs+ 0.3139 f 0.0076 0.173 f 0.008 0.170
see Table XII. Ag+ 0.2417 f 0.0021 0.102 f 0.002 0.115
Selenate. Some aqueous trivalent lanthanide sel- H30+ 0.2755 f 0.0015 0.141 f 0.002
enate solutions were studied by X-ray diffraction, and Mg2+ 0.2090 f 0.0041 0.070 f 0.004 0.072
the SeO(water) internuclear distance of 0.395 nm, with Ca2+ 0.2422 f 0.0052 0.103 f 0.005 0.100
an assumed number of 8 water nearest neighbors of the SrZ+ 0.264 0.125 0.125
selenate anion, was found; see Table XII. Mn2+ 0.2192 f 0.0013 0.080 f 0.001 0.083
Fez+ 0.2114 f 0.0010 0.072 f 0.001 0.078
Molybdate and Tungstate. Table XI1 shows the co2+ 0.2106 f 0.0022 0.072 f 0.002 0.075
results from an X-ray diffraction study, using the iso- Ni2+ 0.2061 f 0.0014 0.067 f 0.001 0.069
morphous substitution principle, according to which a Cu2+eq 0.1968 f 0.0047
definite hydration shell exists around these anions, so Cuz+ ax 0.240 A 0.010
Cu2+mean 0.211 0.072 0.073
that a characteristic Mo-water or W-water distance can Zn2+ 0.2098 f 0.0066 0.070 f 0.007 0.075
be discerned in addition to the intraionic Mo-0 and Cd2+ 0.2301 f 0.0025 0.091 f 0.003 0.095
W-0 distances. Hg2+ 0.242 0.103 0.102
Sn2+ 0.233 0.094 0.093
C. Ionic Radii from Interparticle Distances ~13+ 0.1887 f 0.0015 0.050 f 0.002 0.053
Y3+ 0.2365 0.097 0.101
The mean ion-water distances from Tables I-XII, i.e., La3+ 0.2528 f 0.0048 0.114 f 0.005 0.118
the mean internuclear distances between the monoa- Ce3+ 0.255 0.116 0.114
tomic ions or the central atoms of polyatomic ions and Pr3+ 0.254 0.115 0.114
the oxygen atoms of the water molecules in their first Nd3+ 0.2472 f 0.0033 0.108 f 0.003 0.112
Sm3+ 0.2448 f 0.0021 0.106 f 0.002 0.109
hydration shells, are collected in Table XIII. Appli- Eu3+ 0.245 0.106 0.107
cation of eq 9 to these dion-water
values permits then the Gd3+ 0.239 0.100 0.106
evaluation of the radii of the ions in aqueous solutions, Tb3+ 0.2403 f 0.0005 0.101 f 0.001 0.104
Ri,, provided that the value of Rwateris known. A brief Dy3+ 0.2370 f 0.0003 0.098 f 0.001 0.103
discussion of this problem is given in Section III.B, Er3+ 0.2363 A 0.0005 0.097 f 0.001 0.100
Tm3+ 0.236 0.097 0.099
where some of the results for Rwaters33 are summarized. LU3+ 0.234 0.095 0.097
Further determinations of this quantity216-220 are in Cr3+ 0.1969 f 0.0032 0.058 f 0.003 0.062
essential agreement with the average values quoted Fe3+ 0.2031 f 0.0019 0.064 f 0.002 0.065
previously: Rwakr= 0.1420 f 0.0005 nm for liquid water Rh3+ 0.204 f 0.001 0.065 f 0.001
1~3+ 0.2156 f 0.0002 0.076 f 0.001 0.079
at room temperature. However, Rwaterdepends on the ~13+ 0.2231 f 0.0005 0.084 f 0.001 0.088
temperature, as was found by Narten and Levy29ain
their X-ray diffraction study. Numerical values from Th4+ 0.253 0.114 0.106
this study (the original presented only graphical results) F- 0.2630 f 0.0025 0.124 f 0.003 0.133
were quoted by Lie et al.31band Stillinger and Rah- c1- 0.3187 f 0.0067 0.180 f 0.007 0.181
man32bfor the range 4-200 "C, which can be fitted to Br- 0.3373 f 0.0054 0.198 f 0.005 0.196
the expression Rwakr= 0.1409 + 3.1 X 10-5(t/"C) nm. 1-
NO3- ax
0.3647 f 0.0036
0.265
0.225 f 0.004 0.220
This linear expression agrees, within the experimental NO3- eq 0.3451 f 0.0043
error of 0.002 nm, also with later X-ray diffraction NO3- mean 0.316 A 0.002 0.177 f 0.002 0.179
data,216neutron diffraction data,219and molecular dy- c104- 0.370 0.241 0.240
namics calculation^^^^^^^^ for temperatures from -4 to H2POC 0.377 f 0.011 0.238 f 0.011 0.238
+118 "C. s0:- 0.3815 f 0.0071 0.242 f 0.007 0.230
It can be argued, however, that the water in the first SeOd2- 0.395 0.256 0.243
Mo(W)O:- 0.406 0.267 0.254-0.270
hydration shell of an ion is under much stronger forces
than those due to the fields prevailing in liquid water. diffraction study at 25 "C up to a pressure of 5 kbar
The compressive (electrostrictive) force of the electrical (density of 1.485 g ~ m - gave ~ ) a decrease of Rwaterof
field E near an ion has been translated221(see also ref 0.025 nm relative to ambient pressure.223The electro-
4, p 104) into values of the pressure prevailing there by striction producing these densities and distances cor-
means of the differential expression responds at 25 " C to an electrical field of 2.2 X lolo V
dE, T
d P = (Co/41rK~)(d€ / ~ P ) E (21) m-l or a distance of 0.09 z1I2 nm from the center of an
ion: i.e., to a realistic situation. On this basis, the value
where KT is the isothermal compressibility of water. A of R, at room temperature near an ion would be 0.139
molecular dynamics computer simulation of water un- nm, with an estimated uncertainty of 0.002 nm.
der such a high pressure that the density of water was Application of eq 9 with this value of Rwakryields the
1.3474 g cm-3 at 68 "C yielded2mthe value Rwak, = 0.140 values of Rionlisted in the third column of Table XIII.
nm instead of the 0.143 nm at ambient pressure ex- The uncertainty given reflects that of dion-water in the
pected from the above linear expression for this tem- second column, not compounded by that of Rwater.
perature. The same result was obtained for 77 "C at These values are compared in both Table XI11 and
a density of 1.346 g cm-3 in a similar study.n2 A neutron Figure 2 with the values of EPion, the Pauling-type
Ionic Radii in Aqueous Solutions Chemical Reviews, 1988, Vol. 88, No. 8 1493

that the internuclear distances ion-hydrogen atom


I
0.301
‘ I
(deuterium atom) are known. Still, this does not permit
the apportioning of this distance between the ion and
the water molecule. The agreement between Rionand
RPionfor most of the anions (Table XIII) does indicate
that the average value of Rwater= 0.139 nm is a rea-
sonable compromise for both cations and anions.

V I . Ionic Radii in Nonaqueous Solvents


Work on the sizes of ions and solvent molecules in
nonaqueous solvents is much less extensive than the
corresponding work on water and aqueous ionic solu-
tions. One reason for this is, obviously, the greater
general importance of water and aqueous solutions.
L- Another reason is technical in nature, however, and is
oooooo 0 IO 0 20 0.30 the larger number of atoms, compared with water, in
R,: /nm the nonaqueous solvents. With this larger number,
Figure 2. Ionic radii in solution, Rim = d-& - R,.wi+ R- there are more intramolecular distances that must be
= 0.13g3nm,as a function of the Pauling-typecrystal ionic radii, accounted for in the X-ray or neutron diffraction re-
Rph, for coordination number 6 (except for Sr, Y, the lanthanides, sults, and suitable interaction models required for the
and Th, where Rpi, for coordination number 8 are used). Open computer simulations are that much more complicated.
circles, cations; filled circles, anions. Compared with the large body of information on the
structure of liquid ~ a t e r , from ~ which
~ ~ ~ ~
crystal ionic radii of the monoatomic ions for coordi- Rwaterrequired in eq 9 can be obtained, there is only a
nation number 6. These latter values were taken from limited body for nonaqueous solvents of electrolytes.
Shannon and Prewitt,20except for Sn2+,C1-, Br-, and (Structural information on so-called “inert” solvents
I-, for which the values of Ahrens2I were taken. The such as 2,2-dimethylpropane, benzene, chloroform, and
values for Sr2+,Y3+, the lanthanides, and Th4+are for carbon tetrachloride is not relevant in the present
coordination number 8.20 Values given in the column context.) Following is a brief and nonexhaustive
RPionfor polyatomic ions are thermochemical-type radii presentation of the sizes found for nonaqueous solvents
from Marcus and Loewenschuss.26 by the more recent diffraction and computer simulation
Good agreement between the Ri, and RPionvalues in methods. Except for HF and NH3, methanol and for-
the third and fourth columns of Table XI11 is noted in mamide have the simplest molecules and have received
most cases (see also Figure 2). Noteworthy exceptions the most attention in this respect.
are Ag’, F-, S042-, and Se042-. A possible reason for A neutron diffraction study of DF at 20 “C found in
the disagreement in the case of Ag+ is that the appli- the chain-like aggregates (two nearest neighbors) the
cable coordination number is not 6, but the RPion = hydrogen-bonded F-F internuclear distance to be 0.256
0.102 nm for a square-planar coordination20is the ap- nm.224Half of this can be taken to be the mean radius
propriate one also for the solution. A similar expana- of anhydrous hydrogen fluoride. Liquid ammonia was
tion, i.e., a much lower coordination number than 6 for studied by the molecular dynamics method at 196 K,
F- in the solution to account for the low value Rion = near its triple point,226by neutron diffraction at 208 and
0.124 nm is unlikely, however. In the cases of sulfate 295 K (22 0C),226and by X-ray diffraction at 277 K (4
and selenate the possible reasons for the discrepancies 0C).227The very short hydrogen-bonded N-N inter-
could be wrong estimates of the thermochemical radii, nuclear distance of 0.227 nm was found at the lowest
since the lattice energies of salts of these anions are not temperature, but only much higher distances, 0.34 and
known accurately enough. It should be noted that in 0.37 nm, were compatible with the neutron and X-ray
the case of Li+ the radius is between the average value diffraction data at the higher temperatures. Of the 12
in the solution, 0.206 nm, and the crystal ionic radius neighbors a central ammonia molecule has, about 6 are
for coordination number 6. If a coordination number at the nearer distance and 6 at the farther one. The
of 4 were selected, then the agreement would be be- mean radius of an ammonia molecule in the room tem-
tween RPion= 0.059 nm and dion-water = 0.198 nm, which perature liquid is thus 20.17 nm.
agrees well with data shown in Table I and Figure 1(for The structure of liquid formamide was studied by
low water-to-salt ratio). Therefore the question of what means of X-ray d i f f r a ~ t i o n ~ and
~ * - neutron
~~~ and
the “real” radius of Li+ in dilute aqueous solutions re- electron diffraction.229In the earliest X-ray diffraction
mains open. study228the hydrogen-bonded 0-N internuclear dis-
Another complication that ought to be considered is tance was 0.305 nm; in a later study by all three dif-
the fact that the water molecule is not spherical and fraction techniques229a somewhat shorter distance,
that in the first hydration sphere it directs a hydrogen 0.290 nm, was found. The nature of the aggregates,
atom toward anions, forming a hydrogen bond, whereas whether chainlike, rings, or both, could not be deter-
it directs its oxygen atom toward cations. How this mined unequivocally, but the presence of both kinds
affects the applicable Rwateris impossible to ascertain of aggregates seems now to be the more probable case.232
from simple considerations. The geometry of the ori- Since the formamide molecule is not spherical, it is
entation of the nearest water molecules to certain ions difficult to see whether a radius can be assigned on the
(Li’, Ni2+,and C1-, for instance) has been determined basis of this 0-N distance. However, the ability of the
by the FODNS t e c h n i q ~ e (see~ ~ tsection
~~ IV.A.2), so solvent to approach cations (with the oxygen atom to-
1494 Chemical Reviews, 1988, Vol. 88,No. 8 Marcus

ward the ion) and anions (with the N-bonded hydrogen ammonia in the temperature range 213-283 K and
atom toward the ion) to certain distances is another NH3/Rb ratios of 8-160 indicated the presence of Rb+
question, and probably one-half of the O-N distance ions with 6 ammonia molecules in the first solvation
is not the answer. In liquid N-methylformamide X-ray shell, at an Rb-N internuclear distance of 0.309nm. A
diffraction indicated the presence of a flexible linear similar study244 of ytterbium in ammonia solutions with
chain structure, which was confirmed by ab initio mo- 15.6,151, or 294 NH, per Yb found a first solvation shell
lecular orbital calculations.z33 In N,N-dimethylform- with n slightly >6 and an Yb-N internuclear distance
amide a similar study indicated the absence of any of 0.262 f 0.002 nm. The ytterbium was in the Yb2+
significant structure.z34 state. A Monte Carlo simualtion of a solution
Liquid methanol has been subjected to several stud- of Na+ in 18.45 mol % ammonia in water solution (206
ies, both by Monte Carlo computer simulationmYm and solvent molecules per Na+) indicated the presence of
by X-rayz37i238 and neutronz39diffraction. Chainlike 2.4 water and 4.0 ammonia molecules in the first sol-
aggregates are the main species, with hydrogen-bonded vation shell. The Na-N internuclear distance was taken
0-0 internuclear distances of 0.281 nm235and an av- as 0.235 nm from an ab initio calculation.
erage of 1.85 neighbors per molecule (i.e., the chains are An early X-ray diffraction studyzz8of KI in form-
sufficiently short for terminal molecules, with only one amide at 5.3-31.1 mol of solvent per mol of salt indi-
neighbor, to contribute significantly to the average). cated the presence of 2.1-4.6 solvent molecules in the
Increased pressure increases the number of neighbors first solvation shell of the K+ ion in this concentration
appreciably but decreases the distances only slightly.236 range, at a K-O internuclear distance of about 0.30 nm.
Somewhat shorter distances were obtained by the X-ray In spite of the high relative permittivity of formamide,
diffraction method: 0.278 nm with only an average of extensive ion pairing was postulated to occur in the
1.5 neighborsz37or, in a study with higher accuracy at solutions. Another X-ray diffraction study246 dealt with
20 0C,2380.2798 f 0.000~nm with an average of 1.77 f LiCl in formamide with 4.0 mol of solvent per mol of
0.07 neighbors. The neutron diffraction as- salt. It was found that the Li+ cation has preferentially
signed 0.285 nm to this distance in the attempted fit the 0 atom of formamide as its nearest neighbor, at a
of the data. The nonsphericity of the molecule prevents Li-0 distance of 0.224 f 0.002 nm and n = 5.4 f 0.3
the interpretation of the 0-0 distance in terms of a and that for the C1 anion it is the hydrogen-bonded N
molecular radius, but one-half of this distance, 0.140 atom that solvates it, at a Cl-N distance of 0.327 f
nm, appears to be the distance to which a solvent 0.002 nm and n = 4.5 f 0.1.
molecule can approach an ion (without taking into ac- An electron diffraction of ZnBrz in the
count the electrostriction; see section V.C). In two presence of about equimolar LiBr in N,N-dimethyl-
studies of ethanol, somewhat larger hydrogen-bonded formamide (DMF) at 6.27 mol of solvent per mol of
0-0 distances were found 0.285 nm with 1.9 neighbors ZnBrz indicated the presence of ZnBr3- as the major
in the chain from Monte Carlo calculationsw and 0.2808 species but that the solvent was still present in the first
f 0.0008 nm with l.&f 0.08 neighbors from X-ray solvation shell, with a Zn-0 internuclear distance of
diffraction at 20 0C.238 In tert-butyl alcohol (2- 0.222 nm. An X-ray diffraction studyzMof Cd(I1) in
methyl-2-propanol) a somewhat shorter 0-0 distance, DMF showed n = 6 and Cd-0 internuclear distances
0.274 nm, was found by X-ray diffraction at 26 "C, of 0.2296 nm. A recent X-ray diffraction studyz4'
slightly above the melting point.240 The same comment showed that Cu(I1) cations in Nfl-dimethylformamide
concerning the radius and the approach to ions made are located in the center of a distorted octahedron, with
for methanol applies also for the other alcohols. In an equatorial Cu-0 distances of 0.203 nm and axial ones
X-ray and neutron diffraction study of liquid formic of 0.243 nm (cf. the situation in aqueous solutions
acid the average mutual configuration of the C=O (Table VII)). When acetonitrile is added (to a mole
bonds in two neighboring molecules was determined, ratio of 2 per 1 DMF), the axial DMF solvent molecules
but no data were given from which an average radius are removed, but no acetonitrile ones replace them at
for this solvent could be obtained.z41 distances that can be considered as inside the solvation
An X-ray diffraction study of acetonitrile at 20 "C shell. The structure of the Cu(I1) solvation shell in pure
gave the center-to-center distance between a central acetonitrile could not be studied by X-ray diffraction,
molecule and each of its four nearest neighbors as 0.38 due to the low solubility of Cu(C104)zin this ~olvent.~'
nm1.242 Since, again, acetonitrile is not a spherical but However, an EXAFS studyzmof CuBrz in acetonitrile,
an elongated molecule, it is not immediately obvious without and with added LiBr, showed definite solvation
how this distance relates to the distance of approach of Cu(I1) by acetonitrile, with a Cu-N internuclear
of the solvent to an ion. distance of 0.203 nm, but with Br- anions also present
A brief discussion of structural information on no- in the first solvation shell.
naqueous solutions of ions, from which ion-solvent in- In dimethyl sulfoxide, an X-ray diffraction
ternuclear distances and hence ionic radii can be esti- showed that both Cd(I1) and Hg(I1) formed hexa-
mated, follows. No effort was made to scan the liter- solvates in the first solvation shell, with cation-0 in-
ature on this subject exhaustively. ternuclear distances of 0.2292 and 0.2393 nm, respec-
A neutron diffraction studyzz6of a concentrated tively. An X-ray diffraction study26l of ZnBrz in acetone
7Li-ND3 solution (having 4 ND3 molecules per Li atom) (4.07 mol of acetone per mol of salt) indicated bromide
at 216 K indicated the presence of a Li-N internuclear anions to be present in the first solvation shell and a
distance of 0.2 nm and a Li-D distance of 0.25 nm. If Br-0 internuclear distance of 0.32 nm, but no Zn-0
the electron is delocalized in this solution, the lithium distance was reported.
is ionized to Li+, so that the former distance is relevant Salt solutions in methanol received considerably more
to this review. An EXAFS of rubidium in attention than solutions in other solvents. In early
Ionic Radii in Aqueous Solutions Chemical Reviews, 1988, Vol. 88, No. 8 1495

studies of Wertz and K r ~ h ~cobalt(I1)


~ * ~ chloride
~ ~ r ~ ~of ~ions in solution for all the applications for which
and bromide and iron(II1) chloride solutions in meth- examples were shown in section 11.
anol were studied by X-ray diffraction. The C1-0 and Some questions, however, still remain open. One that
Br-O internuclear distances were found to be the same has already been touched upon in section V.C is
as in aqueous solutions, 0.31 nm27*252 and 0.325 nm,125 whether the same value of RWahapplies for cations and
respectively. In the latter case, also a Co-0 distance anions, in view of the different orientation of the water
of about 0.20 nm could be estimated for a species that molecule toward them. Another question is whether the
is tetrahedral, with two bromide and two methanol internuclear distances dionPwater are temperature and
ligands around the cobalt atom. LindlE2interpreted the pressure dependent, and if so, how much. Many of the
data for FeC13 as indicating the presence of FeC12(sol- diffraction measurements and the computer simulations
vent)4+species (beside FeC14-) with Fe-0 distances of were made at a specified temperature, usually 20 or 25
0.207 nm in both water and methanol. A Monte Carlo “C, but in many other cases only “room temperature”
computer simulation of sodium methoxide in metha- or the like was specified, if at all.
n01,253with 127 solvent molecules per ion, showed that A few determinations were made at several specified
for Na+ n = 6 and the Na-0 internuclear distance is temperatures. For instance, aqueous cadmium sulfate
0.232 f 0.003 nm and for CH30- n = 5 with the 0-0 was studied158by X-ray diffraction at 9 and 62 OC, but
distance 0.267 nm. An X-ray diffraction study2” of diOmaterwas determined to only &0.001 nm. The small
CuCl,, at 7.1-23.1 mol of methanol per mol of salt, differences noted between the results at these tem-
indicated the presence of the species Cu(C1, )2- peratures are probably not significant and could arise
(CH30H,)2(CH30H,)2, with Cu-0 internuclear Ais- from the different number of sulfate ions in the first
tances of 0.194 nm for the equatorial and 0.2465 nm for coordination shell of the cadmium ion at the two tem-
the axial solvent molecules. An X-ray s t u d p of MgC12 peratures, and not from inherent temperature sensi-
in methanol (18.7 mol of solvent per mol of salt) showed tivity of the free ion-solvent distance. An EXAFS
the regular octahedral arrangement of methanol around of some aqueous divalent metal halides at
the Mg2+cation with a Mg-0 internuclear distance of temperatures between 20 and 75 “C indicated that no
0.206 nm, and for C1- n = 6 also, but not necessarily change in n takes place in this range, and presumably
highly symmetrical, with a C1-0 distance of 0.315 nm. also none in dion-water. A Monte Carlo computer simu-
The ion-C distances were also given, so that the geom- lation explored the difference in the radial
etry of the solvent around the ions can also be inferred. distribution functions of water around Li+, Na+, K+, P,
A recent EXAFS studym of CoBr2in methanol showed and C1- between 25 and 75 “C at a ratio of 64 water
that C O B ~ ( C H ~ O His) ~the
+ predominant species in molecules per ion. The height of the first peak in the
concentrated solutions, but at 0.1 mol dm-3 the hexa- gion-water(r) curve for Li’, Na+, and P decreases with
solvate predominates, with Co-0 distances of 0.208 nm. increasing temperatures (as for pure water), but its
In ethanol at 0.2 mol dm-3 and higher concentrations position is independent of the temperature. On the
the predominant species is C O B ~ ~ ( C ~ H ~with
O HCo-0
)~, contrary, the peak heights for K+ and C1- increase
distances decreasing from 0.206 to 0.203 nm as the
slightly with the temperature, as do their r values. In
concentration increases to 3.83 mol dms. An earlier another Monte Carlo study2S8the thermal ellipsoids of
X-ray diffraction studyn of a CoBr2solution in ethanol the water molecules surrounding octahedrally a Zn2+
of a similar high concentration (3.81 mol kg-l) gave only cation were determined at 25 “C with 215 water mole-
the C1-0 distance, 0.31 nm, the same as in methanol
cules per zinc ion. In yet another computer simulation
or water.
this time by the molecular dynamics method,
the effect of pressure, 10 kbar, in dion-water was deter-
V II . Concludlng Remarks +
mined for the system involving 1Na+ + 1C1- 25 H20
The question was raised in the discussion of the at 25 “C. No change in the Na-0 distances was noted,
concept of ionic radii (section 1II.A) whether the in- but the C1-0 distance decreased slightly at the higher
ternuclear distances dionwwater between ions (or the cen- pressure. The value of n for Na+, however, increased
tral atoms of polyatomic ions) and water molecules (Le., from 5.8 to 6.3 as the pressure increased from ambient
the oxygen atom of the water molecule) in their first to 10 kbar. No generalizations can be drawn from these
solvation shell can be determined accurately. The an- rather sporadic, nonsystematic, studies.
swer to this question was yes: section IV presented the One final remark is in place here: the diffraction
diffraction and computer simulation methods, and measurements (by X-rays, neutrons, or electrons or by
section V (Tables I-XII) gave the results that have been the EXAFS method) and the computer simulations (by
obtained. These are summarized in Table XIII. The molecular dynamics and Monte Carlo methods) are
other question that was raised pertained to the alloca- applied to electrolyte solutions in only a relatively small
tion of these distances into the part that “belongs” to number of laboratories. The same names reappear in
the water molecule, Rwater,and the part that “belongs” the publications: see the relevant references 45-259.
to the ion, Rion,i.e., the ionic radius. The answer to that The most active groups are located in Italy (those of
was a qualified yes: if the average, electrostricted value Caminiti, Magini, Clementi, and their respective co-
of 0.1393 nm was used for RmW, then values of Rim were workers), Germany (that of Heinzinger and co-workers),
obtained that agreed well with the Pauling-type crystal England (that of Enderby, Neilson, and co-workers),
ionic radii for coordination number 6, RPion. Some Hungary (that of Palinkas and co-workers), Sweden
specific exceptions were explained ad hoc by the ne- (that of Johansson and co-workers), Japan (that of
cessity to use RPionvalues for other coordination num- Ohtaki and co-workers),while several additional groups
bers. The general conclusion from this is that Paul- made important contributions, too. Remarkable, how-
ing-type crystal ionic radii may serve well for the radii ever, is the constant cooperation between many of these
1496 Chemical Reviews, 1988, Vol. 88, No. 8 Marcus

groups, co-workers from one being temporary guests (45) Yamaguchi, T.; Ohtaki, H.; Spohr, E.; Palinkas, G.; Hein-
with another, so that many of the papers bear addresses zinger, K.; Probst, M. M. 2.Naturforsch. 1986, A41, 1175.
(46) Hewish, N. A.; Neilson, G. W.; Enderby, J. E. Nature (Lon-
of several laboratories. A community of researchers don) 1982, 297, 138.
with interest in this field appears to have been built up (47) Bounds, D. G. Mol. Phys. 1985,54, 1335.
(48) Yamaguchi, T.; Lindqvist, 0.;Claeson, T.; Boyce, J. B. Chem.
with worldwide connections, a very commendable de- Phys. Lett. 1982, 93, 528.
velopment. (49) Narten, A. H.; Vaslow, F.; Levy, H. A. J. Chem. Phys. 1973,
58, 5017.
Acknowledgments. This work was supported in part (50) Licheri, G.; Piccaluga, G.; Pinna, G. J . Appl. Crystallogr.
1973, 6, 392.
by Grant No. 84-00292 from the U.S.-Israel Binational (51) Heinzinger, K. Pure Appl. Chem. 1985,57, 1031.
Science Foundation. (52) Impey, R. W.; Madden, P. A,; McDonald, I. R. J. Phys.
Chem. 1983,87, 5071.
(53) Marchese, F. T.; Beveridge, D. L. J. Am. Chem. SOC.1984,
V II I . References 106. 3713.
I

(54) Bopp, P.; Okada, I.; Ohtaki, H.; Heinzinger, K. Z. Natur-


(1) Stokes, R. H. J. Am. Chem. SOC.1964,86,979. forsch. 1985, A40, 116.
(2) Rashin, A. A.; Honig, B. J . Phys. Chem. 1985,89, 5588.
1981, A%, 1067.
-
(55) Szasz. Gv. I.: Heinzinger. K.: Riede. W. 0. 2. Naturforsch.
(3) Marcus, Y. J. Solution Chem. 1983, 12, 271.
(4) Marcus, Y. Zon Soluation; Wiley: Chichester (UK), 1986; p (56) Heinzinger, K. Stud. Phys. Theor. Chem. 1983,27, 61.
113. (57) Radnai, T.; Heinzinger, K.; Szasz, Gy. I. Z. Naturforsch. 1981,
(5) Marcus, Y. Pure Appl. Chem. 1987, 59, 1093. A36, 1076.
(6) Abraham, M. H.; Marcus, Y. J . Chem. SOC.,Faraday Trans. (58) Paschina, G.; Piccaluga, G.; Pinna, G.; Magini, M. Chem.
1 1986,82, 3255. Phys. Lett. 1983, 98, 157.
(7) Millero, F. J. Chem. Reu. 1971, 71, 147. (59) Musinu, A.; Paschina, G.; Piccaluga, G.; Magini, M. J. Chem.
(8) Marcus, Y.; Loewenschuss, A. Annu. Rep. C , 1984, R. SOC. Phys. 1984,80, 2772.
Chem. (London) 1985, 81. (60) Mezei, M.; Beveridge, D. L. J. Chem. Phys. 1981, 74, 6902.
(9) Burgess, J. Metal Ions in Solution; Ellis horwood: Chi. (61) Ergin, Yu. V.; Koop, 0. Ya.; Khrapko, A. M. Zh. Fiz. Khim.
Chester (UK), 1978; Chapter 11. 1979,53, 2109.
(10) Eigen, M. 2. Ph s. Chem. (FrankfurtlMain) 1964, 1, 176. (62) Okada, I.; Kitsuno, Y.; Lee, H.-G.; Ohtaki, H. Stud. Phys.
.
(11) Nightin ale, E. $ J. Phys. Chem. 1959,63, 1381. Theor. Chem. 1983,27, 81.
(12) Hertz, 8. G.; Spaltoff, W. Z. Elektrochem. 1959, 63, 1096.
(13) Hertz, H. G. Ber. Bunsenges. Phys. Chem. 1963, 67, 311.
(63) Okada, I.; Kitsuno, Y.; Lee, H.-G.; Ohtaki, H. Abstr. VZZnt.
Symp. Solute-Solute-Soluent Interact., Minoo 1982, p 1A-
(14) Blandamer, M. J.; Griffiths, T. R.; Shields, L.; Symons, M. M07.
C. R. Trans. Faraday SOC. 1964,60, 1524. (64) Newsome, J. R.; Neilson, G. W.; Enderby, J. E. J. Phys. C:
(15) Plowman, K. R.; Lagowski, J. J. J. Phys. Chem. 1974,78,143. Solid State Phys. 1980, 13, L923.
(16) Maxey, B. W.; Popov, A. I. J. Am. Chem. SOC. 1967,89,2230. (65) Palinkas, G.; Radnai, T.; Hajdu, F. 2. Naturforsch. 1980, A35,
(17) Coetzee, J. F.; Sharpe, W. R. J. Solution Chem. 1972, I, 77. 107.
(18) Goldschmidt, V. M. Skr. Norske Vid. Akad. Oslo, Math. Nat. (66) Ohtomo, N.; Arakawa, K. Bull. Chem. SOC. Jpn. 1979, 52,
K1. 1926, No. 2; Ber. 1927,60, 1263. 2755.
(19) Pauling, L. J. Am. Chem. SOC. 1927,49, 765; Nature of the (67) Licheri, G.; Piccaluga, G.; Pinna, G. Chem. Phys. Lett. 1975,
Chemical Bond, 3rd ed.; Cornel1 University Press: Ithaca, 35,119.
NY. 1960 Chanter 13. (68) Heinzinger, K.; Vogel, P. C. 2.Naturforsch. 1976, A31,463.
(20) Shannon,'R. D.; Prewitt, C. T. Acta Crystallogr., Sect. B (69) Szasz, Gy. I.; Heinzinger, K.; Palinkas, G. Chem. Phys. Lett.
1969,25, 925; 1970,26, 1046. 1981, 78, 194.
(21) Ahrens, L. H. Geochim. Cosmochim. Acta 1952,2, 155. (70) Kistenmacher, H.; Popkie, H.; Clementi, E. J. Chem. Phys.
(22) Gourary, B. S.; Adrian, F. J. Solid State Phys. 1960,10,128. 1974, 61, 799.
(23) KaDustinskii. A. F.: Yatsimirskii. K. B. Zh. Obshch. Khim. (71) Clementi, E.; Barsotti, R. Chem. Phys. Lett. 1978, 59, 21.
19d9,19,2191. Kapustinskii, A. F. Q.Reu., Chem. SOC.1956, (72) Ohtomo, N.; Arakawa, K. Bull. Chem. SOC. Jpn. 1980, 53,
10. 283.
- - 7
1789.
(24) Yataimirskii, K. B. Zzu. Akad. Nauk SSSR, Ot. Khim. Nauk (73) Caminiti, R.; Licheri, G.; Piccaluga, G.; Pinna, G. Rend. Sem.
1947,453; 1948, 398. Fac. Sci. Cagliari 1977, XLVZ, Suppl. 19.
(25) Jenkins. H. B. D.: Thakur, K. P. J . Chem. Educ. 1979,56, (74) Vogel, P. C.; Heinzinger, K. 2. Naturforsch. 1976, A31,476.
576. (75) Palinkas, G.; Riede, W. 0.; Heinzinger, K. 2. Naturforsch.
(26) Marcus, Y.; Loewenschuss, A. Annu. Rep. C , 1984, R. SOC. 1977, A32, 1137.
Chem. (London) 1985,81. (76) Bopp, P.; Dietz, W.; Heinzinger, K. 2. Naturforsch. 1979,
(27) Levy, H. A.; Danford, M. D. In Molten Salt Chemistry; A34, 1424.
Blander, M., Ed.; Wiley: New York, 1964; Chapter 2. (77) Maeda, M.; Ohtaki, H. Bull. Chem. SOC. Jpn. 1975,48,3755.
(28) Marcus, Y. Introduction to Liquid State Chemistry;Wiley: (78) Caminiti, R.; Licheri, G.; Paschina, G.; Piccaluga, G.; Pinna,
Chichester (UK), 1977; p 121. G. J. Chem. Phys. 1980, 72,4522.
(29) (a) Narten. A. H.: Lew. H. A. J. Chem. Phvs. 1971.55. 2263. (79) Limtakrul, J. P.; Rode, B. M. Monatsh. Chem. 1985, 116,
' id) Narten, A. H: Zbid.' 1972, 56, 5681. " 1377.
(30) Kalman, E.; Lengyel, S.; Haklik, L.; Eke, E. J . Appl. Crys- (80) Brady, G. W. J . Chem. Phys. 1958,28,464.
tallogr. 1974, 7, 442. (81) Neilson, G. W.; Skipper, N. Chem. Phys. Lett. 1985,114, 35.
(a) Lie, G. C.; Clementi, E. J. Chem. Phys. 1975,62,2195. (b) (82) Terekhova, D. S.; Ryss, A. I.; Radchenko, I. V. J. Struct.
Lie, G. C.; Clementi, E.; Yoshimine, M. Zbid. 1976,64, 2314. Chem. 1969,10, 807.
(a) Rahman,. 4.; Stillinger, F. H. J. Che?. Phys. 1971, 55, (83) Eck, C. L. van P. van; Mendel, H.; Boog, W. Discuss.Fara-
3336. (b) Stillinger, F. H.; Rahman, A. Zbid. 1974,60, 1545. day SOC.1957,24, 200,235.
Lonsdale, K. Proc. R. SOC. (London),A 1958, A247, 424. (84) Szasz, Cy. I.; Heinzinger, K. Z. Naturforsch. 1983, A38,214.
Kruh, R. F. Chem. Reu. 1962,62, 319. (85) Szasz, Gy. I.; Heinzinger, K. Z. Naturforsch. 1979, A34,840.
Soper, A. K.; Neilson, G. W.; Enderby, J. E.; Howe, R. A. J . (86) Lawrence, R. M.; Kruh, R. F. J . Chem. Phys. 1967,47,4758.
Phys. (C),Solid State Phys. 1977,10, 1793. (87) Bertagnolli, H.; Weidner, J.-U.; Zimmermann, H. W. Ber.
(36) Enderby, J. E.; Neilson, G. W. Rep. Prog. Phys. 1981,44,593. Bunsenges. Phys. Chem. 1974, 78, 2.
(37) Annis. B. K.: Hahn. R. L.; Narten, A. H. J. Chem. Phys. 1985, (88) Maeda, M.; Maegawa, Y.; Yamaguchi, T.; Ohtaki, H. Bull.
82, 2086. Chem. SOC. Jpn. 1979,52, 2545.
(38) Licheri, G.; Pinna, G. EXAFS and Near Edge Structure; (89) Sandstrom, M.; Neilson, G. W.; Johansson, G.; Yamaguchi,
Bianconi, A,, Incoccia, L., Stipcich, S., Eds.; Springer Ser. T. J . Phys. C: Solid State Phys. 1985, 18, L1115.
Chem. Phys. 1983,27, 240. (90) Yamaguchi, T.; Lindqvist, 0.;Boyce, J. B.; Claeson, T. Acta
(39) Kalman, E.; Serke, I.; Palinkas, G.; Johansson, G.; Kabisch, Chem. Scand. 1984, A38,423.
G.; Maeda, M.; Ohtaki, H. 2. Naturforsch. 1983, A38, 225. (91) Yamaguchi, T.; Johansson, G.; Holmberg, B.; Maeda, M.;
(40) Heinzinger, K. Physica B & C 1985, 131, 196. Ohtaki, H. Acta Chem. Scand. 1984, A38, 437.
(41) Kong, C. L. J . Chem. Phys. 1973,59, 2464. (92) Ohtomo, N.; Arakawa, K.; Takeuchi, M.; Yamaguchi, T.;
(42) Alagona, G.; Ghio, C.; Kollman, P. J . Am. Chem. SOC. 1986, Ohtaki, H. Bull. Chem. SOC.Jpn. 1981,54, 1314.
108, 185. (93) Caminiti, R. Chem. Phys. Lett. 1983, 96, 390.
(43) Briand, C. L.; Burton, J. J. J. Chem. Phys. 1976,64,2888. (a) (94) Caminiti, R.; Cucca, P. Chem. Phys. Lett. 1984, 108, 51.
Ozutsumi, K.; Ohtaki, H.; Kusumegi, A. Bull. Chem. SOC. (95) Lee, H.-G.; Matsumoto, Y.; Yamaguchi, T.; Ohtaki, H. Bull.
Jpn. 1984,57, 2612. Chem. SOC. Jpn. 1983,56,443.
(44) Malenkov, G. G.; Dyakonova, L. P.; Brizhik, L. S. VINITI (96) Caminiti, R.; Atzei, D.; Cucca, P.; Squintu, F.; Bongiovanni,
1980, 346-80. G. Z. Naturforsch. 1985, A40, 1319.
Ionic Radii in Aqueous Solutions Chemical Reviews, 1988, Vol. 88, No. 8 1497

(97) Hewish, N. A.; Neilson, J. E. Chem. Phys. Lett. 1981,84,425. (149) Ohtaki, H.; Maeda, M. Bull. Chem. SOC. Jpn. 1974,47,2197.
(98) Bol, W.; Gerrits, G. J. A,; Eck, C. L. van P. van J. Appl. (150) Johansson, G.; Ohtaki, H. Acta Chem. Scand. 1973,27,643.
Crystallogr. 1970, 3, 486. (151) Eisenberger, P.; Kincaid, B. M. Chem. Phys. Lett. 1975,36,
(99) Caminiti, R. J. Mol. Liq. 1984,28, 191. 134.
(100) Dietz, W.; Riede, W. 0.;Heinzinger,K. 2.Naturforsch. 1982, (152) Musinu, A.; Paschina, G.; Piccaluga, G.; Magini, M. J. Appl.
A37,1038. Crystallogr. 1982, 15, 621.
(101) Palinkas, G.; Radnai, T.; Dietz, W.; Szasz, Gy. I.; Heinzinger, (153) Radnai, T.; Palinkas, G.; Caminiti, R. Z. Naturforsch. 1982,
K. 2.Naturforsch. 1982, A37, 1049. A37, 1247.
(102) Caminiti, R. Chem. Phys. Lett. 1982,88, 103. (154) Dagnall, S. P.; Hague, D. N.; Towl, A. D. C. J. Chem. Soc.,
(103) Caminiti, R.; Cucca, G.; Monduzzi, M.; Saba, G.; Crisponi, G. Faraday Trans. 2 1982, 78, 2161.
J. Chem. Phys. 1984,81, 543. (155) Wertz, D. L.; Bell, J. R. J. Znorg. Nucl. Chem. 1973,35,861.
(104) Caminiti, R.; Licheri, G.; Piccaluga, G.; Pinna, G. J. Appl. (156) Musinu, A.; Paschina, G.; Piccaluga, G.; Magini, M. J. Appl.
Crystallogr. 1979, 12, 34. Crystallogr. 1982, 15, 621.
(105) Caminiti. R.: Licheri. G.: Piccaluga. G.: Pinna. G. Chem.
I
(157) Caminiti,, R.:. Johansson. G. Acta Chem. Scand. 1981., A35.,
Phvs. Lett. 1979. 61. 45. ' 373.
(106) Aliright, J. N. Jl Chem. Phys. 1972, 56, 3783. (158) Caminiti, R. Z. Naturforsch. 1981, A36, 1062.
(107) Ryss, A. I.; Radchenko, I. V. J. Struct. Chem. 1965,6,422. (159) Caminiti, R.; Cucca, P.; Radnai, T. J. Phys. Chem. 1984,88,
(108) Dorosh, A. K.; Skrychevskii, A. F. J. Struct. Chem. 1964,5, 2382.
842. (160) Caminiti, R.; Licheri, G.; Piccaluga, G.; Pinna, G. 2.Natur-
(109) Caminiti, R.; Licheri, G.; Piccaluga, G.; Pinna, G. Chem. forsch. 1980, A35, 1361.
Phys. Lett. 1977, 47, 275. (161) Ohtaki, H.; Maeda, M.; Ito, S. Bull. Chem. SOC. Jpn. 1974,
(110) Probst, M. M.; Radnai, T.; Heinzinger, K.; Bopp, P.; Rode, 47, 2217.
B. M. J. Phys. Chem. 1985,89,753. (162) Sadoc, A.; Lagarde, P.; Vlaic, G. J. Phys. C 1985, 18, 23.
(111) Licheri, G.; Piccaluga, G.; Pinna, G. J. Chem. Phys. 1976,64, (163) van Eck, C. L. van P.; Wolters, H. B. M.; Jaspers, W. J. M.
2437. Recl. Trav. Chim. Pays-Bas 1956, 75,802.
(112) Cu"ings, S.; Enderby, J. E.; Howe, R. A. J.Phys. C Solid (164) Johansson, G. Acta Chem. Scand. 1971,25, 2787.
State Phys. 1980, 13, 1. (165) Sandstrom, M.; Persson, I.; Ahrland, S. Acta Chem. Scand.
(113) Licheri, G.; Piccaluga, G.; Pinna, 0. J. Chem. Phys. 1975,63, 1978. A32.607.
4412. (166) Caminiti, R.; Licheri, G.; Piccaluga, G.; Pinna, G.; Radnai, T.
(114) Shapovalov, I. M.; Radchenko, I. V. J. Struct. Chem. 1971, J. Chem. Phys. 1979, 71, 2473.
12, 705. (167) Caminiti, R.; Radnai, T. 2.Naturforsch. 1980, A35, 1368.
(115) Caminiti, R.; Musinu, A.; Paschina, G.; Pinna, G. J. Appl. (168) Bol. W.: Welzen. T. Chem. Phvs. Lett. 1977.49. 189.
Crystallogr. 1982,15,482. (16$ Joh'ansson, G.; Wakita, H. Zn&g. Chem. 1985,24, 3047.
(116) Licheri, G.; Pinna, G. EXAFS and Near Edge Structure; (170) Johansson, G.; Niinitcro, L.; Wakita, H. Acta Chem. Scand.
Bianconi, A., Incoccia, L., Stipcich, S., Eds.; Springer Ser. 1985, A39, 359.
Chem. Phys. 1983,27, 240. (171) Caminiti, R.; Licheri, G.; Piccaluga, G.; Pinna, G. J. Chem.
(117) Caminiti, R.; Marongin, G.; Paschina, G. 2. Naturforsch. Phys. 1976,65,3134.
1982, A37, 581. (172) Caminiti, R. Chem. Phys. Lett. 1982, 86, 214.
(118) Caminiti, R.; Cucca, G.; Pintori, T. Chem. Phys. 1984,88,155. (173) Caminiti, R.; Licheri, G.; Piccaluga, G.; Pinna, G. J. Chem.
(119) Licheri, G.; Paschina, G.; Piccaluga, G.; Pinna, G. J. Chem. Phys. 1978, 69, 1.
Phys. 1984,81,6059. (174) Caminiti, R.; Licheri, G.; Piccaluga, G.; Pinna, G. Chem.
(120) Ohtaki. H.: Yamarmchi. T.: Maeda. M. Bull. Chem. SOC. hn. Phvs. 1977. 19. 371.
1976,49, 701. " ' ' (175) Cri'stini,-A.ILich&<G.; Piccaluga, G.; Pinna, G. Chem. Phys.
(121) Tajiri, Y.; Ichihashi, M.; Mibuchi, T.; Wakita, H. Bull. Chem. Lett. 1974, 24, 289.
SOC. Jpn. 1986,59,1155. (176) Magini, M. J. Chem. Phys. 1980, 73, 2499.
(122) Ichihashi, M.; Wakita, H.; Masuda, I. J. Solution Chem. 1984, (177) Lee, W. K.; Prohofsky, E. W. J. Chem. Phys. 1981,75,3040.
13, 505. (178) Magini, M.; Caminiti, R. J. Znorg. Nucl. Chem. 1977,39,91.
(123) Magini, M.; Giubileo, G. Gazz. Chim. Ztal. 1981, 111, 449. (179) Magini, M.; Radnai, T. J. Chem. Phys. 1979, 71,4255.
(124) Magini, M. J. Chem. Phys. 1981, 74, 2523. (180) Magini, M. J. Chem. Phys. 1979, 70, 317.
(125) Wertz, D. L.; Kruh, R. F. Znorg. Chem. 1970, 9, 595. (181) Magini, M. J. Znorg. Nucl. Chem. 1978,40,43.
(126) Shapovalov, I. M.; Radchenko, I. V.; Lesovitskaya, M. K. J. (182) Lind, M. D. J. Chem. Phys. 1967,46, 2010.
Struct. Chem. 1972, 13,121. (183) Caminiti, R.; Atzei, D.; Cucca, P.; Anedda, A.; Bongiovanni,
(127) Caminiti, R.; Cucca, P. Chem. Phys. Lett. 1982,89, 110. G. J. Phys. Chem. 1986,90,238.
(128) Neilson, G. W.; Enderby, J. E. Proc. R. SOC.London 1983, (184) Caminiti, R.; Paschina, G. Chem. Phys. Lett. 1981, 82, 487.
A390, 353. (185) Maeda, M.; Ohtaki, H. Bull. Chem. SOC.Jpn. 1977,50,1893.
(129) Enderby, J. E. Pure Appl. Chem. 1985,57, 1025. (186) Tanaka, T.; Ogita, N.; Tamura, Y.; Okada, I.; Ohtaki, H.;
(130) Caminiti, R. J. Chem. Phys. 1982, 77, 5682. Palinkas, G.; Spohr, E.; Heinzinger, K. Z. Naturforsch. 1987,
(131) Licheri, G.; Pinna, G.; Navarra, G.; Vlaic, G. 2.Naturforsch. A42, 29.
1983, A38, 559. (187) Glaser, J.; Johansson, G. Acta Chem. Scand. 1982, A36,125.
(132) Caminiti, R.; Licheri, G.; Piccaluga, G.; Pinna, G. Faraday (188) Glaser, J. Acta Crystallogr. 1978, A34, S197.
Discuss. Chem. SOC.1978, 64, 62. (189) Magini, M.; Cabrini, A.; Scibona, G.; Johansson, G.; Sand-
(133) Neilson, G. W.; Enderby, J. E. J. Phys. C Solid State Phys. strom, M. Acta Chem. Scand. 1976, A30,437.
1978,11, L625. (190) Smith, L. S.; Wertz, D. L. J. Am. Chem. SOC. 1975,97,2365.
(134) Newsome, J. R.; Neilson, G. W.; Enderby, J. E.; Sandstrom, Smith, L. S.; McCain, D. C.; Wertz, D. L. J. Am. Chem. SOC.
M. Chem. Phys. Lett. 1981, 82, 399. 1976,98, 5125.
(135) Magini, M.; DeMoraes, M.; Licheri, G.; Piccaluga, G. J. (191) Habenschuss, A.; Spedding, F. H. J. Chem. Phys. 1979, 70,
Chem. Phys. 1985,83,5797. 3758.
(136) Wakita, H.; Ichihashi, M.; Mibuchi, T.; Masuda, I. Bull. (192) Smith, L. S.; Wertz, D. L. J. Znorg. Nucl. Chem. 1977,39,95.
Chem. SOC.Jpn. 1982,55,817. (193) Caminiti, R.; Cucca, P.; DAndrea, A. Z. Naturforsch. 1983,
(137) Caminiti, R. J. Chem. Phys. 1986, 84, 3336. A38,533.
(138) Sandstrom, D. R. J. Chem. Phys. 1979, 71, 2381. (194) Narten, A. H.; Hahn, R. L. Science (Washington, D.C.) 1982,
(139) Sandstrom, D. R.; Dodgen, H. W.; Lytle, F. W. J. Chem. 21 7, 1249.
Phys. 1977,67,473. (195) Narten, A. H.; Hahn, R. L. J. Phys. Chem. 1983,87, 3193.
(140) Caminiti, R.; Magini, M. Chem. Phys. Lett. 1979, 61, 40. (196) Steele, M. L.; Wertz, D. L. Znorg. Chem. 1977, 16, 1125.
(141) Soper, A. K.; Neilson, G. W.; Enderby, J. E.; Howe, R. A. J. (197) Ryss, A. 1.; Levositskaya, M. K.; Shapovalov, I. M. VZNZTZ
Phys. C: Solid State Phys. 1977,10, 1793. 1976,856-76; Chem. Abstr. 1978,89,95116.
(142) Neilson, G. W. J. Phys. C 1982, 15, L233. (198) Habenschuss, A.; Spedding, F. H. J. Chem. Phys. 1980, 73,
(143) Licheri, G.; Musinu, A.; Paschina, G.; Piccaluga, G.; Sedda, 442.
A. F. J. Chem. Phys. 1984,80, 5308. (199) Steele, M. L.; Wertz, D. L. J. Am. Chem. SOC.1976,98,4424.
(144) Neilson, G. W. Private communication in: Enderby, J. E.; (200) Habenschuss, A.; Spedding, F. H. J. Chem. Phys. 1979, 70,
Neilson, G. W. Rep. Prog. Phys. 1981,44,593. 2797.
(145) Lagarde, P.; Fontaine, A.; Raoux, D.; Sadoc, A.; Migliardo, P. (201) h i s , B. K.; Hahn, R. L.; Narten, A. H. J.Chem. Phys. 1985,
J. Chem. Phys. 1980, 72, 3061. 82, 2086.
(146) Fontaine, A.; La arde, P.; Raoux, D.; Fontana, M. P.; Mais- (202) Brady, G. W. J. Chem. Phys. 1960,33,1079. (a) Marcus, Y.
apo, G.; Migliario, P.; Wanderlingh, F. Phys. Chem. Lett. In Gmelin Handbook of Inorganic Chemistry, Rare Earths
1978, 41, 504. 1981, Suppl. Vol. 0 3 , 1.
(147) Ichihashi, M.; Wakita, H.; Mibuchi, T.; Masuda, I. Bull. (203) Narten, A. H. J. Phys. Chem. 1970, 74, 765.
Chem. SOC. Jpn. 1982,55, 3160. (204) Copestake, A. P.; Neilson, G. W.; Enderby, J. E. J. Phys. C
(148) Fishkis, M. Ya.; Zhmak, V. A. J.Struct. Chem. 1974, 15, 1. 1985, 18,4211.
1498 Chemical Reviews, 1988, Vol. 88, No. 8 Marcus

(205) Lee, S. C.; Kaplow, R. Science (Washington,D.C.) 1970,169, (234) Ohtaki, H.; Itoh, S.; Yamaguchi, T.; Ishiguro, S.; Rode, B. M.
477. Bull. Chem. SOC.Jpn. 1983,56,3406.
(206) Triolo, R.; Narten, A. H. J. Chem. Phys. 1975, 63, 3624. (235) Jorgensen, W. L. J. Am. Chem. SOC.1981, 103, 341, 345.
(207) Cummings, S.; Enderby, J. E.; Neilson, G. VI.; Newsome, J. (236) Jorgensen, W. L.; Ibrahim, M. J. Am. Chem. SOC.1982,104,
R.; Howe, R. A.; Howells, W. S.; Soper, A. K. Nature (Lon- 373.
don) 1980,287, 714. (237) Magini, M.; Paschina, G.; Piccaluga, G. J. Chem. Phys. 1982,
(208) Brady, G. W.; Krause, J. T. J. Chem. Phys. 1957, 27, 304. 77, 2051.
(209) Biggin, S.; Enderby, J. E.; Hahn, R. L.; Narten, A. H. J. Phys. (238) Narten, A. H.; Habenschuss, A. J. Chem. Phys. 1984, 80,
Chem. 1984,88,3634. 3387.
(210) Neilson. G. W.: Schiobere. D.:. Luck. W. A. P. Chem. Phvs.
-I
(239) Tanaka, Y.; Ohtomo, N.; Arakawa, K. Bull. Chem. SOC. Jpn.
Lett. 1985, 122, 475. 1985, 58, 270.
(211) Neilson, G. W.; Enderby, J. E. J. Phys. C Solid State Phys. (240) Narten, A. H.; Sandler, S. I. J. Chem. Phys. 1979, 71, 2069.
1982,15, 2347. (241) Bertagnolli, H.; Chieux, P.; Hertz, H. G. Ber. Bunsenges.
(212) Caminiti, R.; Licheri, G.; Piccaluga, G.; Pinna, G. J. Chem. Phys. Chem. 1984,88,977.
Phys. 1978,68, 1967. (242) Kratochwill, A.; Weidner, J. U.; Zimmermann, H. Ber. Bun-
(213) Caminiti, R.; Cucca, P.; Atzei, D. J. Phys. Chem. 1985, 89, senges. Phys. Chem. 1973, 77,408.
1457. (243) Acrivos, J. V.; Hathaway, K.; Robertson, A.; Thompson, A.;
(214) Caminiti, R.; Paschina, G.; Pinna, G.; Magini, M. Chem. Klein, M. P. J. Phys. Chem. 1980, 84, 1206.
Phys. Lett. 1979, 64, 391. (244) Lelieur, J. P.; Goulon, J.; Cortes, R.; Friant, P. J. Phys.
(215) Johansson, G.; Caminiti, R. 2.Naturforsch. 1986, A41,1325. Chem. 1984,88, 3730.
(216) Hajdu, F.; Lengyel, S.; Palinkas, G. J. Appl. Crystallogr. (245) Tanabe, Y.; Rode, B. M. Abstr. VZZZ Znt. Symp. Solute-So-
1976, 9, 134. lute-Solvent Interact., Regensburg, 1987; Barthel, J.,
(217) Kalman, E.; Palinkas, G.; Kovaa, P. Mol. Phys. 1977,34,505, Schmeer, G., Eds.; p P1.06.
525. (246) Ohtaki, H.; Wada, H. J. Solution Chem. 1985, 14, 209.
(218) Thiessen, W. E.; Narten, A. H. J. Chem. Phys. 1982,77,2656. (247) Kabisch, G.; Kalman, E.; Palinkas, G.; Radnai, T.; Gaizer, F.
(219) Gibson, I. P.; Dore, J. C. Mol. Phys. 1983, 48,1019. Chem. Phys. Lett. 1984,107,463.
(220) Impey, R. W.; Klein, M. L.; McDonald, I. R. J. Chem. Phys. (248) Sandstrom, M.; Persson, I.; Ahrland, S. Acta Chem. Scund.
1981,-74, 647. 1978, A32, 607.
(221) Desnoyers, J. E.; Verall, R. E.; Conway, B. E. J. Chem. Phys. (249) Ohtaki, H. Pure Appl. Chem. 1987,59, 1143.
1965, 43, 243. (250) Sano, M.; Maruo, T.; Yamatera, H. J. Chem. Phys. 1986,84,
(222) Jansco, G.; Bopp, P.; Heinzinger, K. Chem. Phys. 1984,85, 66
377. Wertz, D. L.; Lawrence, R. M.; Kruh, R. F. J. Chem. Phys.
(223) Gaballa, G. A.; Neilson, G. W. Mol. Phys. 1983,50, 97. 1965, 43, 2163.
(224) Deraman, M.; Dore, J. C.; Powles, J. G.; Holloway, J. H.; Wertz, D. L.; Kruh, R. F. J. Chem. Phys. 1969, 50, 4013.
Chieux, P. Mol. Phys. 1985, 55, 1351. Jorgensen, W. L.; Bigot, B.; Chandrasekhar, J. J. Am. Chem.
(225) McDonald, I. R.; Klein, M. L. J. Chem. Phys. 1976,64,4790. SOC.1982, 104, 4584.
(226) Chieux, P.; Berta nolli, H. J. Phys. Chem. 1984, 88, 3726. Ichihashi, M.; Wakita, H.; Masuda, I. Bull. Chem. SOC.Jpn.
(227) Narten, A. H. J. t h e m . Phys. 1977,66, 3117. 1983,56, 3761.
(228) DeSando, R, J.; Brown, G. H. J. Phys. Chem. 1968, 72,1088. Radnai, T.; Kalman. E.: Pollmer. K. Z. Naturforsch. 1984.
(229) Kalman, E.; Serke, I.; Palinkas, G.; %idler, M. D.; Wiesmann, A39, 464.
F. J.;Bertagnolli, H.; Chieux, P. 2.Naturforsch. 1983,A38, (256) Sano, M.; Maruo, T.; Masuda, Y.; Yamatera, H. J. Solution
.
.
n
Zdl. Chem. 1986.15.803.
- 7 - ---
I

(230) Ohtaki, H.; Funaki, A.; Rode, B. M.; Reibnegger, G. J. Bull. (257) Ergin, Yu. V.; Koop, 0. Ya.; Khrapko, A. M. Zh. Fiz. Khim.
Chem. SOC.Jpn. 1983,56, 2116. 1981,55, 173; Russ. J. Phys. Chem. (Engl. Transl.) 1981,55,
(231) Miyake, M.; Ka'i, 0.; Nakagawa, N.; Suzuki, T. J. Chem. sn
SOC.,Faraday d a m . 2 1985,81, 277. (258) Mkchese, F. T.; Beveridge, D. L. Znt. J. Quant. Chem. 1986,
(232) Ohtaki, H.; Itoh, S. Z. Naturforsch. 1985, A40, 1351. 29. 619.
(233) Ohtaki, H.; Itoh, S.; Rode, B. M. Bull. Chem. SOC.Jpn. 1986, (259) Jansco, G.; Heinzinger, K.; Radnai, T. Chem. Phys. Lett.
59, 271. 1984, 110, 196.

Potrebbero piacerti anche