Ion association

Last updated

In chemistry, ion association is a chemical reaction whereby ions of opposite electric charge come together in solution to form a distinct chemical entity. [1] [2] Ion associates are classified, according to the number of ions that associate with each other, as ion pairs, ion triplets, etc. Ion pairs are also classified according to the nature of the interaction as contact, solvent-shared or solvent-separated. The most important factor to determine the extent of ion association is the dielectric constant of the solvent. Ion associates have been characterized by means of vibrational spectroscopy, as introduced by Niels Bjerrum, and dielectric-loss spectroscopy. [3] [4]

Contents

Classification of ion pairs

Ion pairs are formed when a cation and anion, which are present in a solution of an ionizable substance, come together to form a discrete chemical species. There are three distinct types of ion pairs, depending on the extent of solvation of the two ions. For example, magnesium sulfate exists as both contact and solvent-shared ion-pairs in seawater. [5]

In the schematic representation above, the circles represent spheres. The sizes are arbitrary and not necessarily similar as illustrated. The cation is coloured red and the anion is coloured blue. The green area represents solvent molecules in a primary solvation shell; secondary solvation is ignored. When both ions have a complete primary solvation sphere, the ion pair may be termed fully solvated (separated ion pair, SIP). When there is about one solvent molecule between cation and anion, the ion pair may be termed solvent-shared. Lastly, when the ions are in contact with each other, the ion pair is termed a contact ion pair (CIP). Even in a contact ion pair, however, the ions retain most of their solvation shell. The nature of this solvation shell is generally not known with any certainty. In aqueous solution and in other donor solvents, metal cations are surrounded by between 4 and 9 solvent molecules in the primary solvation shell, [6]

An alternative name for a solvent-shared ion pair is an outer-sphere complex. This usage is common in coordination chemistry and denotes a complex between a solvated metal cation and an anion. Similarly, a contact ion pair may be termed an inner-sphere complex. The essential difference between the three types is the closeness with which the ions approach each other: fully solvated > solvent-shared > contact. With fully solvated and solvent-shared ion pairs the interaction is primarily electrostatic, but in a contact ion pair some covalent character in the bond between cation and anion is also present.

An ion triplet may be formed from one cation and two anions or from one anion and two cations. [7] Higher aggregates, such as a tetramer (AB)4, may be formed.

Ternary ion associates involve the association of three species. [8] Another type, named intrusion ion pair, has also been characterized. [9]

Theory

Ions of opposite charge are naturally attracted to each other by the electrostatic force. [10] [11] This is described by Coulomb's law:

where F is the force of attraction, q1 and q2 are the magnitudes of the electrical charges, ε is the dielectric constant of the medium and r is the distance between the ions. For ions in solution this is an approximation because the ions exert a polarizing effect on the solvent molecules that surround them, which attenuates the electric field somewhat. Nevertheless, some general conclusions can be inferred.

Ion association will increase as:
  • the magnitude(s) of the electrical charge(s) q1 and q2 increase,
  • the magnitude of the dielectric constant ε decreases,
  • the size of the ions decreases so that the distance r between cation and anion decreases.

The equilibrium constant K for ion-pair formation, like all equilibrium constants, is related to the standard free-energy change: [12]

where R is the gas constant and T is the temperature in kelvin s. Free energy is made up of an enthalpy term and an entropy term:

The coulombic energy released when ions associate contributes to the enthalpy term, . In the case of contact ion pairs, the covalent interaction energy also contributes to the enthalpy, as does the energy of displacing a solvent molecule from the solvation shell of the cation or anion. The tendency to associate is opposed by the entropy term, which results from the fact that the solution containing unassociated ions is more disordered than a solution containing associates. The entropy term is similar for electrolytes of the same type, with minor differences due to solvation effects. Therefore, it is the magnitude of the enthalpy term that mostly determines the extent of ion association for a given electrolyte type. This explains the general rules given above.

Occurrence

Dielectric constant is the most important factor in determining the occurrence of ion association. A table of some typical values can be found under dielectric constant. Water has a relatively high dielectric constant value of 78.7 at 298K (25 °C), so in aqueous solutions at ambient temperatures 1:1 electrolytes such as NaCl do not form ion pairs to an appreciable extent except when the solution is very concentrated. [13] 2:2 electrolytes (q1 = 2, q2 = 2) form ion pairs more readily. Indeed, the solvent-shared ion pair [Mg(H2O)6]2+SO42− was famously discovered to be present in seawater, in equilibrium with the contact ion pair [Mg(H2O)5(SO4)] [14] Trivalent ions such as Al3+, Fe3+ and lanthanide ions form weak complexes with monovalent anions.

The dielectric constant of water decreases with increasing temperature to about 55 at 100 °C and about 5 at the critical temperature (217.7 °C). [15] Thus ion pairing will become more significant in superheated water.

Solvents with a dielectric constant in the range, roughly, 20–40, show extensive ion-pair formation. For example, in acetonitrile both contact and solvent-shared ion pairs of Li(NCS) have been observed. [16] In methanol the 2:1 electrolyte Mg(NCS)2 is partially dissociated into a contact ion pair, [Mg(NCS)]+ and the thiocyanate ion. [17]

The dielectric constant of liquid ammonia decreases from 26 at its freezing point (80 °C) to 17 at 20 °C (under pressure). Many simple 1:1 electrolytes form contact ion pairs at ambient temperatures. The extent of ion pairing decreases as temperature decreases. With lithium salts there is evidence to show that both inner-sphere and outer-sphere complexes exist in liquid-ammonia solutions. [18]

Of the solvents with dielectric constant of 10 or less, tetrahydrofuran (THF) is particularly relevant in this context, as it solvates cations strongly with the result that simple electrolytes have sufficient solubility to make the study of ion association possible. In this solvent ion association is the rule rather than the exception. Indeed, higher associates such as tetramers are often formed. [19] Triple cations and triple anions have also been characterized in THF solutions. [20]

Ion association is an important factor in phase-transfer catalysis, since a species such as R4P+Cl is formally neutral and so can dissolve easily in a non-polar solvent of low dielectric constant. In this case it also helps that the surface of the cation is hydrophobic.

In SN1 reactions the carbocation intermediate may form an ion pair with an anion, particularly in solvents of low dielectric constant, such as diethylether. [21] This can affect both the kinetic parameters of the reaction and the stereochemistry of the reaction products.

Experimental characterization

Vibrational spectroscopy provides the most widely used means for characterizing ion associates. Both infrared spectroscopy and Raman spectroscopy have been used. Anions containing a CN group, such as cyanide, cyanate and thiocyanide have a vibration frequency a little above 2000 cm−1, which can be easily observed, as the spectra of most solvents (other than nitriles) are weak in this region. The anion vibration frequency is "shifted" on formation of ion pairs and other associates, and the extent of the shift gives information about the nature of the species. Other monovalent anions that have been studied include nitrate, nitrite and azide. Ion pairs of monatomic anions, such as halide ions, cannot be studied by this technique. Standard NMR spectroscopy is not very useful, as association/dissociation reactions tend to be fast on the NMR time scale, giving time-averaged signals of the cation and/or anion. However, diffusion ordered spectroscopy (DOSY), with which the sample tube is not spinning, can be used as ion pairs diffuse more slowly than do single ions due to their greater size. [22]

Nearly the same shift of vibration frequency is observed for solvent-shared ion pairs of LiCN, Be(CN)2 and Al(CN)3 in liquid ammonia. The extent of this type of ion pairing decreases as the size of the cation increases. Thus, solvent-shared ion pairs are characterized by a rather small shift of vibration frequency with respect to the "free" solvated anion, and the value of the shift is not strongly dependent on the nature of the cation. The shift for contact ion pairs is, by contrast, strongly dependent on the nature of the cation and decreases linearly with the ratio of the cations charge to the squared radius: [18]

Cs+ > Rb+ > K+ > Na+ > Li+;
Ba2+ > Sr2+ > Ca2+.

The extent of contact ion pairing can be estimated from the relative intensities of the bands due to the ion pair and free ion. It is greater with the larger cations. [18] This is counter to the trend expected if coulombic energy were the determining factor. Instead, the formation of a contact ion pair is seen to depend more on the energy needed to displace a solvent molecule from the primary solvation sphere of the cation. This energy decreases with the size of the cation, making ion pairing occur to a greater extent with the larger cations. The trend may be different in other solvents. [18]

Higher ion aggregates, sometimes triples M+XM+, sometimes dimers of ion pairs (M+X)2, or even larger species can be identified in the Raman spectra of some liquid-ammonia solutions of Na+ salts by the presence of bands that cannot be attributed to either contact- or solvent-shared ion pairs. [18]

Evidence for the existence of fully solvated ion pairs in solution is mostly indirect, as the spectroscopic properties of such ion pairs are indistinguishable from those of the individual ions. Much of the evidence is based on the interpretation of conductivity measurements. [23] [24]

See also

Related Research Articles

<span class="mw-page-title-main">Acid–base reaction</span> Chemical reaction between an acid and a base

In chemistry, an acid–base reaction is a chemical reaction that occurs between an acid and a base. It can be used to determine pH via titration. Several theoretical frameworks provide alternative conceptions of the reaction mechanisms and their application in solving related problems; these are called the acid–base theories, for example, Brønsted–Lowry acid–base theory.

<span class="mw-page-title-main">Salt (chemistry)</span> Chemical compound involving ionic bonding

In chemistry, a salt or ionic compound is a chemical compound consisting of an ionic assembly of positively charged cations and negatively charged anions, which results in a neutral compound with no net electric charge. The constituent ions are held together by electrostatic forces termed ionic bonds.

<span class="mw-page-title-main">Solvation</span> Association of molecules of a solvent with molecules or ions of a solute

Solvation describes the interaction of a solvent with dissolved molecules. Both ionized and uncharged molecules interact strongly with a solvent, and the strength and nature of this interaction influence many properties of the solute, including solubility, reactivity, and color, as well as influencing the properties of the solvent such as its viscosity and density. If the attractive forces between the solvent and solute particles are greater than the attractive forces holding the solute particles together, the solvent particles pull the solute particles apart and surround them. The surrounded solute particles then move away from the solid solute and out into the solution. Ions are surrounded by a concentric shell of solvent. Solvation is the process of reorganizing solvent and solute molecules into solvation complexes and involves bond formation, hydrogen bonding, and van der Waals forces. Solvation of a solute by water is called hydration.

In chemistry, hydronium (hydroxonium in traditional British English) is the common name for the cation [H3O]+, also written as H3O+, the type of oxonium ion produced by protonation of water. It is often viewed as the positive ion present when an Arrhenius acid is dissolved in water, as Arrhenius acid molecules in solution give up a proton (a positive hydrogen ion, H+) to the surrounding water molecules (H2O). In fact, acids must be surrounded by more than a single water molecule in order to ionize, yielding aqueous H+ and conjugate base. Three main structures for the aqueous proton have garnered experimental support: the Eigen cation, which is a tetrahydrate, H3O+(H2O)3, the Zundel cation, which is a symmetric dihydrate, H+(H2O)2, and the Stoyanov cation, an expanded Zundel cation, which is a hexahydrate: H+(H2O)2(H2O)4. Spectroscopic evidence from well-defined IR spectra overwhelmingly supports the Stoyanov cation as the predominant form. For this reason, it has been suggested that wherever possible, the symbol H+(aq) should be used instead of the hydronium ion.

In chemistry, an acid dissociation constant is a quantitative measure of the strength of an acid in solution. It is the equilibrium constant for a chemical reaction

Solubility equilibrium is a type of dynamic equilibrium that exists when a chemical compound in the solid state is in chemical equilibrium with a solution of that compound. The solid may dissolve unchanged, with dissociation, or with chemical reaction with another constituent of the solution, such as acid or alkali. Each solubility equilibrium is characterized by a temperature-dependent solubility product which functions like an equilibrium constant. Solubility equilibria are important in pharmaceutical, environmental and many other scenarios.

<span class="mw-page-title-main">Aqueous solution</span> Solution in which the solvent is water

An aqueous solution is a solution in which the solvent is water. It is mostly shown in chemical equations by appending (aq) to the relevant chemical formula. For example, a solution of table salt, also known as sodium chloride (NaCl), in water would be represented as Na+(aq) + Cl(aq). The word aqueous means pertaining to, related to, similar to, or dissolved in, water. As water is an excellent solvent and is also naturally abundant, it is a ubiquitous solvent in chemistry. Since water is frequently used as the solvent in experiments, the word solution refers to an aqueous solution, unless the solvent is specified.

<span class="mw-page-title-main">Galvanic cell</span> Electrochemical device

A galvanic cell or voltaic cell, named after the scientists Luigi Galvani and Alessandro Volta, respectively, is an electrochemical cell in which an electric current is generated from spontaneous oxidation–reduction reactions. A common apparatus generally consists of two different metals, each immersed in separate beakers containing their respective metal ions in solution that are connected by a salt bridge or separated by a porous membrane.

<span class="mw-page-title-main">Solvation shell</span> Solvent interface of a solute

A solvation shell or solvation sheath is the solvent interface of any chemical compound or biomolecule that constitutes the solute in a solution. When the solvent is water it is called a hydration shell or hydration sphere. The number of solvent molecules surrounding each unit of solute is called the hydration number of the solute.

In chemistry, the intimate ion pair concept, introduced by Saul Winstein, describes the interactions between a cation, anion and surrounding solvent molecules. In ordinary aqueous solutions of inorganic salts, an ion is completely solvated and shielded from the counterion. In less polar solvents, two ions can still be connected to some extent. In a tight, intimate, or contact ion pair, there are no solvent molecules between the two ions. When solvation increases, ionic bonding decreases and a loose or solvent-shared ion pair results. The ion pair concept explains stereochemistry in solvolysis.

<span class="mw-page-title-main">Counterion</span> Ion which negates another oppositely-charged ion in an ionic molecule

In chemistry, a counterion is the ion that accompanies an ionic species in order to maintain electric neutrality. In table salt the sodium ion is the counterion for the chloride ion and vice versa.

A solvated electron is a free electron in a solution, in which it behaves like an anion. An electron's being solvated in a solution means it is bound by the solution. The notation for a solvated electron in formulas of chemical reactions is "e". Often, discussions of solvated electrons focus on their solutions in ammonia, which are stable for days, but solvated electrons also occur in water and many other solvents – in fact, in any solvent that mediates outer-sphere electron transfer. The solvated electron is responsible for a great deal of radiation chemistry.

The Debye–Hückel theory was proposed by Peter Debye and Erich Hückel as a theoretical explanation for departures from ideality in solutions of electrolytes and plasmas. It is a linearized Poisson–Boltzmann model, which assumes an extremely simplified model of electrolyte solution but nevertheless gave accurate predictions of mean activity coefficients for ions in dilute solution. The Debye–Hückel equation provides a starting point for modern treatments of non-ideality of electrolyte solutions.

In chemistry, the hydron, informally called proton, is the cationic form of atomic hydrogen, represented with the symbol H+
. The general term "hydron", endorsed by the IUPAC, encompasses cations of hydrogen regardless of their isotopic composition: thus it refers collectively to protons (1H+) for the protium isotope, deuterons (2H+ or D+) for the deuterium isotope, and tritons (3H+ or T+) for the tritium isotope.

<span class="mw-page-title-main">Double layer (surface science)</span> Molecular interface between a surface and a fluid

In surface science, a double layer is a structure that appears on the surface of an object when it is exposed to a fluid. The object might be a solid particle, a gas bubble, a liquid droplet, or a porous body. The DL refers to two parallel layers of charge surrounding the object. The first layer, the surface charge, consists of ions which are adsorbed onto the object due to chemical interactions. The second layer is composed of ions attracted to the surface charge via the Coulomb force, electrically screening the first layer. This second layer is loosely associated with the object. It is made of free ions that move in the fluid under the influence of electric attraction and thermal motion rather than being firmly anchored. It is thus called the "diffuse layer".

<span class="mw-page-title-main">Conductivity (electrolytic)</span> Measure of the ability of a solution containing electrolytes to conduct electricity

Conductivity or specific conductance of an electrolyte solution is a measure of its ability to conduct electricity. The SI unit of conductivity is siemens per meter (S/m).

Equilibrium chemistry is concerned with systems in chemical equilibrium. The unifying principle is that the free energy of a system at equilibrium is the minimum possible, so that the slope of the free energy with respect to the reaction coordinate is zero. This principle, applied to mixtures at equilibrium provides a definition of an equilibrium constant. Applications include acid–base, host–guest, metal–complex, solubility, partition, chromatography and redox equilibria.

A metal ion in aqueous solution or aqua ion is a cation, dissolved in water, of chemical formula [M(H2O)n]z+. The solvation number, n, determined by a variety of experimental methods is 4 for Li+ and Be2+ and 6 for most elements in periods 3 and 4 of the periodic table. Lanthanide and actinide aqua ions have higher solvation numbers (often 8 to 9), with the highest known being 11 for Ac3+. The strength of the bonds between the metal ion and water molecules in the primary solvation shell increases with the electrical charge, z, on the metal ion and decreases as its ionic radius, r, increases. Aqua ions are subject to hydrolysis. The logarithm of the first hydrolysis constant is proportional to z2/r for most aqua ions.

In chemistry, ion transport number, also called the transference number, is the fraction of the total electric current carried in an electrolyte by a given ionic species i:

<span class="mw-page-title-main">Hydration number</span> Measure of solvency/solution

The hydration number of a compound is defined as the number of molecules of water bonded to a central ion, often a metal cation. The hydration number is related to the broader concept of solvation number, the number of solvent molecules bonded to a central atom. The hydration number varies with the atom or ion of interest.

References

  1. Davies, C. W. (1962). Ion Association. London: Butterworths.
  2. Wright, Margaret Robson (2007). "Chapter 10: concepts and theory of non-ideality". An introduction to aqueous electrolyte solutions. Wiley. ISBN   978-0-470-84293-5.
  3. Untersuchungen über Ionenassoziation. I. Der Einfluss der Ionenassoziation auf die Aktivität der Ionen bei Mittleren Assoziationsgraden
  4. Earley, J. D.; Zieleniewska, A.; Ripberger, H. H.; Shin, N. Y.; Lazorski, M. S.; Mast, Z. J.; Sayre, H. J.; McCusker, J. K.; Scholes, G. D.; Knowles, R. R.; Reid, O. G. (2022-04-14). "Ion-pair reorganization regulates reactivity in photoredox catalysts". Nature Chemistry. 14 (7): 746–753. Bibcode:2022NatCh..14..746E. doi:10.1038/s41557-022-00911-6. ISSN   1755-4349. PMID   35422457. S2CID   248152234.
  5. Burgess, John (1978). Metal Ions in Solution. Chichester: Ellis Horwood. ISBN   978-0-85312-027-8.Chapter 12, Kinetics and Mechanism: Complex formation"
  6. Burgess, Chapter 5, "Solvation numbers"
  7. Fuoss, R. M.; Kraus, C. A. (1935). "Properties of Electrolytic Solutions. XV. Thermodynamic Properties of Very Weak Electrolytes". J. Am. Chem. Soc. 57: 1–4. doi:10.1021/ja01304a001.
  8. Alexandrov, A.; Kostova, S. (1984). "Extraction-spectrophotometric and radiometric investigation of the ternary ion-association complex of niobium(V) with pyrocatechol and triphenyl-tetrazolium chloride". Journal of Radioanalytical and Nuclear Chemistry . 83 (2): 247–255. doi:10.1007/BF02037138. S2CID   97372470.
  9. Fletcher, R. J.; Gans, P.; Gill, J. B.; Geyer, C. (1997). "Spectrochemistry of solutions. part 29. Intrusion ion pairing: identification of a new form of ion pair in transition metal salt solutions in pyridine through their visible spectra". J. Mol. Liquids. 73–74: 99–106. doi:10.1016/S0167-7322(97)00060-3.
  10. Hans Falkenhagen, Theorie der Elektrolyte, S. Hirzel Verlag, Leipzig, 1971.
  11. S. Petrucci, ed. (2012). "III. Foundations of Modern Statistical Theories". Ionic Interactions: From Dilute Solution to Fused Salts. Physical Chemistry: A Series of Monographs. Vol. 22. Elsevier. p. 424. ISBN   978-0-323-15092-7.
  12. Klotz, I. M. (1964). Chemical Thermodynamics. W. A. Benjamin. Chapter 10.
  13. Assuming that both Na+ and Cl have 6 water molecules in the primary solvation shell at ambient temperatures, a 5 M solution (5 mol/L) will consist almost entirely of fully solvated ion pairs.
  14. Manfred Eigen, Nobel lecture.
  15. Clifford, A. A. "Changes of water properties with temperature". Archived from the original on 2008-02-13. Retrieved 2009-05-02.
  16. Gans, P.; Gill, J. B.; Longdon, P. J. (1989). "Spectrochemistry of solutions. Part 21. Inner- and outer-sphere complexes of lithium with thiocyanate in acetonitrile solutions". J. Chem. Soc. Faraday Trans. I. 85 (7): 1835–1839. doi:10.1039/F19898501835.
  17. Gans, P; Gill, J. B.; Holden, K. M. L. (1994). "Spectrochemistry of solutions. Part 27. Formation of [Mg(NCS)]+ in solutions of Mg(NCS)2 in methanol". J. Chem. Soc., Faraday Trans. 90 (16): 2351–2352. doi:10.1039/FT9949002351.
  18. 1 2 3 4 5 Gill, J. B. (1981). "Solute-solute interactions in liquid ammonia solutions: a vibrational spectroscopic view". Pure Appl. Chem. 53 (7): 1365–1381. doi: 10.1351/pac198153071365 . S2CID   55513823.
  19. Goralski, P.; Chabanel, M. (1987). "Vibrational study of ionic association in aprotic solvents. 11. Formation and structure of mixed aggregates between lithium halides and lithium thiocyanate". Inorg. Chem. 26 (13): 2169–2171. doi:10.1021/ic00260a032.
  20. Bacelon, P.; Corset, J.; de Loze, C. (2004). "Triple ion formation in solutions of alkaline sulfocyanides". J. Solution Chem. 9 (2): 129–139. doi:10.1007/BF00644484. S2CID   93697320. (sulfocyanides = thiocyanates).
  21. Winstein, S.; Clippinger, E.; Fainberg, A. H.; Heck, R.; Robinson G. C. (1956). "Salt Effects and Ion Pairs in Solvolysis and Related Reactions. III.1 Common Ion Rate Depression and Exchange of Anions during Acetolysis". Journal of the American Chemical Society. 78 (2): 328–335. doi:10.1021/ja01583a022.
  22. Pregosin, Paul S. (May 2017). "Applications of NMR diffusion methods with emphasis on ion pairing in inorganic chemistry: a mini-review: Applications of NMR diffusion methods". Magnetic Resonance in Chemistry. 55 (5): 405–413. doi:10.1002/mrc.4394. PMID   26888228. S2CID   3739280.
  23. Raymond M. Fuoss (1957). "Ionic Association. I. Derivation of Constants from Conductance Data". J. Am. Chem. Soc. 79 (13): 3301–3303. doi:10.1021/ja01570a001.
  24. Miyoshi, K. (1973). "Comparison of the Conductance Equations of Fuoss–Onsager, Fuoss–Hsia and Pitts with the Data of Bis(2,9-dimethyl-1,10-phenanthroline)Cu(I) Perchlorate". Bull. Chem. Soc. Jpn. 46 (2): 426–430. doi: 10.1246/bcsj.46.426 .