Proton affinity

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]

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 affinity HHe+(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 acid HHe+(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 proton H+(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 base He(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 pKa 63   +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]
BaseAffinity
(kJ/mol)
Neutral molecules
Helium178
Neon201
Argon371
Dioxygen422
Dihydrogen424
Krypton425
Hydrogen fluoride490
Dinitrogen495
Xenon496
Nitric oxide531
Carbon dioxide548
Methane552
Hydrogen chloride564
Hydrogen bromide569
Nitrous oxide571
Sulfur trioxide589[15]
Carbon monoxide594
Ethane601
Nitrogen trifluoride602
Hydrogen iodide628
Carbonyl sulfide632
Acetylene641
Arsenic trifluoride649
Silane649
Sulfur dioxide676
Hydrogen peroxide678
Ethylene680
Phosphorus trifluoride697
Water697
Carbon disulfide699
Phosphoryl trifluoride702
2,4-Dicarba-closo-heptaborane(7)703
Hydrogen sulfide712
Hydrogen selenide717
Hydrogen cyanide717
Formaldehyde718
Carbon monosulfide732
Cyanogen chloride735
Arsine750
Benzene759
Methanol761
Methanethiol784
Ethanol788
Acetonitrile788
Phosphine789
Nitrogen trichloride791
Ethanethiol798
Propanol798
Propane-1-thiol802
Hydroxylamine803
Dimethyl ether804
Glyceryl phosphite812
Borazine812
Acetone823
Diethyl ether838
Dimethyl sulfide839
Iron pentacarbonyl845
Ammonia854
Methylphosphine854
Hydrazine856
Diethyl sulfide858
1,6-Dicarba-closo-hexaborane(6)866
Aniline877
P(OCH2)3CCH3877
Ferrocene877
Dimethyl sulfoxide884
Dimethyl formamide884
Trimethyl phosphate887
Trimethylarsine893
Methylamine896
Tri-O-methyl thiophosphate897
Dimethylphosphine905
Trimethyl phosphite923
Dimethylamine923
Pyridine924
Trimethylamine942
Trimethylphosphine950
Triethylphosphine969
Triethylamine972
Lithium hydroxide1008
Sodium hydroxide1038
Potassium hydroxide1100
Caesium hydroxide1125
Anions
Trioxophosphate(1)1301
Iodide1315
Pentacarbonylmanganate(1)1326
Trifluoroacetate1350
Bromide1354
Nitrate1358
Pentacarbonylrhenate(1)1389
Chloride1395
Nitrite1415
Hydroselenide1417
Formate1444
Acetate1458
Phenoxide1470
Cyanide1477
Hydrosulfide1477
Cyclopentadienide1490
Ethanethiolate1495
Nitromethanide1501
Arsinide1502
Methanethiolate1502
Germanide1509
Trichloromethanide1515
Formylmethanide1533
Methylsulfonylmethanide1534
Anilide1536
Acetonide1543
Phosphinide1550
Silanide1554
Fluoride1554
Cyanomethanide1557
Propoxide1568
Acetylide1571
Trifluoromethanide1572
Ethoxide1574
Phenylmethanide1586
Methoxide1587
Hydroxide1635
Amide1672
Hydride1675
Methanide1743

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. 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. Lias, S. G.; Liebman, J. F.; Levin, R. D. (1984). Title J. Phys. Chem. Ref. Data. 13':695.
  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". Inorganic Syntheses. p. 113. doi:10.1002/9780470132425.ch22. ISBN 9780470132425. {{cite book}}: |journal= ignored (help)
  10. Estimated to be the same as for Li+(aq) → Li+(g).
  11. 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".
This article is issued from Wikipedia. The text is licensed under Creative Commons - Attribution - Sharealike. Additional terms may apply for the media files.