You are on page 1of 25

What is Electronegativity?

Juan Camilo Martínez González


Universidad Nacional de Tres de Febrero, Buenos Aires, Argentina.

Klaus Ruthenberg
Coburg University of Applied Sciences and Arts, Coburg, Germany.

Alles, was uns eine Wirkung der sogenannten Wahlverwandschaft


(Affinitas electiva) zu seyn scheint, kann mithin nichts anderes
seyn, als die Wirkung der, bei manchen Körpern stärker, als bei
andern, hervortretenden elektrischen Polarität.
J. J. Berzelius, 1820

Introduction

1. Sketch of the prehistory of electronegativity

“Our experience about the mutual electrical behaviour of the bodies has shown to us that the latter
can be separated into two classes, namely into electropositive and electronegative ones”. Jöns Jacob
Berzelius (1779-1848), one of the most influential chemists of the 19th century and beyond, wrote
these words in a treatise on chemical proportions, atomic weights, and the relation between
electricity and stuff behaviour (Berzelius, 1820, p. 81 1). Already before the early 19th century,
several scientists considered chemical phenomena to be in fact of electrical nature. In 1798, for
example, Johann Wilhelm Ritter (1776-1810), one of the pioneers of electrochemistry, speculated
about whether the “system of electricity” and the “system of chemistry” might become one in the
future (s. Ostwald, 1896, p. 70). According to Berzelius, electricity is a kind of imponderable, liquid
matter: “This stuff has no property by which it would be perceivable for our senses, rather it
becomes manifest only if it is dismembered into its constituents.” (Berzelius, 1825, p. 70 2) These
constituents – positive and negative electricity – can be differentiated by their different behaviour or
properties. As to the latter, Berzelius intriguingly mentions, among others, the sour taste of positive
electricity and the basic taste (alkalinischer Geschmack) of negative electricity – both are obtainable
by bringing the particular electrode into contact with the tongue. Well in accordance with this
1
“Unsere Erfahrungen über das gegenseitige Verhalten der Körper haben uns gelehrt, daß letztere sich in zwei
Klassen, nämlich in electropositive und in electronegative eintheilen lassen.“ (emphases original)
2
“Dieser Stoff besitzt keine Eigenschaft, wodurch er für unsere äußeren Sinne wahrnehmbar wird, sondern er
offenbart sich erst dann, wenn er in seine Bestandteile zerlegt wird.“ Though it does not look like this on the face of it,
Berzelius was quite reluctant with respect to speculations about the nature of electricity (cf. Russell, 1963). Note that
the picture of electricity as liquid matter was considered even earlier, for example by Descartes (cf. Farber, 1952, p.
216).

1
description, the change of colour of litmus paper is reported as follows: “ + E” yields red, “- E” blue
(p. 71). The latter observations show the conceptual vicinity of Berzelius´ dual theory of chemical
constitution on the one hand and the the theory of acidity of the same scholar on the other. Like
acidity, electricity is a polar (dual) theoretical conception. During the electrolytic experiments with
the voltaic pole the acids appear at one electrode, and the bases at the opposite. Berzelius described
the elements as consisting of more or less of the two sorts of electricity, and he interpreted the
chemical behaviour with explicit reference to electricity: “Everything that seems to be an effect of
the so-called elective affinity (Affinitas electiva) cannot be something else than the impact of, for
some bodies stronger than for others, obtruding electrical polarity” (Berzelius, 1820, p. 98,
emphasis original3).

Like Amedeo Avogadro (1776-1856) – with his oxygenicity scale – before him, Berzelius tried to
align the known elements with respect to their chemical affinity. It is important to be aware that –
like Avogadro – Berzelius gave no operational description or criteria how to get to his scale. For
comparison, we present the succession of the first elements of a later version from 1831 (cf.
Russell, 1963, p. 1304), the non-metal part of the electrochemical series, and the elemental order
referring to modern electronegativity values (after Allred and Rochow) in one graphical
representation (Fig. 1):

Fig. 1 Comparison of the elemental order according to Berzelius (dark line; O = oxygen = 1; F =
fluorine = 4), the electrochemical series (medium dark line; O = 2; F = 1), and the modern
electronegativity sequence according to calculations of Allred and Rochow (light line; O = 2; F =
1). Sulphur, for example, is no. 2 in the Berzelian, no. 6 in the electrochemical series (cf. Lowry and
Cavell, 19495), and no. 7 in the modern electronegativity scale after Allred and Rochow.

According to that picture, for example, Arsenik (no. 10 for Berzelius in 1831) contains a bigger
proportion of positive electricity than, for example, nitrogen. Berzelius is obviously not taking into
account what later will be called the “law of electroneutrality”.

Although there is a rough similarity between the “classical” and the modern alignment – typical
non-metals are electronegative, the more metallic elements electropositive – it is not possible to
conclude that both sequences are referring to the same underlying regularity or “law”: Even a
3
See our motto.
4
Berzelius was improving that scale during his whole scientific life. Russell 1963 gives a valuable overview.
5
In a more or less classical (“Berzelian”) vein, the authors claim:“By means of electrical measurements it is
possible to arrange the elements in a electrochemical series in which the metals are placed in descending order of
electropositiveness, and the non-metals in ascending order of electronegativiness” (Lowry and Cavell 1949, p. 53).

2
comparison of, say, boiling points, would yield quite similar trends. 6 In a paper on the origins of the
electronegativity concept, the American historian of chemistry William Jensen claims the following:
“... whatever electronegativity really is – and the 30 or more definitions currently in the literature
still haven´t given us a conclusive answer – it really does reflect some intuitive chemical reality – a
reality already perceived in its essentials by a master chemist more than 150 years ago.” (Jensen,
1996, p. 18) He seems to hold that there is a straight, linear and somewhat uniform development
with respect to electronegativity. We do not subscribe to this claim. What Berzelius describes is not
commensurable with what modern chemistry is calling electronegativity although he uses the same
word. This incommensurability is at least threefold. Firstly, the assumed theory of a dual kind of
liquid electrical matter is only in part and formally similar to the modern views of electricity –
roughly speaking: in both pictures there is a plus and a minus. Berzelius of course has no access to
electrons or protons, and he is not explicitly talking about chemical bonding. Rather he is
ingeniously speculating about the electrical nature of chemical substances and their affinity or
reactivity. Secondly, the operational part of his achievements remains nebulous. Other than the
chemists since Pauling, which will be shown in the following sections, he is depending on his huge
chemical experience and intuition only. Thirdly, and not surprisingly, not even the results show
correlations that could be considered non-ambiguous (see Fig. 1). Only if we would take a whiggish
point of view, we might come to the conclusion that Berzelius on the one hand, and for example
Mulliken or Pauling, on the other, speak about the same property or disposition. As to this stance,
our modern picture of substances, acidity, radicals, periodicity, reactivity, and so on, is of course the
only correct one and chemical elements and their periodic classification are simply natural kinds.

Accordingly, Berzelius´ electronegativity concept should be considered part of the merely loose
prehistory rather than an integral part of a linear scientific development.

2. Current ambiguity of electronegativity in quantum chemistry

The modern electronegativity concept is only possible within a theory of chemical bond that
contains electrons. The American chemist Gilbert Lewis (1875-1946) was the first to discuss
chemical bonds as structures consisting of shared electron pairs. Already in his first foundational
texts of modern chemistry, Lewis employed the terms ‘electropositive group’ and ‘negative
elements’ or ‘radicals’ to refer to the tendency to attract or repel a pair of electrons (Lewis, 1916, p.
781-782). In his book Valence and the Structure of Atoms and Molecules from 1923 the term

6
It shall be noted, however, that some relative positions are very similar in any of these scales, like the
sequence of halogens (fluorine, chlorine, bromine, iodine), which were mentioned explicitly by Berzelius since 1931.
For a modern observer, this fact may suggest a connection to periodicity.

3
“electronegative” is used without any concrete definition (probably because it appeared common to
him already). Obviously, the use of that term was very common already around 1920. 7 In modern
terms the first quantitative definition of electronegativity is usually attributed to Pauling. 8 He wrote
"numbers representing their [elements] power of attraction for the electrons in a covalent bond, by
means of which the amount of partial ionic character of the bond may be estimated" (Pauling, 1950,
p. 236). In this definition, he clarifies that the concept is applicable only to covalent bonds and that
electronegativity is a property of elements related to the attraction of electrons in bonding
situations. In the initial article from 1932, Pauling built in his first scale positions for 19 elements.
Two years later, Mulliken (1934) called this property "electroaffinity" to relate electronegativity
with the atomic attraction power for electrons. Unlike the description of Pauling's electronegativity
for elements, Mulliken's approach as "the sum or average of ionization potential plus electron
affinity" (Mulliken, 1934, p. 783) is an explicit description for isolated atoms. It is worth
mentioning that although electronegativity is related to the concept of covalent bond, the
multiplicity of qualitative descriptions exceeds the description of molecular systems, and can be
made without using a notion of bonds. For example, Gordy defines “the electronegativity of a
neutral atom in a stable molecule as the potential at a distance r (covalent radius) from its nucleus
which is caused by the nuclear charge effective at that distance” (Gordy, 1946, p. 604). This
definition is the first to use a potential to define electronegativity for atoms in molecules. It shows
that the chemical bond in a molecule ceases to have the leading role in the description of
electronegativity: more and more it becomes an intrinsic property of atoms and is related to the
core/valence electrons relationship. These first characterizations define a first period in the
description of electronegativity in the so called semi-empirical methods in quantum chemistry. The
aim of this approach is to relate electronegativity with other physical properties or observables.
There are different ways to establish this relationship which will be examined in the next section in
which the empirical determinations and their difficulties are discussed.

With the development of ab initio methods in quantum chemistry9 in the decade of 1950
electronegativity enters into an even more complex theoretical field in which its qualitative
description is gradually slipping away from the unity it enjoyed before. For this approaches,
7
Even physicists like Max Born (1882-1970, Nobel prize 1954) used it. In an intriguing article about the bridge
between chemistry and physics (long before the famous Born-Oppenheimer approximation) the latter wrote:
”Regrettably, no method is known to measure electron affinities of the electronegative elements directly or to calculate
these from conceptions about the atomic structure.” (Born, 1920, p. 380, our emphasis)
8
Jensen (2012) claims that the first qualitative description of electronegativity was due to the American
physical chemist Worth H. Rodebush who in 1925 wrote: “If it might be permissible to introduce a qualitative formula
into science which is rapidly becoming exact, we might represent the electronegativity as a function of V/S where V is
the number of valence electrons and S the number of shells. The basis of this formula is Coulomb’s law and I believe
that in a few years we shall calculate the energy changes in chemical reactions by means of it”. (1925, p 383)
9
In 1950, p. 1561, Parr et al. introduced the term ab initio to name exact calculations in which the only
parameters allowed are those for fundamental constants. What fundamental constants are still is subject to discussions.
The ab initio quantum chemistry is also related to mathematical chemistry in actual research.

4
electronegativity seems to have lost its physical content. In a recent article, for example,
electronegativity is defined as “a physical quantity which can be calculated directly by quantum-
chemical methods (at least, in principle). ...is a formal parameter which has no physical sense but
can be used in correlation analysis of physical properties and reactivity of various classes of organic
and heteroelement-containing compounds” (Zueva et al., 2000, p. 613). But even if we accept that
the electronegativity is rather a “formal parameter” than a property, there is no clarity about how to
describe it properly. In the field of density functional theory (DFT 10) it is also possible to find a
wide range of non-equivalent descriptions. In the book Density-Functional Theory of Atoms and
Molecules (1989), Parr and Yang, for example, define electronegativity as a concept that can be
reduced to the chemical potential11: “the chemical potential concept is the same as the
electronegativity concept” (Parr and Yang, 1989, p. 74). Yet within the same approach we can find a
definition that differs to some extent from the formal description of Parr and Yang, for example
“electronegativity χ is a negative value of the chemical potential of electrons μ in an N-electron
system” (Urusov, 1994, p. 103) or as “as the minus eigen-energy of the unperturbed occupied
valence state involved in addition and release of electrons by atoms-in-molecules interactions”
(Putz, 2009, p. 733). DFT is not the only approach that claims to be an entirely theoretical
description of electronegativity. The definition of Allen for example is given as follows:
“Electronegativity Is the Average One-Electron Energy of the Valence-Shell Electrons in Ground-
State Free Atoms” (Allen, 1989, p. 9003).

3. Measuring electronegativity

Although the qualitative descriptions of electronegativity are a reasonable problem for philosophers
and scientists, this concept has not ceased to be tangible for any chemist. However if
electronegativity is a property is not one susceptible to direct measurement. For this reason bringing
together a qualitative electronegativity concept and a measurement domain has not been an easy
task. It is expected that as long as there is no clarity about what is electronegativity, attempts to
measure it will have the same degree of ambiguity and lack of unity. There are at least three
domains in which the measurement of electronegativity are preferred. The first is the
thermodynamic domain. Electronegativity is thought as a relational property of substances.

10
Density-functional theory is a recent approach in quantum chemistry about ground states. This method is
based on the electron density ρ(r) as the basic variable instead of the wave function ψ (r 1, s1, r2, s2,..., rn, sn). The density
ρ is just the three-dimensional single-particle density evidenced in diffraction experiments. The ground states are
described in terms of it.
11
In density-functional theory the electrons in an atom or molecule are treated analogically to statistical gas of
particles. In this approach the chemical potential measure the “escaping” tendency of an electron cloud. It is a
constant through all space for the ground state of and atom, molecule or solid.

5
Thermodynamic calculations of electronegativity are based on the idea to subtract from the
empirically measured bonding (or dissociation) energy all its covalent fraction in order to recover
what is considered the “ionic character” of that bond. According to this approach resonance12 the
chemical kind is the result of electronegativity differences between elements.

example Δχ
type of substance Bonding predictability
electron
of
pairs
chemical kind

homopolar chlorine, Cl2 (g) 0 yes yes


heteropolar hydrogen chloride, HCl (g) 0.9 yes yes
ionic sodium chloride, NaCl (s) 2.1 no yes
metallic sodium, Na (s) 0 no no

Tab. 1 Applicability of electronegativity differences

Pauling (1932) pioneered the use of such measurements based on the heats of dissociation or
formation of AnBm type substances, where n and m are the stoichiometric ratio of the compound.
The difference between the actual and the purely covalent bond between A and B is a measure of
the partial ionic energy (Δ) in the AB bond:

Δ = EAB – ½ [EAA + EBB] (1)

This partial ionic energy is related to the different way in which bonding electrons are shared in
substances A and B and is a measure of electronegativity (χ).

∆A:B = (χA – χB)2 (2)

Thus, this formal description gives only differences of electronegativity. To obtain the absolute
values it was necessary to postulate hydrogen as – feasible – central element. For this role, Pauling

12
Form the point of view of quantum mechanics one system can be described by a wave function represented as
a superposition of all it's possible states. This concept is interpreted in the valence bond method in quantum chemistry,
associating to every term in the wave function a Lewis structure that describes the molecule. For example, in a normal
covalent bond between atoms AB described by the wave function: ΨAB=aΨA:B + bΨA+B- there is at least two structures
A:B and A+B-. The normal state of AB is described using both structures, it's said that AB is a hybrid of resonance.

6
assigned it with χ = 0 (later with χ = 2.2, to avoid negative χ for most metals). In 1961 Allred13
updated the Pauling scale incorporating more accurate thermodynamic measurements. Despite the
widespread use of the Pauling scale and its subsequent improvement by Allred a major difficulty in
developing a thermodynamic scale of electronegativities is the absence of measured values for
many single bonds and especially for metallic bonds. However, the thermodynamic scale is one of
the most used and generally serves as a point of reference for other electronegativity scales.
Finemann (1958, p. 948) has extended the range of applicability of this scale giving values for
radical using the relationship:

ΔAR = EAR – ½ [EAA + ERR] (3)

With this relation it becomes clear that electronegativity varies with bond multiplicity in radicals.
Table 2 summarizes average values for some radicals:

R χ R χ R χ R χ
CH3 2.6 NH2 3.1 BH2 1.9 [HCO3] 3.4
CF3 2.9 NF2 3.2 PH2 2.3 [HPO4] 3.4
SiF3 2.0 NCS 3.2 SiCH3 1.9 [NO3] 3.7
CHCH2 2.7 NNN 3.3 OCH3 3.4 [SO4] 3.7
CCH 2.8 NC 3.3 OC6H5 3.5 O2 3.5
CHO 2.9 NO2 3.4 OH 3.5

Tab. 2 Average thermo-chemical electronegativities χ of radicals R (Batsanov and Batsanov, 2012,


p. 133)

The second measurement domain that has been used to build a scale of electronegativity is the
spectroscopic. In turn, this domain can be divided into two main fields, one that uses the atomic
radius (r) and the effective nuclear charge (Z eff) as a geometric relationship between the valence
electrons and the nuclei, and one that uses the ionization potential and the electron affinity. The first
subgroup of spectroscopic measurements was developed on the basis of calculations of the bond
length made by Schomaker and Stevenson (1941) as an application of thermodynamic values of
Pauling's electronegativity:

rAB = rA + rB – 0.09∆χAB (4)


13
Allred worked in electronegativity measurements in the thermo-chemical domain (1961) using actual data on
single bonds absent in Pauling's former work and in the spectroscopic domain in the joined work with Rochow
(1958).

7
This relation made possible to relate covalent or ionic radius to electronegativity as an atomic
related property. The Scales of Gordy (1946) and Allred and Rochow (1958) are based on this basic
assumption in which the atomic electronegativity is calculated as:

χ = a Zeff / rn + b (5)

a and b are the linear coefficients that relate this electronegativity values with those of thermo-
chemical measurements. This approach is also called the geometrical electronegativity concept
because it relates the geometrical disposition between the nuclei and the valence electrons. We
prefer to refer to it as spectroscopic measurement because it contains some values for r and Zeff from
empirical determinations.

The differences between the approaches of Allred and Rochow on the one hand and Gordy on the
other are based on how to obtain values for Z eff. According to Gordy the value of the effective
charge of the nucleus can be obtained by the difference between the valence electrons (n) and the
shielding constant of the valence shell (s):

Zeff ≈ n – s (n – 1) (6)

Furthermore, Allred and Rochow (1958) built their value for n in eq. 5 using values obtained
empirically with X-ray spectroscopy or using Slater's rules14. These two scales were the first to
introduce the concept of force for successful measurements of electronegativities and allowed to
develop electronegativity values for elements that have not being measured.

Gordy has also developed a useful relation to measure electronegativities for metals relating
equation 5 with the bond-stretching force constant k measured in radio frequency spectroscopy, the
bond length d and the bond order N:

k = aN (χAχB / d2)3/4 + b (7)

Another method utilizing the covalent radius to develop an electronegativity scale was developed
by Sanderson (1951). According to this author electronegativity is associated with the average
electron density of the atoms (ED). The latter is representative of the average degree of
compactness of the electronic sphere around the atomic nucleus. This number is a measure
associated with each atom and can be calculated using the number of electrons (Z) and within either
covalent or ionic radius r as follows:

ED = Z / 4.19 r3 (8)

14
According to this set of rules, the electron shells have their own values of n1 the sum of which determines the
shielding constant of the atoms n. Within this scale the value of n in eq. 5 is 2 while that in Gordy's scale is 1

8
The electronegativity of an atom can be calculated through a stability radius, obtained as the ratio
ED/EDi, where EDi is the electron density value of an inert isoelectronic atom. In order to calculate
EDi it is necessary to obtain a covalent radius that is not accessible empirically and which is
interpolated from known values, therefore it is a fictional value. Sanderson claims that the stability
radius is a measure of the degree of compactness of electrons and a measure of the force of
attraction to the nucleus, thus a measure of electronegativity. Following this approach, Sanderson is
one of the first to obtain values for the noble gases. Table 3 summarizes different ways to relate Zeff
and r to obtain values for electronegativity:

Year Authors Equation Notes


1942 Liu χ=a(N* + b)r2/3 N* is the number of e-shells
1946 Gordy χ=a(n + b)/r + c N is the number of electrons
1951 Cottrell, Sutton χ=a(Z*/r)1/2 + b dimensionality of E1/2
1952 Sanderson χ=a(N/r3) + b N = Σe
1956 Williams χ=a(n/r)b n is the number of valence electrons
1958 Allred, Rochow χ=a(Z* - b) r2 + c Z* of Slater

Tab. 3 Pioneering works of the development of the electronegativity concept using Z and r (cf.
Batsanov and Batsanov, 2012, p 134).

The second subgroup of spectroscopic measurements of electronegativity was developed by the


American chemist Robert Mulliken (1896-1986). Within this approach electronegativity is
measured as the average of the sum of the electron affinity A and ionization potential I:

χA = ½ (IA + AA) (9)

This relation allows to obtain values for isolated atoms at any valency state. However, the lack of
empirical measurements for the ionization potential and electron affinity have been a large obstacle
to the construction of a complete scale (in his 1934 article, Mulliken gives values of just 11
elements). In contrast to the lack of empirical data this approach was considered to be of reasonable
value: "Mulliken's method has much greater theoretical support" (Mullay, 1987, p. 4). Intuitively,
this approach has been described as the tension of one atom between the tendency to lose or gain
electrons. This tendency Mulliken calls "electroaffinity", and assumes that both I and A can be a
correct measure of that property. As the ionization potential can be measured and calculated for

9
different valence states, Hinze and Jaffe (1962) developed the concept of orbital electronegativity to
denote the orbital energy used for an atom in a chemical combination, the description associated
with an orbital allows I to establish different electronegativities in valence variations for the same
atom. This orbital electronegativity can be obtained as the partial derivative of the energy of the
atomic orbital j with respect to electrons occupying n:

χA,j = ∂EA / ∂nj (10)

Allen (1989) has developed another scale of electronegativity using the ionization potential for
Allen electronegativity should be considered the third dimension of the periodic table because “It is
most likely that this new third dimension is an energy because the Schrödinger equation itself
identifies energy as the central parameter for describing the structure of matter” (Allen, 1989, p.
9003). Allen's argument is that the periodic table itself is reducible to the foundations of quantum
mechanics, and proof of this is the periodicity that the electronegativity presents. For Allen
electronegativity is obtained as:

χspec = mεp + mεs / m + n (11)

M and n are the number of valence electrons of p and s orbitals; mεp and mεs are the one-electron
energies obtained as the difference between the ground state and the first ionized state. The third
measurement domain for electronegativity is the electrostatic which actually was following the
thermochemical approach very early.

Because the heats of dissociation (or of formation) of two-atomic compounds are used to set it up
the Pauling scale of electronegativity initially is a relative (or relational) model. In other words, to
obtain values for single elements (or atoms) makes necessary a kind of calibration or
standardization. Accordingly, the Pauling scale is generally restricted to those elements which form
two-atomic compounds, and it excludes the noble gases (no compounds of the noble gases were
known at the time Pauling was publishing his original work) 15. Electronegativities could be assigned
to noble gases by so-called “absolute” concepts that is concepts which refer to atomic observables
and “properties” rather than to relative or relational – for example thermochemical – means.

While noble gases had to wait quite long to get on any electronegativity scale, for the metals the
story began much earlier with the discussion of the dipole moment as a calculational reference, and
thus supplemented the original data published by Pauling, who – as we have mentioned already – in
his original work did refer to covalent compounds only (Pauling 1932). As reference to discuss

15
Intriguingly, Pauling himself (1933, p. 1899) predicted the possible existence of KrF 6 and XeF6, noble gas
compounds which were synthesized successfully only in the 1960s by the British-American chemist Neil Bartlett (1932-
2008).

10
electronegativity, the American chemist J. Gilbert Malone – who, as it were, was the very first to
react on Pauling´s concept – chose the dipole moment (electric moment) of two-atomic and more-
atomic molecular substances in a short article on The Electric Moment as a Measure of the Ionic
Nature of Covalent Bonds (Malone 1933).

Dipole moments are inferred from the measurement of dielectric constants (permittivity) of
substances. Malone is referring to such values from the published literature and aims at another
measure (or explanation) for electronegativity: “The perfection of two independent measures of
electronegativity, one based on thermal and one on dipole moment data, would be of considerable
value.” (Malone, 1933, p. 199) According to this value, Malone mentions two aspects, the
supplementary aspect of differently incomplete scales and the perspective to calculate “bond angles
in polyatomic atoms”. However, “Before this can be done a clearer definition of electronegativity
must be given.” (Malone, 1933, p. 199) Unfortunately, he is not elaborating on what he means by a
“clearer definition” in that article. On the contrary, he seems to follow the Pauling concept without
any reasonable criticism. He even revised his first approach (which he began to develop before
Pauling´s paper appeared) when he recognized that Pauling has used hydrogen as the reference for
his scale. What might be considered a critical statement can be found in the very first paragraph of
Malone´s paper: “The idea that the electric moment bears some relation to the ionic quality of the
two-electron type of bond has been known for a number of years. It is true that the terms used in
describing a bond have been loosely applied, and it is uncertain in some cases just what meaning
has been intended.” (Malone, 1933, p. 197)

Element Pauling Mulliken/Jaffe Allred/Rochow Gordy Sanderson Allen


H 2.2 2.1 (s) 2.20 2.17 2.31 2.3
Li 0.98 0.84 (s) 0.97 0.96 0.86 0.912
Be 1.57 1.40 (sp) 1.47 1.38 1.61 1.576
B 2.04 1.93 (sp2) 2.01 1.91 1.88 2.051
C 2.55 2.48 (sp3) 2.50 2.52 2.47 2.544
N 3.04 2.28 (p) 3.07 3.01 2.93 3.066
O 3.44 3.04 (p) 3.50 3.47 3.46 3.610
F 3.98 3.90 (p) 4.10 3.94 3.92 4.193
Na 0.93 0.74 (s) 1.01 0.90 0.85 0.869
Si 1.90 2.25 (sp3) 1.74 1.82 1.74 1.916
Cl 3.16 2.95 (p) 2.83 3.00 3.28 2.869

11
Ge 2.01 2.50 (sp3) 2.02 1.77 2.31 1.994
Br 2.96 2.62 (p) 2.74 2.68 2.96 2.685

Tab. 4 Selected electronegativity values from different scales (Mullay 1987). Allen values are
quoted from Allen (1989).

Importantly, although the values shown in this table may vary in different degrees, the orders of the
elements in the different electronegativity scales is the same, but the way in which this values has
been obtained is completely different in the experimental measurements, and in cases where it is not
measured directly calculated values also tend to converge.

4. The meaning of electronegativity – discussion

There are various published reviews about electronegativity from the point of view of theoretical
chemistry (e.g., Pritchard and Skinner, 1955 Ferreira, 1967; Batsanov, 1968; Mullay, 1987) but
only rarely more general issues (e.g., Jensen 1996, 2003, 2012) nor questions referring to the
philosophy of chemistry have been raised. On the basis of the brief systematical remarks on the
recent, mainly post-quantum developments in the former paragraph, we shall discuss four crucial
issues or questions in order to clarify the meaning of electronegativity: persistence, reference,
status, and experimentation. We admit that there might be some overlap between those issues and
the suggested interpretation.

(1) What is the support for the unification of the concept which is assumed in the community of the
chemical sciences, despite of the many different qualitative and quantitative attempts to give
meaning to the statement a is more electronegative than b?

The persistence or stability of the electronegativity concept in chemistry with respect to its
applications in education and science, the wide-spread acceptation as a reasonable and unified
theoretical concept, and the on-going interest in theoretical chemistry is something very surprising.
As we have pointed out, there is no consistent and uniform qualitative or quantitative definition or
description up to the present day. On the contrary, with the development of both semi-empirical and
so-called ab initio quantum chemistry we observe a proliferation of a variety of electronegativities.
Nevertheless, the results of this variety of approaches tend to converge, or in other words, seem to
represent something very similar. However, the idea about this something rather emerges from
intuition than from clear, logically comprehensible conceptions. Anyhow, it is possible to build

12
full-scales of electronegativities for all chemical elements – containing even those which are not
active chemically – and any group of atoms using different measurement domains. All these
different physical descriptions, as covalent radius, ionization potential, electron affinity, dipole
moment, dissociation energy, and spectroscopical data can be interpreted as if having something in
common. As to the interpretation of this puzzling situation two aspects should be taken into
account. Firstly, aside from Pauling´s (dissociation energy) and Malone`s (dipole moment)
approach, the measurements referring to the mentioned physical descriptions explicitly aim at
atoms. Because of this strong “simplification” – which is of course typical for the scientific method
in physics – the obtained data address processes very close to each other. Adjoining or removing of
outer shell electrons is the basis for ionizations, and the same electrons do have a crucial impact
with respect to spectroscopy and atomic radii measurements. Thermochemical descriptions
indirectly refer to more simplified systems, too: Dissociation of a diatomic structures leads to
atoms, and during that process – using the modern view – again the outer electrons are involved
(during the cleavage of electron pairs). Slightly more complicated is the situation with respect to
the use of dipole moments as a starting point, as already the work of Malone 1933 has shown. In
order to get to Δ-measures from dipole moments of hydrides, he had to assume molecule bond
angles.16 From that point of view, it is not really surprising that there is similarity or convergence
between the different scales. That leads to a second important aspect which refers to periodicity.
We know various means to reveal systematic and periodic trends of chemical elements when
aligned according to the structure of the periodic system, particularly when main block elements
are taken into account.17 Among those are atomic radii, ionization potentials, and electrical
conductivity. Hence, to represent the trends of elemental behavior which are expressed in
electronegativity scales, it is not necessary to use this ill-defined concept. To put it in other words:
Periodicity can be described using direct measurements rather than electronegativity.18

Accordingly, a proper explanation of “a is more electronegative than b” can only be given if the
carrier of that property is well-defined and if a concept of periodicity is taken into account, too.

(2) What is the target, reference, or carrier of electronegativity, in other words: what kind of entities
are a and b?

Many chemists consider electronegativity to be a molecular concept:

16
The dipole moment is a resultant of the electrical impact of the constituents. There are heteropolar substances
with a dipole moment of zero (e.g., CCl 4), and homopolar substances with a dipole moment different from zero (e.g.,
O3).
17
The atomic number is of course not periodic. Intriguingly, electron affinities do not show a through-going
periodic correlation.
18
Mark Leach is claiming a strong role for electronegativity with respect to the explanation of chemical
periodicity (Leach, 2012). We will be discussing this issue and the applicability of a Kantian point of view in more
detail elsewhere.

13
“Electronegativity is an important part of the intuitive approach to understanding nature that sets
chemists off from other physical scientists…The reason for the sustained interest appears to lie in
the fact that the idea of electronegativity is practically a direct consequence of foundation concepts
of modern chemistry, specifically the following three:

(1) molecules are made up of atoms held together by chemical bonds.

(2) chemical bonds involve a sharing of electrons between the atoms.

(3) The electrons are not always shared equally.” (Mullay, 1987, p. 2)

Although the first sentence appears to say something substantial (“intuitive approach”, see next
point below), the claim that the molecular picture is central for the “idea of electronegativity” is
incorrect historically as well as systematically. As to illustrate the systematical incorrectness, we
might take a closer look at the electronegativity of metals. Because metals under normal
circumstances do not form covalent bonds (or molecules), Pauling´s theoretical approach of the
extra energy in an atomic bond – we might call it an explanation – cannot be applied to them (see
eq. 1). In the “ionic” state, be it a salt or a metal, the model of a bonding electron pair makes no
sense. Consequently, the first scale of Pauling did not contain metals. If a suitable calculation
method is found, however, the empirical approach of Pauling can very well be accommodated
because the measurement of thermochemical data is independent of any atomistic interpretation.
Hence, we find electronegativity values for metals in “absolute” (like Mulliken´s) and relative
scales (like Pauling´s), too. However, there is one restriction to the common application of the
differences of electronegativities: These differences are of heuristic value for the prediction of
bonding kinds or chemical kinds – except metallic states (see Tab. 1). If the differences are
reasonably high (like for sodium fluoride, NaF), the predicted chemical kind is ionic (and the
substance occurs in the solid form), if the differences are low or even zero (like in dioxygen, O 2),
the predicted kind of bonding is molecular (and the state of aggregate at standard conditions is
gaseous) – if and only if two non-metallic elements are compared. This very useful rough
estimation is not applicable to pairs of metal atoms, independently of the chosen approach or scale.
We have to put in some chemical knowledge in order not to jump to conclusions like: If metal
atoms are brought together chemically, the result is something molecular which must be gaseous or
liquid (because the difference of electronegativities is zero). In addition, there is another problem:
“It is extremely convenient in the teaching of elementary chemistry to divide the elements into
metals and nonmetals … However, a study of the electronegativity values as determined from
electronic densities shows that some elements whose physical properties identify them beyond
question as metals are more electronegative than some elements whose physical properties identify

14
them as nonmetals.” (Sanderson, 1952, p. 543) Though this statement is not easy to be verified with
respect to the common scales, it gives an indirect hint to some similar taxonomic concepts in
chemistry. The concepts of states of aggregate, acidity, and of noble/ignoble metals (later:
electromotive series, cf. Fig. 1), for example, were forged in order to classify chemical substances,
first in a qualitative manner, later more and more quantitatively. From this point of view,
electronegativity is similar to those concepts.

Considering the indicated line of interpretation, electronegativity appears to be referring to atoms,


in particular to the outer shell electrons of those atoms. The recent historical developments seem to
point out that the more reduced and the more abstract the approach, the more “pure” is the access to
electronegativity (see, for example, the concept of orbital electronegativity, eq. 10, and Allen´s
quantum-chemical approach, eq. 11). If we take a superficial look at the obtained measures (see
table 4), it indeed looks as if absolute electronegativity values really exist. However, a closer look
reveals that this is not correct, because all mentioned physical descriptions or measurements
actually do compare (at least) two states. This is obvious regarding Pauling´s calculations which
refer to a substance and its dissociated parts; this is not so obvious with respect to the atomic
physical measures, but here as well two states are compared, for example, the uncharged particle
and the anion, if the electron affinity is to be investigated. Even less obvious is this relationality if a
concept like electron density is applied and covalent radii are needed for calculations (cf.
Sanderson, 1951, and Table 3).

Χ (ionization Χ Electron
Element potential) (Mulliken) affinity

Eea / kJ/mol

He 5.65 - - 21
Ne 4.95 4.60 - 29
Ar 3.62 3.36 - 35
Kr 3.21 2.98 - 39
Xe 2.79 2.59 - 41

Tab. 5 Noble gas electronegativity values (Χ) as inferred from ionization potentials alone
(reference: ΧFluorine {Pauling} = 4) and from Mulliken (cf. eq. 9), together with their electron
affinities (Eea). Note that the latter have to be read conversely to the common convention: negative
values mean endothermic processes (values from Atkins and de Paula, 2013).

15
As to this case, atoms and their electronic shells are separated conceptually from their environment,
that is, from their chemical vicinity and all pertinent relations. As a result, it becomes – through eq.
8, for example – possible to calculate electronegativities even for noble gases, which might be
considered an advantageous achievement from the side of reductionist theoreticians, but it could as
well be regarded as a theoretical weakness if the core of chemistry is taken seriously. The lower
noble gases do not take part in the chemical life, they do not form compounds. As to those
elements, even simplified calculations of “electronegativities” (like those inferred from ionization
energies, see Tab. 5) yield reasonable results in comparison with more sophisticated approaches.
However, the electron affinity measurements indicate that noble gas atoms do show just the
opposite behavior than that expected for entities of that elevated electronegativity because they do
not tend at all to take up additional (bonding) electrons. Thus, electronegativity as the phenomenon
to attract or repulse electrons in bonding states is neglected by pure atomic-physical approaches, or,
to put it differently, this approach is only pretending to say something typically chemical, but in
fact it only reformulates well-known atomic-physical facts (as does the trend of decreasing values
with increasing atomic numbers in a main-element group) without chemical relevance.

With the development of “absolute” values for electronegativities the concept is mainly applied to
atoms, and, with respect to the surveyed theoretical developments, the former relational character is
lost. In order to come to terms with those two faces of electronegativity – we might call them the
“chemical” and the “physical” face – different possible ways can be chosen with respect to a
philosophical interpretation. Firstly, one might take the physicalistic stance, according to which the
belief is prevailing that everything can be explained by referring to the energetical properties of
elementary particles and electromagnetic fields. According to electronegativity, this stance may
lead to an exaggeration of the applicability of theoretical reduction, en el cual electronegatividad
refiere a un item en el dominio ontologico de la fisica de particulas que puede ser propiamente
descrito por un observable. O de lo contrario admitir que la electronegativividad no puede ser
descrita enterminos mecanico cuantivos, en cuyo caso hay que admitir its removal from chemistry.
However, there are – secondly – more neutral ways to think the interdiscourse relation. The Dutch
philosopher Jaap van Brakel, for example, tends to avoid the classical relational descriptions like
reduction, supervenience, and emergence, because according to his investigations – particularly in
the philosophy of chemistry – they cannot be applied convincingly (van Brakel, 2000, pp. 50-57).
Together with the neglect of these interdiscourse relational descriptions, he suggests an
intercultural approach which does not subscribe to the widely believed opinion that the scientific –
that is, physical – description is the only true and really fundamental one. Instead of assuming –

16
more or less explicitly – that it is enough to say something formally correct about isolated atoms,
electrons, orbitals, or the like, he prefers to consider different approaches like macroscopic,
microscopic, and submicroscopic views to explain or represent the behavior of entities to be at eye
level, and to analyze the relations between them carefully and in detail. 19 If we follow this non-
reductionistic suggestion, the intended chemical core of electronegativity comes to the fore because
the more substantial aspects and the relationality can be emphasized without essentializing
preconceptions. So far, we can conclude that the carriers of modern electronegativity are atoms in
their interaction with the environment. There are approaches which address this interaction
explicitly, and those approaches might not even need any reference to atoms (Pauling). There are
other approaches which address relationality only implicitly (Mulliken), and approaches which do
not address relationality at all (Allen).

Accordingly, the role of the particular chemical bond in the description of electronegativity should
not be underestimated: Different orbitals – atomic or molecular – have different values of
electronegativity, and the same class of atoms with variations in hybridization (multiple bond
states), too.

(3) Which epistemological status is to ascribe to electronegativity? Is it a classificatory concept, a


property, a disposition; does it embody a theory, or a chemical law?

As we have seen, electronegativity certainly has a classificatory or taxonomic character to it. Some
elements (or their atoms) are electronegative, others electropositive. There are trends we can easily
follow by using the periodic table background: Along the horizontals, electronegativity is
increasing, along the verticals, it is decreasing. Hence, it is coherent to connect it with what might
be called periodicity (although it is open, how)20. As we have also seen already , however, this
classification can be – at least in some cases and in approximation (s. Tab. 5) – inferred from other
data, too. In contrast to electronegativity though, these other data might be directly measurable
(ionization energy) or more easily derivable (atomic radii). The more relationality is introduced by
the applied empirical background – or projected theoretically – into the concept, the more it
becomes a chemical property. Descriptions of isolated particles can only be of complementary
service, that is, these descriptions cannot be used as a basis to build up a proper electronegativity
scale.

Only shortly we shall refer to the usefulness of the notion of disposition21 here. This notion is

19
For a brief account of this philosophical point of view, see Ruthenberg and Harré, 2012.
20
See footnote 18 (Leach).
21
In the discussion about the unifying aspects underlying chemical concepts as Acidity, Chang (2012) has leaved
open the same path to discuss the propensity of acids in Lewis' approach. If acids and bases are defined as donor
acceptor of electron pairs, electronegativity becomes a measure of that propensity. It is interesting that the results of
Chang with respect of this matter are very close to the conclusions presented in this article and the actual ambiguity

17
intensely discussed particularly in contemporary analytical philosophy (cf. Vetter and Schmid,
2014). With respect to the natural sciences, particularly to chemistry, dispositional descriptions
become important because characterizations of substances usually contain claims like dissolvable,
explosive, inflammable, acidic, analgetic, reductive, liquefiable and so forth, which are considered
to be dispositions by many philosophers involved in this discussions. Dispositions of that kind
seem to be better applicable to our subject than a blunt assignment of a static property at least with
respect to one aspect: There is not one carrier which simply has this property (or ability, or
capability), rather we are talking about a kind of response function which is very close
conceptually to what chemists address as reactivity. Whether or not the chemical realm as a whole
should or must be interpreted as consisting of a web of such dispositions (dispositional monism)
shall be left open at the moment.

As to the ambiguous status of electronegativity some scientists deny the claim that it is a proper
theoretical achievement: “Interpretative chemistry still operates in terms of empirical concepts such
as electronegativity and molecular shape, commonly said to be without theoretical underpinning.
Admittedly, there are those who pontificate against the use of such classical concepts, while
remaining singularly unsuccessful themselves to account for the predictive powers of these ideas,
or to provide alternatives that are theoretically more soundly based” (Boeyens, 2003, p. vi).
Consistingly, the ab initio effort tries to eliminate or replace electronegativity by other – allegedly
more appropriate – concepts such as the chemical potential. From that point of view a defense of
the physical meaning of the term electronegativity would require a set of statements in the form of
laws that explain and predict the behavior of (even sub-microscopic) chemical systems. As a result
of the impossibility to accomplish this deduction, electronegativity for reductionists cannot be a
real property: “Of all of the many concepts developed in the past to provide a theoretical
foundation for chemistry, electronegativity is one of the more enduring. Purportedly introduced two
centuries ago by Berzelius, it still features in many theories, albeit like the concepts valence, bond
and structure, without ‘first principle underpinning’. This means that they are not represented by
quantum-mechanical observables, have served a noble purpose in the past and are now obsolete”
(Boeyens and Du Toit, 1997, p. 296). Referring to the line of argument presented here, this
statement is questionable both historically and epistemologically. As we have seen already, non-
reductionist approaches do indeed have very promising prospects, and “chemical theories” do not
necessarily need any specific and particular physical certification to acquire scientific and
philosophical legitimacy22.

of electronegativity but Chang doesn't even mention it at all: “It [Lewis acidity] would have to be something like
the propensity to accept electron pairs, but there would be several different theoretical ways of making that notion
properly quantitative, not to mention linking those ways to performable measurement operations.” (p. 696)
22
With respect to the non reducibility of the “chemical world” to the (supposed) more fundamental laws of

18
In good harmony with a non-reductionist stance is scientific perspectivism. According to the
approach of Ronald Giere (2006), scientific descriptions are all perspectival, that is, dependent on
the context, the agent and the purpose. Scientific representations are four-place relation of the form:
S uses X to represent W for purposes P. The difficulty with this proposal is that it presupposes the
existence of a single item W to be represented from different perspectives. As to our subject, it is
more than questionable whether we are allowed to refer to isolated atoms as manipulated real
entities. Rather, electronegativity is a concept that describes processual behavior or dispositions
than properties of single individuals in chemical systems. If we were to represent a W of
electronegativity it would be more complex than that isolated atom could represent and, moreover,
for many of the cases referred to in the present study the pertinent characterizations are
incompatible because the chosen experimental frameworks and models are entirely different.
Obviously, it is not the theoretical identity which supports the assumed unity or convergence of the
concept: Even the common approaches of quantum chemistry point at plurality because there is no
single way to achieve solutions of quantum laws for isolated molecules, like Valence Bond and
Molecular Orbital methods, Quantum Theory of Atoms and Molecules, and Density-Functional
Theory. It is not expected that quantum chemistry can be unified into a single theory.

In another attempt to find a suitable interpretation, we turn another time from submicroscopic to
macroscopic if not manifest descriptions in (applied) science, and shall briefly consider
dimensionless numbers for comparison. Very well introduced in chemistry is the pH value for the
characterization of acidity or basicity in aqueous solutions. It is dimensionless because it is the
logarithm of the molar concentration of hydrogen ions (times minus one). Hence, this number is an
“observable in disguise”, and the reset into the measured value and the retranslation into the
corresponding meaning are easy. The derivation of electronegativity as a dimensionless number is
not that clear-cut and very diverse (see eqs. 1-11). Clearly, pH and electronegativity serve different
functions. However, there indeed are other dimensionless numbers which have much more
functional similarity with the latter: numbers which are used in chemical engineering to describe
characteristics of fluid flow, heat transfer, and mass transfer, like the Reynolds number, the Nusselt
number, and the Prandtl number. The Reynolds number, for example, is applied to differentiate
between laminar and turbulent flow of fluids:

Re = ρ d ν / η (12)

In eq. 12, ρ is the density of a fluid, ν its velocity, d the flown through length, and η the dynamic

quantum mechanics, Lombardi and Labarca (2005) developed an argument for the use of a Kantian rooted ontological
pluralism based on the supposed conceptual rupture between the interpretations of the orbital concept in chemistry and
physics (Lombardi and Labarca, 2006). Cf. three responses to the arguments of Labarca and Lombardi can be found in:
Needham (2006), Córdoba and Martínez (2014) and Hettema (2012, chapter 10 section 10.3.3)

19
viscosity. The Reynolds number is a classification number and a characteristic ratio at the same
time. In the original paper, Osborne Reynolds states “(...) that the general character of the motion of
fluids in contact with solid surfaces depends on the relation between a physical constant of the fluid
and the product of the linear dimensions of the space occupied by the fluid and the velocity.”
(Reynolds, 1883, p. 935; our emphases). The physical constant Reynolds refers to is not specified in
much detail here. If Re is less than circa 2100, the fluid flow is laminar, and above that quantity, the
flow turns to become turbulent. By no strict theoretical means it can be explained exactly why and
when the flow of a liquid is in the one or the other state because laminarity and turbulence are
macroscopic descriptions. Different than pH and χ, Re is read quasi binary, that is, it helps to
differentiate between precisely two opposite states. Another difference is even more relevant in the
present discussion: As we have emphasized already, electronegativity has no manifest or direct
reference, rather this reference is speculative and abstract. The basis of the pH is a direct
measurement, that of the Reynolds number the correlation of measurements and manifest
observations of the behaviour of moving fluids. Hence, electronegativity is a dimensionless number
without clear reference, and because of this the pertinent field of its modelling is very diverse.
However, the attempted reductionistic or materialistic modelling does not look very promising. In
his book Philosophy of Chemistry, Jaap van Brakel discusses the role and function of models like
Re in chemical engineering in detail. He says:“What we have is a world of interrelated models,
where no matter which model or description one picks out and tries to say what it is that is being
modelled, what is being modelled is itself a model of something else.” (van Brakel, 2000, p. 189)23

(4) Which is the relevance of experimentation and measurement?

We have emphasized that some approaches to infer electronegativity from measurements are based
more macroscopically, whereas others explicitly refer to an abstracted level. In the framework of
the classical distinction between empirical and theoretical representations, a clear sequence –
systematically and historically – can be set up which starts with the early, more or less pre-quantum
attempts of Pauling and Malone, has a middle section with for example Mulliken and Allred, and a
final, modern end occupied by the likes of Sanderson and Allen.

According to the functional taxonomy of instruments and experiments proposed by the German
philosopher Michael Heidelberger, there are two basic types of experiments, those to improve or
expand knowledge and those to adjust actual knowledge to a theoretical context (Heidelberger,
2003). The first type can be differentiated into a productive, a representative, and a constructive or
imitative function. Productive is an experimental setting that produces phenomena which usually
could not be observed, representative we call instrumental techniques that transform one

23
. The author does not address electronegativity in that context.

20
phenomenon into another, and imitative experiments are built up to present the phenomenon in a
“pure” form in order to mimic nature. Although we are talking about the creative interpretation of
measures (like dissociation energies, ionization potentials, dipole moments, and electron affinities)
rather than a real measurement of a property, this taxonomy might be of heuristic service: We could
consider the electronegativities as mainly productive because they are produced through an artificial
derivation or construction from measured observables. Just as well the representative point of view
might be emphasized in our case: The transformation of ionization energies into a dimensionless
number, for example (see Tab. 5), is something like a transformation of an experimentally forced
phenomenon into a property-like or disposition-like description or model. One main conclusion
Heidelberger draws from his investigations in the history and philosophy of experiment is that the
nature of experimental attempts lies not solely in the testing of theories (like earlier philosophies of
science – verificationism and falsificationism – have suggested) but in the making of reality.

The assumed convergence of the different approaches is much weaker than we do observe in other
common measurement procedures like, for example, the Avogadro constant or analytical
determinations of chemical species. If we are to determine the concentration of, for example,
mercury in a certain matrix, we might apply different techniques like, to mention few opportunities,
X-ray-fluorescence spectrometry, redox titration, atomic absorption spectrometry, photometry,
atomic emission spectrometry, and mass spectrometry. If the sample pretreatment is performed
appropriately and accurately, the results may well converge (although in practice they rarely do). In
this example, the target has a well-defined material background, namely the number of mercury
atoms per volume. Contrasting that, for electronegativity the convergence is only a calculated one
because the target is only intuitively characterized. Hence, in the case of electronegativity the
involved experiments are certainly not applied just to test hypotheses (or theoretical predictions).
Rather, a number of extremely diverse measures are considered to describe – or, to put it more
provocative: to make – an assumed property.

6. Conclusion: What is electronegativity?

- “its modern history spans about 50 years. But even to this day there is no definite answer to the
question, what is electronegativity? Why electronegativity is so useful to chemist? And why has it
had such a long existence?” (Mullay, 1987 p. 2)

- The more “stuffy” the method is, the more complicated the derivation.

- The more “quantum” the description, the more remote are stuff properties.

21
- tendency towards abstraction, reduction, atomism, QM (change in definition?)

- description of isolated atoms leads to nowhere (cf. noble gases)

- The more relationality is projected into the concept, the more it becomes a chemical property.

- “Electronegativity is an important part of the intuitive approach to understanding nature that sets
chemists off from other physical scientists (Mullay)

- no need/reason to essentialize the one and only (real) version

- EN is a dimensionless number with a complicated referential background

- convergence of different approaches not necessarily leads to an existence proof

8. References

8. References

Allen, L. C. (1989). Electronegativity Is the Average One-Electron Energy of the Valence-Shell


Electrons in Ground-State Free Atoms. Journal of the American Chemical Society, 111,
9003–9014.

Allred, A. L. (1961). Electronegativity Values from Thermochemical Data. Journal of Inorganic


Nuclear Chemistry ,17, 215–221.

Allred, A. L. & Rochow, E. G. (1958). A Scale of Electronegativity Based on Electrostatic Force.


Journal of Inorganic and Nuclear Chemistry, 5, 264–268.

Atkins, P. and de Paula, J. (2013). Elements of Physical Chemistry. Oxford: Oxford University
Press.

Batsanov, S. S. (1968). The Concept of Electronegativity. Conclusions and Prospects. Russian


Chemical Reviews, 37, 332–351.

Batsanov, S. & Batsanov, A. (2012). Introduction to Structural Chemistry. New York: Springer.

Berzelius, J. J. (1825). Lehrbuch der Chemie, (trans. by F. Wöhler). Ersten Bandes erste Abtheilung.
Dresden: Arnoldische Buchhandlung.

Berzelius, J. J. (1820). Versuch über die Theorie der chemischen Proportionen und über die
chemischen Wirkungen der Electricität; nebst Tabellen über die Atomgewichte der meisten
unorganischen Stoffe und deren Zusammensetzungen, (Transl. K. A. Blöde). Dresden:
Arnoldische Buchhandlung.

Boeyens, J. C. (2003). The Theories of Chemistry. Amsterdam: Elsevier Science and Technology
Books.

Boeyens, J. C. & du Toit, J. (1997). The Theoretical Basis of Electronegativity. Electronic Journal
of Theoretical Chemistry, 2, 296–301.

22
Born, M. (1920) Die Brücke zwischen Chemie und Physik. Die Naturwissenschaften, 8, 373-382.

Chang, H. (2012) Acidity: The persistence of Everyday in the Scientific. Philosophy of Science, 79:
690 – 700.

Córdoba, M., Martínez, J.C. (2014) Los orbitales cuánticos y la autonomía del mundo químico.
Theoria, 80, 261-282.

Farber, E. (1952). The Evolution of Chemistry. New York: The Ronald Press Company.

Ferreira, R. (1967). Electronegativity and Chemical Bonding. Advances in Chemical Physics, 13,
55–84.

Finemann M, A. (1958) Correlation of bond dissociation energies of polyatomic molecules using


Pauling’s electronegativity concept. Journal of Physical Chemistry, 62, 947–951

Giere, R. (2006). Scientific Perspectivism. Chicago: University of Chicago Press.

Gordy, W. (1946). A New Method of Determining Electronegativity from Other Atomic Properties.
Physical Review, 69, 604-607.

Heidelberger, M. (2003). Theory-Ladenness and Scientific Instruments in Experimentation, in: The


Philosophy of Scientific Experimentation, Hans Radder (Ed.), Pittsburgh: University of
Pittsburgh Press, 138-151 (chapter 7).

Hettema, H. (2012). Reducing Chemistry to Physics. Limits, Models, Consequences. PhD


dissertation. University of Groningen: Createspace Publishing Platform.

Hinze, J. & Jaffe, H. H. (1962). Electronegativity. I. Orbital Electronegativity of Neutral Atoms.


Journal or the American Chemical Society, 84, 540–546.

Jensen, W. B. (1996). Electronegativity from Avogadro to Pauling. Part I: Origins of the


Electronegativity Concept. Journal of Chemical Education, 73, 11–20.

Jensen, W. B. (2003). Electronegativity from Avogadro to Pauling: II. Late Nineteenth- and Early
Twentieth-Century Developments. Journal of Chemical Education, 80, 279 – 287.

Jensen, W. B. (2012) When Was Electronegativity First Quantified?, Part I. Journal of. Chemical.
Education, 89, 94 – 96.

Leach, M. R. (2012). Concerning electronegativity as a basic elemental property and why the
periodic table is usually represented in its medium form. Foundations of Chemistry, 15, 13-
29.

Lewis, G. N. (1916). The Atom and the Molecule. Journal of the American Chemical Society, 38,
762-785.

Lewis, G. N. (1923). Valence and the Structure of Atoms and Molecules. New York: The Chemical
Catalog Company.

Lombardi, O. & Labarca, M. The ontological autonomy of the chemical world. Foundations of
Chemistry, 7, 125 – 48.

23
Lowry, T. M. and Cavell, A. C. (1949). Intermediate Chemistry. London: Macmillan and Company.
Malone, J. G. (1933). The Electric Moment as a Measure of the Ionic Nature of Covalent
Bonds. The Journal of Chemical Physics, 1, 197–199.

Mullay, J. (1987). Estimation of Atomic and Group Electronegativities. Structure and Bonding 66,
1–25.

Mulliken, R. (1934). A New Electronaffinity Scale; together with Data on Valence States and on
Valence Ionization Potentials and Electron Affinities. Journal of Chemical Physics, 2, 782–
793.

Needham, P. (2006). Ontological reduction: A comment on Lombardi and Labarca. Foundations of


Chemistry, 8, 73−80.

Ostwald, W. (1896) Elektrochemie – Ihre Geschichte und Lehre. Leipzig: Verlag von Veit & Comp.

Parr, R. G. Craig, P. D & Ross, I. G. (1950) Molecular Orbital Calculations of Lower excited
electronic Levels of Benzene, Configuration interaction included. Journal of Chemical
Physics, 18, 1561–1563.

Parr, R. G. and Yang, W. (1989). Density-Functional Theory of Atoms and Molecules. New York:
Oxford University Press.

Pauling, L. (1932). The Nature of the Chemical Bond. IV. The Energy of Single Bonds and the
Relative Electronegativity of Atoms. Journal of the American Chemical Society, 54, 3570-
3582.

Pauling, L. (1993). The Formula of Antimonic Acid and Antimonates. Journal of the American
Chemical Society, 55, 1895 – 1900.

Pauling, L. (1950). College Chemistry: An Introductory Textbook of General Chemistry. Second


Edition. San Francisco: W. H. Freeman & Company.

Pritchard, H. O. & Skinner, H. A. (1955) The Concept of Electronegativity. Chemical Reviews, 55,
745 – 786.

Putz, M. V. (2009). Electronegativity: Quantum Observable. International Journal of Quantum


Chemistry, 109, 733–738.

Reynolds, O. (1883). An Experimental Investigation of the Circumstances Which Determine


Whether the Motion of Water Shall be Direct or Sinuous, and of the Law of Resistance in
Parallel Channels. Philosophical Transactions of the Royal Society Londo, 174, 935-982.

Rodebush,W. H.(1925). Compact Arrangement of the Periodic Table. Journal of Chemical


Education, 2, 381-383.

Russell, C.A. (1963). The electrochemical theory of Berzelius. Annals of Science, 19, 127-145.

Ruthenberg, K. and Harré, R. (2012) Philosophy of Chemistry as Intercultural Philosophy: Jaap van
Brakel. Foundations of Chemistry, 14, 193-203.

24
Sanderson, R. (1951). An Interpretation of Bond Lengths and a Classification of Bonds. Science,
114, 670–672.

Sanderson, R. (1952). Electronegativity in Organic Chemistry. Journal of Chemical Education, 29,


539 – 544.

Schomaker, V & Stevenson, P. (1941). Some Revisions of the Covalent Radii and the Additivity
Rule for the Lengths of Partially Ionic Single Bonds. Journal of the American Chemical
Society, 63, 37–40.

Urusov, V. S. (1994). Orbital Electronegativity Concept and its Role in Energetic Crystal Chemistry.
Journal of Structural Chemistry,35, 101–114.

van Brakel, J. (2000) Philosophy of Chemistry. Leuven: Leuven University Press.

Vetter, B. and Schmid S., eds. (2014). Dispositionen. Berlin: Suhrkamp Taschenbuch Wissenschaft.

Zueva, E. M., Galkin V. I., Cherkasov, R. A. and Cherkasov, A. R. (2000). Physical Interpretation of
the Electronegativity. Russian Journal of Organic Chemistry 38: 613–623.

25

You might also like