Proton affinity

Last updated

The proton affinity (PA, Epa) of an anion or of a neutral atom or molecule is the negative of the enthalpy change in the reaction between the chemical species concerned and a proton in the gas phase: [1]

Contents

These reactions are always exothermic in the gas phase, i.e. energy is released (enthalpy is negative) when the reaction advances in the direction shown above, while the proton affinity is positive. This is the same sign convention used for electron affinity. The property related to the proton affinity is the gas-phase basicity, which is the negative of the Gibbs energy for above reactions, [2] i.e. the gas-phase basicity includes entropic terms in contrast to the proton affinity.

Acid/base chemistry

The higher the proton affinity, the stronger the base and the weaker the conjugate acid in the gas phase. The (reportedly) strongest known base is the ortho-diethynylbenzene dianion (Epa = 1843 kJ/mol), [3] followed by the methanide anion (Epa = 1743 kJ/mol) and the hydride ion (Epa = 1675 kJ/mol), [4] making methane the weakest proton acid [5] in the gas phase, followed by dihydrogen. The weakest known base is the helium atom (Epa = 177.8 kJ/mol), [6] making the hydrohelium(1+) ion the strongest known proton acid.

Hydration

Proton affinities illustrate the role of hydration in aqueous-phase Brønsted acidity. Hydrofluoric acid is a weak acid in aqueous solution (pKa = 3.15) [7] but a very weak acid in the gas phase (Epa (F) = 1554 kJ/mol): [4] the fluoride ion is as strong a base as SiH3 in the gas phase, but its basicity is reduced in aqueous solution because it is strongly hydrated, and therefore stabilized. The contrast is even more marked for the hydroxide ion (Epa = 1635 kJ/mol), [4] one of the strongest known proton acceptors in the gas phase. Suspensions of potassium hydroxide in dimethyl sulfoxide (which does not solvate the hydroxide ion as strongly as water) are markedly more basic than aqueous solutions, and are capable of deprotonating such weak acids as triphenylmethane (pKa = ca. 30). [8] [9]

To a first approximation, the proton affinity of a base in the gas phase can be seen as offsetting (usually only partially) the extremely favorable hydration energy of the gaseous proton (ΔE = 1530 kJ/mol), as can be seen in the following estimates of aqueous acidity:

Proton affinityHHe+(g)H+(g)+ He(g)+178 kJ/mol [6]    HF(g)H+(g)+ F(g)+1554 kJ/mol [4]    H2(g)H+(g)+ H(g)+1675 kJ/mol [4]
Hydration of acidHHe+(aq)HHe+(g) +973 kJ/mol [10]  HF(aq)HF(g) +23 kJ/mol [7]  H2(aq)H2(g) 18 kJ/mol [11]
Hydration of protonH+(g)H+(aq) 1530 kJ/mol [7]  H+(g)H+(aq) 1530 kJ/mol [7]  H+(g)H+(aq) 1530 kJ/mol [7]
Hydration of baseHe(g)He(aq) +19 kJ/mol [11]  F(g)F(aq) 13 kJ/mol [7]  H(g)H(aq) +79 kJ/mol [7]
Dissociation equilibrium  HHe+(aq)H+(aq)+ He(aq)360 kJ/mol  HF(aq)H+(aq)+ F(aq)+34 kJ/mol  H2(aq)H+(aq)+ H(aq)+206 kJ/mol 
Estimated pKa63 +6 +36

These estimates suffer from the fact the free energy change of dissociation is in effect the small difference of two large numbers. However, hydrofluoric acid is correctly predicted to be a weak acid in aqueous solution and the estimated value for the pKa of dihydrogen is in agreement with the behaviour of saline hydrides (e.g., sodium hydride) when used in organic synthesis.

Difference from pKa

Both proton affinity and pKa are measures of the acidity of a molecule, and so both reflect the thermodynamic gradient between a molecule and the anionic form of that molecule upon removal of a proton from it. Implicit in the definition of pKa however is that the acceptor of this proton is water, and an equilibrium is being established between the molecule and bulk solution. More broadly, pKa can be defined with reference to any solvent, and many weak organic acids have measured pKa values in DMSO. Large discrepancies between pKa values in water versus DMSO (i.e., the pKa of water in water is 14, [12] [13] but water in DMSO is 32) demonstrate that the solvent is an active partner in the proton equilibrium process, and so pKa does not represent an intrinsic property of the molecule in isolation. In contrast, proton affinity is an intrinsic property of the molecule, without explicit reference to the solvent.

A second difference arises in noting that pKa reflects a thermal free energy for the proton transfer process, in which both enthalpic and entropic terms are considered together. Therefore, pKa is influenced both by the stability of the molecular anion, as well as the entropy associated of forming and mixing new species. Proton affinity, on the other hand, is not a measure of free energy.

List of compound affinities

Proton affinities are quoted in kJ/mol, in increasing order of gas-phase basicity of the base.

Proton affinity [14]
Base Affinity
(kJ/mol)
Neutral molecules
Helium 178
Neon 201
Argon 371
Dioxygen 422
Dihydrogen 424
Krypton 425
Hydrogen fluoride 490
Dinitrogen 495
Xenon 496
Nitric oxide 531
Carbon dioxide 548
Methane 552
Hydrogen chloride 564
Hydrogen bromide 569
Nitrous oxide 571
Sulfur trioxide 589 [15]
Carbon monoxide 594
Ethane 601
Nitrogen trifluoride 602
Hydrogen iodide 628
Carbonyl sulfide 632
Acetylene 641
Arsenic trifluoride 649
Silane 649
Sulfur dioxide 676
Hydrogen peroxide 678
Ethylene 680
Phosphorus trifluoride 697
Water 697
Carbon disulfide 699
Phosphoryl trifluoride 702
2,4-Dicarba-closo-heptaborane(7) 703
Hydrogen sulfide 712
Hydrogen selenide 717
Hydrogen cyanide 717
Formaldehyde 718
Carbon monosulfide 732
Cyanogen chloride 735
Arsine 750
Benzene 759
Methanol 761
Methanethiol 784
Ethanol 788
Acetonitrile 788
Phosphine 789
Nitrogen trichloride 791
Ethanethiol 798
Propanol 798
Propane-1-thiol 802
Hydroxylamine 803
Dimethyl ether 804
Glyceryl phosphite 812
Borazine 812
Acetone 823
Diethyl ether 838
Dimethyl sulfide 839
Iron pentacarbonyl 845
Ammonia 854
Methylphosphine 854
Hydrazine 856
Diethyl sulfide 858
1,6-Dicarba-closo-hexaborane(6) 866
Aniline 877
P(OCH2)3CCH3877
Ferrocene 877
Dimethyl sulfoxide 884
Dimethyl formamide 884
Trimethyl phosphate 887
Trimethylarsine 893
Methylamine 896
Tri-O-methyl thiophosphate 897
Dimethylphosphine 905
Trimethyl phosphite 923
Dimethylamine 923
Pyridine 924
Trimethylamine 942
Trimethylphosphine 950
Triethylphosphine 969
Triethylamine 972
Lithium hydroxide 1008
Sodium hydroxide 1038
Potassium hydroxide 1100
Caesium hydroxide 1125
Anions
Trioxophosphate(1) 1301
Iodide 1315
Pentacarbonylmanganate(1) 1326
Trifluoroacetate 1350
Bromide 1354
Nitrate 1358
Pentacarbonylrhenate(1) 1389
Chloride 1395
Nitrite 1415
Hydroselenide 1417
Formate 1444
Acetate 1458
Phenoxide 1470
Cyanide 1477
Hydrosulfide 1477
Cyclopentadienide 1490
Ethanethiolate 1495
Nitromethanide 1501
Arsinide 1502
Methanethiolate 1502
Germanide 1509
Trichloromethanide 1515
Formylmethanide 1533
Methylsulfonylmethanide 1534
Anilide 1536
Acetonide 1543
Phosphinide 1550
Silanide 1554
Fluoride 1554
Cyanomethanide 1557
Propoxide 1568
Acetylide 1571
Trifluoromethanide 1572
Ethoxide 1574
Phenylmethanide 1586
Methoxide 1587
Hydroxide 1635
Amide 1672
Hydride 1675
Methanide 1743

Related Research Articles

<span class="mw-page-title-main">Acid</span> Chemical compound giving a proton or accepting an electron pair

An acid is a molecule or ion capable of either donating a proton (i.e. hydrogen ion, H+), known as a Brønsted–Lowry acid, or forming a covalent bond with an electron pair, known as a Lewis acid.

<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">Hydroxide</span> Chemical compound

Hydroxide is a diatomic anion with chemical formula OH. It consists of an oxygen and hydrogen atom held together by a single covalent bond, and carries a negative electric charge. It is an important but usually minor constituent of water. It functions as a base, a ligand, a nucleophile, and a catalyst. The hydroxide ion forms salts, some of which dissociate in aqueous solution, liberating solvated hydroxide ions. Sodium hydroxide is a multi-million-ton per annum commodity chemical. The corresponding electrically neutral compound HO is the hydroxyl radical. The corresponding covalently bound group –OH of atoms is the hydroxy group. Both the hydroxide ion and hydroxy group are nucleophiles and can act as catalysts in organic chemistry.

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

<span class="mw-page-title-main">Base (chemistry)</span> Type of chemical substance

In chemistry, there are three definitions in common use of the word "base": Arrhenius bases, Brønsted bases, and Lewis bases. All definitions agree that bases are substances that react with acids, as originally proposed by G.-F. Rouelle in the mid-18th century.

In organic chemistry, a carbanion is an anion in which carbon is negatively charged.

Deprotonation (or dehydronation) is the removal (transfer) of a proton (or hydron, or hydrogen cation), (H+) from a Brønsted–Lowry acid in an acid–base reaction. The species formed is the conjugate base of that acid. The complementary process, when a proton is added (transferred) to a Brønsted–Lowry base, is protonation (or hydronation). The species formed is the conjugate acid of that base.

<span class="mw-page-title-main">Neutralization (chemistry)</span> Chemical reaction in which an acid and a base react quantitatively

In chemistry, neutralization or neutralisation is a chemical reaction in which acid and a base react with an equivalent quantity of each other. In a reaction in water, neutralization results in there being no excess of hydrogen or hydroxide ions present in the solution. The pH of the neutralized solution depends on the acid strength of the reactants.

The Brønsted–Lowry theory (also called proton theory of acids and bases) is an acid–base reaction theory which was first developed by Johannes Nicolaus Brønsted and Thomas Martin Lowry independently in 1923. The basic concept of this theory is that when an acid and a base react with each other, the acid forms its conjugate base, and the base forms its conjugate acid by exchange of a proton (the hydrogen cation, or H+). This theory generalises the Arrhenius theory.

<span class="mw-page-title-main">1,8-Bis(dimethylamino)naphthalene</span> Chemical compound

1,8-Bis(dimethylamino)naphthalene is an organic compound with the formula C10H6(NMe2)2 (Me = methyl). It is classified as a peri-naphthalene, i.e. a 1,8-disubstituted derivative of naphthalene. Owing to its unusual structure, it exhibits exceptional basicity. It is often referred by the trade name Proton Sponge, a trademark of Sigma-Aldrich.

<span class="mw-page-title-main">Dissociation (chemistry)</span> Separation of molecules or ionic compounds into smaller constituent entities

Dissociation in chemistry is a general process in which molecules (or ionic compounds such as salts, or complexes) separate or split into other things such as atoms, ions, or radicals, usually in a reversible manner. For instance, when an acid dissolves in water, a covalent bond between an electronegative atom and a hydrogen atom is broken by heterolytic fission, which gives a proton (H+) and a negative ion. Dissociation is the opposite of association or recombination.

<span class="mw-page-title-main">Hydrogen iodide</span> Chemical compound

Hydrogen iodide (HI) is a diatomic molecule and hydrogen halide. Aqueous solutions of HI are known as hydroiodic acid or hydriodic acid, a strong acid. Hydrogen iodide and hydroiodic acid are, however, different in that the former is a gas under standard conditions, whereas the other is an aqueous solution of the gas. They are interconvertible. HI is used in organic and inorganic synthesis as one of the primary sources of iodine and as a reducing agent.

<span class="mw-page-title-main">Fluoroantimonic acid</span> Chemical compound

Fluoroantimonic acid is a mixture of hydrogen fluoride and antimony penta­fluoride, containing various cations and anions. This mixture is a superacid that, in terms of corrosiveness, is trillions of times stronger than pure sulfuric acid when measured by its Hammett acidity function. It even protonates some hydro­carbons to afford pentacoordinate carbo­cations. Like its precursor hydrogen fluoride, it attacks glass, but can be stored in containers lined with PTFE (Teflon) or PFA.

<span class="mw-page-title-main">Nitroxyl</span> Chemical compound

Nitroxyl or azanone is the chemical compound HNO. It is well known in the gas phase. Nitroxyl can be formed as a short-lived intermediate in the solution phase. The conjugate base, NO, nitroxide anion, is the reduced form of nitric oxide (NO) and is isoelectronic with dioxygen. The bond dissociation energy of H−NO is 49.5 kcal/mol (207 kJ/mol), which is unusually weak for a bond to the hydrogen atom.

An acidity function is a measure of the acidity of a medium or solvent system, usually expressed in terms of its ability to donate protons to a solute. The pH scale is by far the most commonly used acidity function, and is ideal for dilute aqueous solutions. Other acidity functions have been proposed for different environments, most notably the Hammett acidity function, H0, for superacid media and its modified version H for superbasic media. The term acidity function is also used for measurements made on basic systems, and the term basicity function is uncommon.

In chemistry, solvent effects are the influence of a solvent on chemical reactivity or molecular associations. Solvents can have an effect on solubility, stability and reaction rates and choosing the appropriate solvent allows for thermodynamic and kinetic control over a chemical reaction.

Transition metal oxides are compounds composed of oxygen atoms bound to transition metals. They are commonly utilized for their catalytic activity and semiconducting properties. Transition metal oxides are also frequently used as pigments in paints and plastics, most notably titanium dioxide. Transition metal oxides have a wide variety of surface structures which affect the surface energy of these compounds and influence their chemical properties. The relative acidity and basicity of the atoms present on the surface of metal oxides are also affected by the coordination of the metal cation and oxygen anion, which alter the catalytic properties of these compounds. For this reason, structural defects in transition metal oxides greatly influence their catalytic properties. The acidic and basic sites on the surface of metal oxides are commonly characterized via infrared spectroscopy, calorimetry among other techniques. Transition metal oxides can also undergo photo-assisted adsorption and desorption that alter their electrical conductivity. One of the more researched properties of these compounds is their response to electromagnetic radiation, which makes them useful catalysts for redox reactions, isotope exchange and specialized surfaces.

Acid strength is the tendency of an acid, symbolised by the chemical formula , to dissociate into a proton, , and an anion, . The dissociation of a strong acid in solution is effectively complete, except in its most concentrated solutions.

<span class="mw-page-title-main">Carborane acid</span> Class of chemical compounds

Carborane acidsH(CXB
11
Y
5
Z
6
)
(X, Y, Z = H, Alk, F, Cl, Br, CF3) are a class of superacids, some of which are estimated to be at least one million times stronger than 100% pure sulfuric acid in terms of their Hammett acidity function values (H0 ≤ –18) and possess computed pKa values well below –20, establishing them as some of the strongest known Brønsted acids. The best-studied example is the highly chlorinated derivative H(CHB
11
Cl
11
)
. The acidity of H(CHB
11
Cl
11
)
was found to vastly exceed that of triflic acid, CF
3
SO
3
H
, and bistriflimide, (CF
3
SO
2
)
2
NH
, compounds previously regarded as the strongest isolable acids.

References

  1. "Proton affinity." Compendium of Chemical Terminology .
  2. "Gas-phase basicity." Compendium of Chemical Terminology .
  3. Poad, Berwyck L. J.; Reed, Nicholas D.; Hansen, Christopher S.; Trevitt, Adam J.; Blanksby, Stephen J.; MacKay, Emily G.; Sherburn, Michael S.; Chan, Bun; Radom, Leo (2016). "Preparation of an ion with the highest calculated proton affinity: ortho-diethynylbenzene dianion". Chem. Sci. 7 (9): 6245–6250. doi:10.1039/C6SC01726F. PMC   6024202 . PMID   30034765.
  4. 1 2 3 4 5 Bartmess, J. E.; Scott, J. A.; McIver, R. T. (1979). "Scale of acidities in the gas phase from methanol to phenol". J. Am. Chem. Soc. 101 (20): 6046. doi:10.1021/ja00514a030.
  5. The term "proton acid" is used to distinguish these acids from Lewis acids. It is the gas-phase equivalent of the term Brønsted acid.
  6. 1 2 Lias, S. G.; Liebman, J. F.; Levin, R. D. (1984). Title J. Phys. Chem. Ref. Data. 13':695.
  7. 1 2 3 4 5 6 7 Jolly, William L. (1991). Modern Inorganic Chemistry (2nd Edn.). New York: McGraw-Hill. ISBN   0-07-112651-1.
  8. Jolly, William L (1967). "The intrinsic basicity of the hydroxide ion". J. Chem. Educ. 44 (5): 304. Bibcode:1967JChEd..44..304J. doi:10.1021/ed044p304.
  9. Jolly, William L (1968). "σ‐Methyl‐π‐Cyclopentadienylmolybdenum Tricarbonyl". Inorg. Synth. 11: 113. doi:10.1002/9780470132425.ch22. ISBN   9780470132425.
  10. Estimated to be the same as for Li+(aq) → Li+(g).
  11. 1 2 Estimated from solubility data.
  12. Meister, Erich C.; Willeke, Martin; Angst, Werner; Togni, Antonio; Walde, Peter (2014). "Confusing Quantitative Descriptions of Brønsted-Lowry Acid-Base Equilibria in Chemistry Textbooks – A Critical Review and Clarifications for Chemical Educators". Helvetica Chimica Acta. 97 (1): 1–31. doi:10.1002/hlca.201300321. ISSN   1522-2675.
  13. Silverstein, Todd P.; Heller, Stephen T. (2017-06-13). "pKa Values in the Undergraduate Curriculum: What Is the Real pKa of Water?". Journal of Chemical Education. 94 (6): 690–695. Bibcode:2017JChEd..94..690S. doi:10.1021/acs.jchemed.6b00623. ISSN   0021-9584.
  14. Jolly, William L. (1991). Modern Inorganic Chemistry (2nd Edn.). New York: McGraw-Hill. ISBN   0-07-112651-1
  15. "Proton affinity of SO3".