Electron affinities
   HOME

TheInfoList



OR:

The electron affinity (''E''ea) of an atom or molecule is defined as the amount of energy released when an electron attaches to a neutral atom or molecule in the gaseous state to form an anion. ::X(g) + e → X(g) + energy Note that this is not the same as the enthalpy change of electron capture ionization, which is defined as negative when energy is released. In other words, the enthalpy change and the electron affinity differ by a negative sign. In solid state physics, the electron affinity for a surface is defined somewhat differently (Electron affinity#"Electron affinity" as defined in solid state physics, see below).


Measurement and use of electron affinity

This property is used to measure atoms and molecules in the gaseous state only, since in a solid or liquid state their energy levels would be changed by contact with other atoms or molecules. A list of the electron affinities was used by Robert S. Mulliken to develop an electronegativity scale for atoms, equal to the average of the electrons affinity and ionization potential. Other theoretical concepts that use electron affinity include electronic chemical potential and chemical hardness. Another example, a molecule or atom that has a more positive value of electron affinity than another is often called an electron acceptor and the less positive an electron donor. Together they may undergo Intervalence charge transfer, charge-transfer reactions.


Sign convention

To use electron affinities properly, it is essential to keep track of sign. For any reaction that ''releases'' energy, the ''change'' Δ''E'' in total energy has a negative value and the reaction is called an Exothermic reaction, exothermic process. Electron capture for almost all non-noble gas atoms involves the release of energy and thus are exothermic. The positive values that are listed in tables of ''E''ea are amounts or magnitudes. It is the word "released" within the definition "energy released" that supplies the negative sign to Δ''E''. Confusion arises in mistaking ''E''ea for a change in energy, Δ''E'', in which case the positive values listed in tables would be for an endo- not exo-thermic process. The relation between the two is ''E''ea = −Δ''E''(attach). However, if the value assigned to ''E''ea is negative, the negative sign implies a reversal of direction, and energy is ''required'' to attach an electron. In this case, the electron capture is an endothermic process and the relationship, ''E''ea = −Δ''E''(attach) is still valid. Negative values typically arise for the capture of a second electron, but also for the nitrogen atom. The usual expression for calculating ''E''ea when an electron is attached is : This expression does follow the convention Δ''X'' = ''X''(final) − ''X''(initial) since −Δ''E'' = −(''E''(final) − ''E''(initial)) = ''E''(initial) − ''E''(final). Equivalently, electron affinity can also be defined as the amount of energy ''required'' to detach an electron from the atom while it holds a Electric charge, single-excess-electron thus making the atom a ion, negative ion, i.e. the energy change for the process :X → X + e If the same table is employed for the forward and reverse reactions, ''without switching signs'', care must be taken to apply the correct definition to the corresponding direction, attachment (release) or detachment (require). Since almost all detachments ''(require +)'' an amount of energy listed on the table, those detachment reactions are endothermic, or Δ''E''(detach) > 0. :


Electron affinities of the elements

Although ''E''ea varies greatly across the periodic table, some patterns emerge. Generally, nonmetals have more positive ''E''ea than metals. Atoms whose anions are more stable than neutral atoms have a greater ''E''ea. Chlorine most strongly attracts extra electrons; neon most weakly attracts an extra electron. The electron affinities of the noble gases have not been conclusively measured, so they may or may not have slightly negative values. ''E''ea generally increases across a period (row) in the periodic table prior to reaching group 18. This is caused by the filling of the valence shell of the atom; a halogen, group 17 atom releases more energy than a Group 1 element, group 1 atom on gaining an electron because it obtains a filled electron shell, valence shell and therefore is more stable. In group 18, the valence shell is full, meaning that added electrons are unstable, tending to be ejected very quickly. Counterintuitively, ''E''ea does ''not'' decrease when progressing down most columns of the periodic table. For example, ''E''ea actually increases consistently on descending the column for the Group 2 element, group 2 data. Thus, electron affinity follows the same "left-right" trend as electronegativity, but not the "up-down" trend. The following data are quoted in Joule per mole, kJ/mol.


Molecular electron affinities

The electron affinity of molecules is a complicated function of their electronic structure. For instance the electron affinity for benzene is negative, as is that of naphthalene, while those of anthracene, phenanthrene and pyrene are positive. ''In silico'' experiments show that the electron affinity of hexacyanobenzene surpasses that of fullerene.


"Electron affinity" as defined in solid state physics

In the field of solid state physics, the electron affinity is defined differently than in chemistry and atomic physics. For a semiconductor-vacuum interface (that is, the surface of a semiconductor), electron affinity, typically denoted by ''E''EA or ''χ'', is defined as the energy obtained by moving an electron from the vacuum just outside the semiconductor to the bottom of the conduction band just inside the semiconductor: :E_ \equiv E_ - E_ In an intrinsic semiconductor at absolute zero, this concept is functionally analogous to the chemistry definition of electron affinity, since an added electron will spontaneously go to the bottom of the conduction band. At nonzero temperature, and for other materials (metals, semimetals, heavily doped semiconductors), the analogy does not hold since an added electron will instead go to the Fermi level on average. In any case, the value of the electron affinity of a solid substance is very different from the chemistry and atomic physics electron affinity value for an atom of the same substance in gas phase. For example, a silicon crystal surface has electron affinity 4.05 eV, whereas an isolated silicon atom has electron affinity 1.39 eV. The electron affinity of a surface is closely related to, but distinct from, its work function. The work function is the thermodynamic work that can be obtained by reversibly and isothermally removing an electron from the material to vacuum; this thermodynamic electron goes to the ''Fermi level'' on average, not the conduction band edge: W = E_ - E_. While the work function of a semiconductor can be changed by doping (semiconductor), doping, the electron affinity ideally does not change with doping and so it is closer to being a material constant. However, like work function the electron affinity does depend on the surface termination (crystal face, surface chemistry, etc.) and is strictly a surface property. In semiconductor physics, the primary use of the electron affinity is not actually in the analysis of semiconductor–vacuum surfaces, but rather in heuristic electron affinity rules for estimating the band bending that occurs at the interface of two materials, in particular metal–semiconductor junctions and semiconductor heterojunctions. In certain circumstances, the electron affinity may become negative. Often negative electron affinity is desired to obtain efficient cathodes that can supply electrons to the vacuum with little energy loss. The observed electron yield as a function of various parameters such as bias voltage or illumination conditions can be used to describe these structures with band diagrams in which the electron affinity is one parameter. For one illustration of the apparent effect of surface termination on electron emission, see Figure 3 in Marchywka Effect.


See also

* Ionization energy — a closely related concept describing the energy required to ''remove'' an electron from a neutral atom or molecule * One-electron reduction * Electron-capture mass spectrometry * Electronegativity * Valence electron * Vacuum level * Electron donor


References

*Tro, Nivaldo J. (2008). ''Chemistry: A Molecular Approach'' (2nd Edn.). New Jersey: Pearson Prentice Hall. . pp. 348–349.


External links


Electron affinity
definition from the IUPAC Gold Book {{Authority control Atomic physics Chemical bonding Chemical properties Ions Molecular physics Physical chemistry