Halogen Oxidizing Powers

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Struct Chem

DOI 10.1007/s11224-015-0647-z

ORIGINAL RESEARCH

Which halogen is the strongest oxidant? A study with systematics


and surprises
Maja Ponikvar-Svet1 • Diana N. Zeiger2 • Joel F. Liebman2

Received: 9 July 2015 / Accepted: 17 July 2015


 Springer Science+Business Media New York 2015

Abstract Although it is generally accepted that fluorine Introduction


is the strongest oxidant among the halogens and so among
all of the elements, it has not been explained in the liter- The title of the current paper asks the simple question,
ature why this is the case. In this paper, we ask whether this ‘‘Which halogen is the strongest oxidant?’’ Upon remem-
‘‘common knowledge’’ is indeed true; we also explore bering that the category ‘‘halogen’’ includes fluorine,
various means of determining the oxidation strength of the chlorine, bromine, iodine and astatine, there is an imme-
halogens by considering the energetics of binary halides, diate and simple answer: fluorine. The grade of F would
hypohalites and dihalogens. In addition, we use Latimer plausibly be given to the beginning student of chemistry
diagrams to evaluate the redox chemistry of these com- who does not provide this answer. However, why is fluo-
pounds. Ultimately, we find that the conventional wisdom rine the strongest oxidant? What evidence is there for this
holds and that our investigations have provided a clear assertion? An attempt to answer this question is offered in
understanding of why fluorine is, indeed, the strongest the present paper by looking closer at the formation of
oxidant among the halogens. halogen-containing compounds and by investigating both
some experimental and some theoretical variables.
Keywords Fluorine  Halogen  Oxidant  Latimer
diagram  Enthalpy of formation  Redox potential
Definition of oxidation process and related
reactions

In inorganic chemistry, the term oxidation usually


describes a process in which electrons are removed from
a molecule or atom. This criterion is clearly sufficient for
This paper is dedicated to Magdolna (Magdi) Hargittai on the discussion of the oxidation properties of halogens with the
occasion of her 70th birthday. emphasis on fluorine; however, for the sake of com-
& Maja Ponikvar-Svet pleteness, let us also provide the official, more sophisti-
maja.ponikvar-svet@ijs.si cated and precise, definition of oxidation as given by the
& Diana N. Zeiger International Union of Pure and Applied Chemistry
dzeige1@umbc.edu (IUPAC) [1]:
& Joel F. Liebman 1. The complete, net removal of one or more electrons
jliebman@umbc.edu
from a molecular entity (also called ‘‘deelectronation’’).
1
Department of Inorganic Chemistry and Technology, Jožef 2. An increase in the oxidation number of any atom
Stefan Institute, Jamova 39, 1000 Ljubljana, Slovenia within any substrate.
2
Department of Chemistry and Biochemistry, University of 3. Gain of oxygen and/or loss of hydrogen of an organic
Maryland, Baltimore County, Baltimore, MD 21250, USA substrate.

123
Struct Chem

All oxidations meet criteria 1 and 2, and many meet Chlorine forms ClI, and iodine forms I2, but we wonder
criterion 3, but this is not always easy to demonstrate. whether these species qualify as monoiodides. We know of
no classical chemistry for any astatine counterparts; asta-
tine is an artificial element with a half-life of 7.5 h and is
Formation of binary compounds too radioactive for many plausible studies to have been
performed. AtCl, AtBr and AtI have been observed using
Fluorine forms a binary compound with most elements mass spectrometry [24] and thermochromatography [25].
(i.e., all elements except for He, Ne and Ar, as well as those AtF and other astatine fluorides remain uncharacterized in
elements with the highest of atomic numbers that have too the experimental primary chemical literature [26] (this is
short a half-life and/or too low abundance to perform the mainly due to astatine radioactive decay and short lifetime,
necessary synthetic chemistry). Additionally, the resulting perhaps also because these last species are too involatile to
binary fluoride has the same, and even not uncommonly a have been observed by the employed techniques).
higher, oxidation state, than the formally related binary Thus, we deduce that fluorine encourages high oxidation
chloride, bromide and iodide, and (presumably) astatide. states more than do the other halogens and so is the most
For example, the alkali metal halides are generally oxidizing of all of them. Perhaps, we should not be sur-
assumed to have 1:1 M/X stoichiometry for all species prised by this conclusion. After all, the electronegativity of
(e.g., Refs. [2, 3]); the apparent exceptions such as LiF2 fluorine is so high—is it not the highest of all of the ele-
[4], NaCl3 [5], CsBr3 [6] and KI3 [7] contain polyatomic ments? The electronegativity of the halogens drops pre-
anions. Higher polyhalides, such as Cs2I8 [8], will also not cipitously going down the periodic table,
be discussed in the current paper (we acknowledge now F  Cl [ Br [ I. No reference need be given for this
that there are predictions of higher valence cesium fluo- truism from elementary chemistry courses. Now, it is found
rides [9] formed under high-pressure conditions). There are that the electronegativity of fluorine is the second highest
also nearly 20 molecular hexafluorides (16 of which are of all of the elements; so what element is first? This is
well-defined species), but only four hexachlorides and one unequivocally neon, while helium and fluorine are roughly
hexabromide [10], and we note neither hexaiodides nor the same using Allen’s contemporary definition [27, 28].
hexaastatides. Most of these hexahalides are metal, or ‘‘at Electronegativity cannot suffice as an explanation in that
least’’ metalloid, derivatives—sulfur and xenon being the neon and helium are definitively nonoxidizing, if not
sole nonmetals to form such species; indeed, these last two chemically inert. Furthermore, the highest oxidation state
hexahalides are both hexafluorides. for most elements as found in neutral binary compounds is
Let us now consider the nonmetals sulfur, chlorine, with oxygen. While this special value is very commonly
iodine and xenon. They readily form stable and conve- shared with those of fluorine, notable exceptions include
niently isolated hexafluorides, pentafluorides, heptafluo- ðVÞ
nitrogen, chlorine, manganese and xenon, such as N2 O5
rides and hexafluorides, respectively, with central atom
ðVIIÞ ðVIIÞ
oxidation states VI, V, VII and VI, respectively. Xenon versus NðIIIÞ F3 ; Cl2 O7 versus ClðVÞ F5 ; Mn2 O7 versus
octafluoride with oxidation state VIII is a source of con- MnðIVÞ F4 ; and XeðVIIIÞ O4 versus XeðVIÞ F6 . The corre-
tention as to its isolability under any set of experimental sponding high-valence binary fluorides do not exist. Should
conditions [11–13]. With chlorine, sulfur forms only the we allow the consideration of binary ionic species? While
dichloride and the thermally labile tetrachloride [14]; N(V) is found in both [NO3]- and [NF4]?salts, and Cl(VII) is
chlorine forms the monochloride [dare we use the name, or found in [ClO4]- and [ClF6]? salts, Xe(VIII) is found in
even the descriptor chlorine(I) chloride for Cl2]; iodine Xe(VIII)O4; however, its existence remains unknown, or at
forms the trichloride with an early suggestion of the pen- best debatable, in both binary ionic and neutral fluorides.
tachloride [15]; finally, xenon forms the fragile tetrachlo- The record for the experimentally observed highest oxi-
ride [16, 17]. With bromine, sulfur forms the dibromide, dation state is IX, found in admittedly gaseous [IrO4]?
and chlorine and bromine form the monobromide [again, [29]; this oxidation state has yet to be found in any salt and
dare we refer to BrCl as chlorine(I) bromide or even use the is unmatched by any iridium fluoride. It may be argued that
formula ClBr]. There is evidence for ClBr3 [18], but the there is too much steric hindrance for [Xe(VIII)F7]? for this
suggested structure is Cl–Br(Br)–Br, and so it is better species to exist, for example as a product from the plau-
understood as a bromine(III)-mixed halide than as chlorine sible reaction of [KrF]? and XeF6 [30]. However, the
tribromide. Iodine tribromide [19] and xenon dibromide isoelectronic IF7 has long been known, and [XeF7]- with
[20] have also been proposed. With iodine, sulfur forms an additional lone pair is a viable anion and has been
S2I2 [21] and has also been suggested to form another spectroscopically characterized as its Cs? and NF4? salts
binary species, SI2, recognized as a possible product in a [31]. Relatedly, Xe(VIII)F8 remains unformed from the
classical synthesis and cryogenic matrix reaction [22, 23]. decomposition of the aforementioned [NF4]?[Xe(VI)F7]-,

123
Struct Chem

while [Xe(VI)F8]2- is well established in salts wherein that F \ Cl. While fluorine has thus been spoken of as
NF4? is among accompanying cations. (We also note that perhaps anomalous [44–46], and an approximately
[I(VII)F8]-, an ion isoelectronic to XeF8
ðVIIIÞ
, forms stable 110 kJ mol-1 ‘‘correction’’ factor has been suggested
salts [32]). therein for fluorine with eight valence electrons in any
environment, these data are experimentally incontrovert-
ible. There is an approximately 55 kJ mol-1 spread
Enthalpies of formation of binary halides between the electron affinities of all four elements. We
now cite a set of earlier measurements [47] for the electron
The same conclusion that fluorine is the most oxidizing of affinities of F, Cl, Br, I. These values are less precise than
the halogens may be drawn by looking at the enthalpies of those noted above, but are more than accurate enough to
formation of related binary halides. Consider the binary confirm the order F \ Cl [ Br [ I. This study additionally
halides of carbon, CF4, CCl4, CBr4 and CI4 (all from Ref. allowed for distinguishing the energies of the two processes
[33], save the last from Ref. [34]); of boron BF3, BCl3, BBr3 2
P3=2 XðgÞ þ e ! X ð2Þ
ðgÞ
and BI3 (all from Ref. [35]); of sodium halides, NaF, NaCl,
NaBr, NaI (all from Ref. [35]); and those of the other alkali
2
P1=2 XðgÞ þ e ! X
ðgÞ ; ð3Þ
metals (all from Ref. [35]). The same order is found for the
where it is acknowledged that few discussions of atomic
gas phase phosphorus(III) halides [35], where the data for
halogens distinguish between the two spin–orbit states. If
PI3 were derived as the sum of the enthalpies of formation
one considers the 2P3/2–2P1/2 spin–orbit averaged value for
[35] and of sublimation [36]. The enthalpies of formation
the two atomic states, the order I \ F \ Br \ Cl is found,
become less negative or equivalently, more positive, and the
where the total spread is reduced to approximately
species become less thermodynamically stable, in the order
30 kJ mol-1, and I and F are found to differ by only
fluoride \ chloride \ bromide \ iodide. This is consistent
10 kJ mol-1. Do we wish to conclude that fluorine and
with the electron transfer to halogen, and hence the ionicity
iodine have essentially equal oxidizing power?
and electrostatic stability being the greatest for fluoride, and
With what other species can comparisons be made?
then for chloride, bromide and iodide, decreasing in that
What other species educate us as to the oxidizing power of
order. These statements are all consistent and sensible, but
the halogens? Let us remain, however, temporarily, in the
… The same order of enthalpies of formation is experi-
gas phase. The first comparison is made with the diatomic
mentally found for OF2 [35] \ OCl2 [37] \ OBr2 [38], and
halogens, F2, Cl2, Br2, I2. That is, we consider the ener-
in addition, from the analysis of high-level quantum
getics of the related reactions
chemical calculations that are accurate to approximately
10 kJ mol-1 [39] for both species: OBr2 * OI2. Most X2ðgÞ þ e ! ½X2 
ðgÞ : ð4Þ
assuredly, the oxygen in OF2 is expected to have positive
From any of three references [48–50] with alternative
oxygen and negative fluorine, while the other oxygen
methods of determining the corresponding electron affini-
halides likewise have negative oxygen and positive halo-
ties for all four of these species, we find the following order
gen. We may ask how, then, can we compare oxygen
of electron affinities: F2  Cl2 * Br2 \ I2, with a total
difluoride to the other oxygen dihalides?
spread of nearly 100 kJ mol-1. That the electron affinity
value for the fluorine-containing species is greater than
those for the corresponding compounds with chlorine and
Electron affinities and comparison of dihalogens the other halogens is sensible and satisfying. However, it is
perhaps disconcerting that the molecular electron affinities
At this point, we turn to a discussion of the electron are smaller than those of the corresponding atomic quan-
affinities of the halogens as part of the understanding of the tities until it is realized that it is an antibonding, and hence
electron-attracting power of these elements. After all, is it relatively destabilizing, molecular orbital in which the
not the halogen’s electron affinity, the exothermicity of the ‘‘new’’ electron goes. Indeed, it may well be surprising that
gas phase reaction the diverse [X2]- are bound at all.
XðgÞ þ e ! X
ðgÞ ; ð1Þ

the simplest, most unambiguous measure of the electron Diatomic halogens and aqueous halides
attraction, the oxidizing power of an element? Let us take
the electron affinity values of F, Cl, Br, I from exquisitely We now turn to the simplest of all of the redox couples that
precise studies [40–43]. Doing so recovers the expected connecting the diatomic halogens to the aqueous halide
order of oxidizing power Cl [ Br [ I, but then, it is seen ions. To return us to more fundamental thermochemical

123
Struct Chem

quantities, we consider the separate standard enthalpies, decreasing, well-known, conventional, oxidizing power
Gibbs energies and temperature ‘‘corrected’’ entropies of order F  Cl [ Br [ I.
formation, i.e., DfH, DfG and TDfS where T = 298.15 K.
Here we consider the process
Latimer diagrams
X2 þ 2e ! 2X : ð5Þ
The exothermicity and exoergicity decline significantly The redox chemistry associated with both [OX]- and HOX
in the classical order F  Cl [ Br [ I whether the dia- (as well as with other halogen oxygen-containing species)
tomic halogen is taken in the proper standard phase (F2, gas; may be put into context by the use of reduction potential
Cl2, gas; Br2, liquid; I2, solid), all in gaseous or all in diagrams originally given by Latimer [53]. The diagrams
aqueous solution themselves. It may be surprising to the show the standard reduction potentials connecting the
reader that the choice of phase is quite unimportant. Using various oxidation states of a given element including the
the data in Wagman [35] for what follows here, we confirm skip potentials. They make it possible to deal with complex
that the enthalpy and entropy are in opposition to each systems; the table of redox potentials is organized by the
other. Gaseous dihalogens possess higher entropy than voltage of the process, rather than by the chemical species
those in the aqueous phase, and so more entropy is lost upon involved, and so the table does not provide a qualitative
reduction of the ‘‘lighter species’’ gaseous F2 and Cl2 than understanding of the redox system. Changes in free energy
of the ‘‘heavier’’ species, liquid Br2 and solid I2, to form (dependent on concentration, temperature, the presence of
aqueous halide ion. However, gaseous halogens have more other non-inert reagents and pH) affect redox potentials. In
positive enthalpy than do their condensed-phase counter- practice, pH changes are the most important factor in
parts: to the defined 0.0 kJ mol-1 of the standard state are aqueous chemistry that influence redox potentials because
added the enthalpies of vaporization and sublimation of large numbers of H? (or, to be more precise and correct,
liquid Br2 and solid I2, respectively. We know that the [H3O]?) and OH- ions are usually involved in redox half-
enthalpy and entropy must roughly cancel for the unsol- reactions. For this reason, it has become conventional to
vated elements because the free energies of vaporization construct Latimer diagrams for the two extremes, at
and sublimation necessarily must equal 0.0 kJ mol-1 at the pH = 0 and pH = 14 (or more precisely, 13.996). Poten-
phase change temperature, and these temperatures are quite tial diagrams in acid and alkaline solution constructed by
close to the 298.15 K that defines the thermodynamic using the standard electrode potentials as reported by
standard state. The exothermicity associated with the for- Bratsch [54] are presented in Figs. 1 and 2.
mation of the aqueous halide ions, or equivalently with the The Latimer diagrams clearly illustrate that fluorine has
negative of their enthalpies of formation, declines steeply in the highest oxidizing potential not only among halogen
the order F  Cl [ Br [ I; the reader may recall that the elements but also among all the elements of the periodic
four values were not that different for the gaseous species. It table. Is this surprising because fluorine is not the most
is not surprising that anion solvation energies decrease in electronegative element? Actually, it is not, because, unlike
the order F  Cl [ Br [ I as befits ever-increasing atomic electronegativity, oxidizing potential is a measured vari-
size, decreasing localized size/charge ratio and decreasing able. The potentials for astatine are estimated values only
ion-water hydrogen bond energies. Measurements of the and for this reason will be omitted from the following
halide–water hydrogen bond/complexation enthalpies for discussion. The given redox potentials illustrate the higher
one halide ion with 1, 2, 3,…, n water molecules in the gas oxidizing potentials of halogens and their anions in acid
phase were reported (see the pioneering studies [51, 52]). than in alkaline solution. The only exception to this is the
This solvation corresponds to ‘‘freezing’’ or localizing ClO3-/ClO2 couple with a negative (reducing properties)
the solvating waters around the anion. Solvation entropies potential. In fact, ClO2 (as OClO, not its isomer ClOO) is
significantly increase in the order F \ Cl \ Br \ I. Indeed, the only isolable halogen-containing compound with
the absolute entropy of aqueous F- is negative. Numeri- halogen in the (IV) oxidation state. When relating [XO]-
cally, the value is –13.8 J mol-1 K-1, while that of aque- and HXO to X2 in acid media, a decrease in redox poten-
ous Cl-, Br- and I- are positive (56.5, 82.4 and tials from Cl to I is observed. In alkaline media, the
111.3 J mol-1 K-1, respectively) and their neutral, iso- potential of the BrO-/Br2 couple is, however, the highest:
electronic, noble gas counterparts (Ne, Ar, Kr, Xe) are the order is Br [ Cl [ I  At. It is interesting to note that
approximately equal, clustered around 62 ± 4 J mol-1 - the redox potentials for [XO]-/X- couples are given [54]
K-1. We noted previously that fluorine has an anomalously for alkaline media only and follow the expected order of
low electron affinity, and it apparently also has an Cl [ Br [ I; in acid media, we must use the data for the
anomalously negative entropy of solvation. It appears that conjugate acids and the HXO/X- couples. A look at higher
it is the solvation energy order that generates the oxidation states shows that the potentials of XO3-/XO-

123
Struct Chem

Fig. 1 Standard potentials for


systems involving halogens and
their anions in acid solution at
pH = 0.000 (E in V).
Estimated values are given in
parentheses

couples are given for acid media only, and again, the [XO]-, we will use both interchangeably. Relatedly,
potential of the BrO-/Br2 couple is the highest; the order is hypohalous acids have been written as HOX, reflecting the
Br [ Cl [ I. Figure 1 clearly illustrates the diversity of oxygen–hydrogen bond, and as HXO, such as chlorous,
redox reactions proceeding in aqueous media. It is there- chloric and perchloric acids as well as most other oxyacids
fore not surprising that redox titrations involving mostly such as sulfuric and nitric acids, H2SO4 and HNO3. As in
bromine and iodine and their anions in acid solutions are the case of [OX]- and [XO]- also for hypohalous acids,
advantageously used in classical analytical chemistry [55, HOX and HXO will be used interchangeably.
56]. The reader will note that we ignore here any system- Consider now the oxygen monohalides, OX, in which
atic discussion of redox potentials involving heptavalent there is a half-filled antibonding molecular orbital in which
halogen because the derived oxoanions are not the same for the ‘‘new’’ electron goes. That is, we consider the ener-
all three halogens. Cl(VII) and Br(VII) are unequivocally getics of the related reactions
tetracoordinate unprotonated ions, while I(VII) includes OXðgÞ þ e ! ½OX ð6Þ
ðgÞ
both tetra and (hydrated) hexacoordinate ions of diverse
degrees of protonation. Nonetheless, we note that there is The measured [57] electron affinity order here is
no ambiguity in the assertion that Br [ Cl for the [XO4]-/ F * Cl \ Br * I, and the total spread is but 10 kJ mol-1.
[XO3]- couples. Do we wish to conclude that the oxidizing power of all four
halogens is numerically nearly the same? Or do we con-
clude that the oxidizing power of iodine is higher than that
Oxygen monohalides and hypohalites of fluorine?
We recognize [OX]- as the long-known, trivially named
Let us now turn to a discussion of compounds containing hypohalite ions, all recognized in salts and aqueous solu-
both oxygen and halogens. Before proceeding, we note that tions alike except for the case of X = F, for which neither
hypohalite ions have been written in the literature both as salts nor solutions are known to the experimental chemist
[OX]- and as [XO]-. The former, written with the ‘‘O’’ (LiOF has been seen in a cryogenic matrix [58]); see Ref.
first, is reminiscent of the well-known hydroxide and [59, 60] for theoretical discussion of such compounds).
acetate ions, [OH]- and [OAc]-, respectively. The latter is These species are the conjugate bases of the hypohalous
more consistent with other oxohalogen anions such as acids, HOX, and indeed the anions and these acids have
chlorite, chlorate and perchlorate [ClO2]-, [ClO3]- and been studied separately, except for the case of X=F, for
[ClO4]-. Rather than deciding between using [OX]- or which ‘‘the pK of HOF is probably [7’’ [61]. We find that

123
Struct Chem

Fig. 2 Standard potentials for


systems involving halogens and
their anions in alkaline solution
at pH = 13.996 (E in V).
Estimated values are given in
parentheses

the reduction potentials for the [OX]- ions (X=Cl, Br, I) in Cl, Br and I species. Whether from high-level quantum
alkaline media are all rather close, as are the corresponding chemical calculations [62–64] or from experimental ref-
reduction potentials of HOX (X=Cl, Br, I) in acid media. erences cited therein, the enthalpy of formation of gaseous
[OCl]- is a common oxidant and is found in many HFO is slightly more negative than that of gaseous HClO
household bleach products. However, this choice of species (although the individual values vary from reference to
does not reflect its power as an oxidant; rather, consider- reference, the value for HFO is found in each source to be
ations of ease, safety and economics presumably dictate roughly 10 kJ mol-1 more negative than that for HClO).
this choice of species. Equivalently, as the enthalpies of formation of gaseous F2
and Cl2 are 0.0 kJ mol-1 by definition, the gas phase
reaction
Hypofluorous acid, HFO and hypofluorite ion, HFOðgÞ þ 1=2Cl2ðgÞ ! HClOðgÞ þ 1=2F2ðgÞ ð7Þ
[FO]2
is endothermic by the aforementioned *10 kJ mol-1.
We now shift our attention to the comparison of hypoflu- Given the range of literature values for the enthalpies of
orous acid, HFO, and hypofluorite ion, [FO]-, with the formation of HFO and HClO, we believe that it is safe to
corresponding species containing the other halogens. More say that the reaction is roughly thermoneutral if not
properly, we need only compare HFO and [FO]- with endothermic. If we assign a positive oxidation state to the
HClO and [ClO]- because we earlier compared the related halogen atoms in both HFO and HClO, we would conclude

123
Struct Chem

that F(I) is comparably if not less oxidizing than Cl(I). Of that HFO and [FO]- are much more powerful oxidants than
course, the reader may protest and assert that hypofluorous their chlorine-containing counterparts? HFO needs ‘‘tam-
acid does not have positive fluorine, and an immediate ing’’ [68] and [FO]- salts cannot even be isolated, whereas
conclusion may be that any comparison between HFO and the chlorine-containing species are household products.
HClO is inconsistent and hence invalid. We accept the There is a simple answer: elemental fluorine and elemental
inconsistency but still proceed. chlorine are not the products of any experimentally realized
Turning now to aqueous media, consider now the redox reaction of interest. Rather, the products are aqueous
reaction F- and Cl-, or their protonated equivalents HF and HCl,
HFOðaqÞ þ 1=2 Cl2ðgÞ ! HClOðaqÞ þ 1=2 F2ðgÞ ð8Þ and we have already shown that F2 is a vastly superior
oxidant to Cl2. The energetics of the total redox reactions
We recall that the enthalpy of solvation is the difference of hypohalite and hypohalous acids to halide ion vastly
between the enthalpy of formation of a gaseous species and favor F(I) over Cl(I).
that of its infinitely dilute aqueous solution. The enthalpy
of solvation of HFO is not known from experiment. Can we
derive this quantity from the archival [35] value for HClO, Closing remarks
or can we at least extrapolate the enthalpy of solvation of
HFO through calculating the difference between the We conclude that fluorine is the most powerful oxidant
enthalpy of formation of HFO and HClO? It is plausible among the halogens. However, the reasons and the sup-
that the difference between the enthalpies of solvation of porting data are often obscure, if not contradictory. We
ClOH and FOH is approximately the same difference as provide the necessary reconciliation here, accompanied by
that between ClCH2COOH and FCH2COOH (more pre- clarification.
cisely, the enthalpies of solvation of substituted acetic Given Magdi Hargittai’s numerous writings on the sci-
acids, XCH2COOH [65], have been successfully approxi- ence-humanities interface, we opt to close our paper with
mated [66] by calculating the sum of the solvation the words of the poet Eliot [69], who wrote in ‘‘The Little
enthalpies of XCH3 and HCOOH. The HCOOH cancels Gidding.’’
out; thus, we are saying that the difference between ClOH
We shall not cease from exploration
and FOH is comparable to that of the difference between
And the end of all our exploring
ClCH3 and FCH3. Alternatively, we are saying that the
Will be to arrive where we started
contribution of the component CH2CO is roughly the same
And know the place for the first time.
for both acetic acids). The difference between the enthal-
pies of solvation of FCH2COOH and ClCH2COOH is small
Acknowledgments One of the authors (MPS) gratefully acknowl-
[65, 67], and so, the difference between the enthalpies of edges the Slovenian Research Agency (ARRS Grant P1-0045, Inor-
solvation of HFO and HClO is, presumably, also small. ganic Chemistry and Technology) for financial support.
Therefore, the difference between the enthalpies of for-
mation of aqueous HFO and HClO is small. We therefore
conclude that the reaction of interest is roughly ther- References
moneutral and that F(I) and Cl(I) are comparably oxidizing.
In the previous paragraph, we derived that the enthalpy of 1. International Union of Pure and Applied Chemistry, Com-
pendium of chemical terminology, Gold Book, Version 2.3.3
formation of aqueous HFO and HClO are nearly the same.
(Accessed July, 2015)
So, too, are the enthalpies of formation of gas phase [FO]- 2. Karapet’yants MK (1960) Zh Strukt Khim 1:399–403
and [ClO]- [62], with the latter being somewhat more stable. 3. Ladd MFC, Lee WH (1961) J Inorg Nucl Chem 20:163–165
From this information, we deduce that the reaction 4. Andrews L (1976) J Am Chem Soc 98:2147–2152
5. Zhang W, Oganov ER, Goncharov AF, Zhu Q, Boulfelfel SE,
½FO 1 
ðgÞ þ =2Cl2 ðgÞ ! ½ClOðgÞ þ =2F2ðgÞ
1 ð9Þ Lyakhov AO, Stavrou E, Somayazulu M, Prakapenka VB,
Konopkova Z (2013) Science 342:1502–1505
is roughly thermoneutral. If we now assume that the dif- 6. Harris IWH (1932) J Chem Soc 2709-2713
ference between the hydration enthalpies of [ClO]- and 7. Jakowkin AA (1896) Z Phys Chem Stoch Ve 20:19–39
8. Babkov AV, Stepin BD (1968) Zhu Neorg Khim 13:11–14
[FO]- is approximately the same as that between
9. Miao M (2013) Nat Chem 5:846–852
[ClCH2COO]- and [FCH2COO]-, we deduce that the 10. Seppelt K (2015) Chem Rev 115:1296–1306
reaction 11. Slivnik J, Volavšek B, Marsel J, Vrščaj V, Šmalc A, Frlec B,
Zemljič Z (1963) Croat Chem Acta 35:81–82
½FO 1 
ðaqÞ þ =2 Cl2ðgÞ ! ½ClOðaqÞ þ =2 F2ðgÞ
1 ð10Þ 12. Frlec B, Holloway JH, Slivnik J, Šmalc A, Volavšek B, Zemljič
A (1970) J Inorg Nucl Chem 32:2521–2527
is roughly thermoneutral as well. F(I) and Cl(I) are com- 13. Grant DJ, Wang TH, Dixon DA, Christe KO (2010) Inorg Chem
parably oxidizing. Again, then, why is it normally assumed 49:261–270

123
Struct Chem

14. Steudel R, Jensen D, Plinke B (1987) Zeit Naturforsch B 42. Blondel C, Cacciani P, Delsart C, Trainham R (1989) Phys Rev A
42:163–168 40:3698–3701
15. Hannay JB (1879) J Chem Soc Trans 35:169–171 43. Pelaez RJ, Blondel C, Delsart C, Drag C (2009) J Phys B Atom
16. Perlow GJ, Perlow MR (1964) J Chem Phys 41:1157–1158 Mol Opt Phys 42:125001/1–125001/7
17. Haner J, Schrobilgen GJ (2015) Chem Rev 115:1255–1295 44. Politzer P (1969) J Am Chem Soc 91:6235–6237
18. Nelson LY, Pimentel GC (1968) InorgChem 7:1695–1699 45. Politzer P, Timberlake JW (1972) J Org Chem 37:3557–3559
19. Forbes GS, Faull JH Jr (1933) J Am Chem Soc 55:1820–1830 46. Balighian ED, Liebman JF (2002) J Fluor Chem 116:35–39
20. Perlow GJ, Yoshida H (1968) J Chem Phys 49:1474–1478 47. Berry SR, Reimann CW (1963) J Chem Phys 38:1540–1543
21. Feher F, Münzner H (1963) Chem Ber 96:1131–1149 48. DeCorpo JJ, Franklin JL (1971) J Chem Phys 54:1885–1888
22. Schneider E (1948) Süddeut Apothek-Zeit 88:449–450 49. Chupka WA, Berkowitz J, Gutman D (1971) J Chem Phys
23. Feuerhahn M, Vahl G (1980) Inorg Nucl Chem Lett 16:5–8 55:2724–2733
24. Appelman EH, Sloth EN, Studier MH (1966) Inorg Chem 50. Ayala JA, Wentworth WE, Chen ECM (1981) J Phys Chem
5:766–769 85:768–777
25. Merinis J, Legoux Y, Bouissires G (1972) Radiochem Radioanal 51. Arshadi M, Yamdagni R, Kebarle P (1970) J Phys Chem
Lett 11:59–64 74:1475–1482
26. Riedel S (2013) In: Reedijk J, Poeppelmeier K (eds) Compre- 52. Hiraoka K, Mizuse S, Yamabe S (1988) J Phys Chem
hensive inorganic chemistry II, vol 2, 2nd edn. Elsevier, 92:3943–3952
Amsterdam 53. Latimer WM (1938) Oxidation potentials. Prentice Hall, New
27. Allen LC (1989) J Am Chem Soc 111:9003–9014 York
28. Mann JB, Meek TL, Allen LC (2000) J Am Chem Soc 54. Bratsch SG (1989) J Phys Chem Ref Data 18:1–21
122:2780–2783 55. Beck CM II (1994) Anal Chem 66:224A–239A
29. Wang G, Zhou M, Göttel JT, Schrobilgen GJ, Su J, Li J, Schlöder 56. Beck CM II (1997) Metrologia 34:19–30
T, Riedel S (2014) Nature 514:475–477 57. Gilles MK, Polak ML, Lineberger WC (1992) J Chem Phys
30. Christe KO, Dixon DA (1992) J Am Chem Soc 114:2978–2985 96:8012–8020
31. Christe KO, Wilson WW (1982) Inorg Chem 21:4113–4117 58. Andrews L, Raymond JI (1971) J Chem Phys 55:3078–3086
32. Adams CJ (1974) Inorg Nucl Chem Lett 10:831–835 59. O’Hare PAG, Wahl AC (1970) J Chem Phys 53:2469–2478
33. Pedley JB (1994) Thermochemical data and structures of organic 60. Liebman JF (1972) J Chem Phys 56:4242–4243
compounds, vol 1., TRC data seriesTRC, College Station 61. Appelman EH, Thompson RC (1984) J Am Chem Soc
34. Carson AS, Laye PG, Pedley JB, Welsby AM (1993) J Chem 106:4167–4172
Thermodyn 25:261–269 62. Glukhovtsev MN, Pross A, Radom L (1996) J Phys Chem
35. Wagman DD, Evans WH, Parker VB, Schumm RH, Halow I, 100:3498–3503
Bailey SM, Churney KL, Nuttall RL (1982) The NBS tables of 63. Rayne S, Forest K (2011) Comput Theor Chem 974:163–179
chemical thermodynamic properties: selected values for inorganic 64. Karton A, Parthiban S, Martin JML (2009) J Phys Chem A
and C1 and C2 organic substances in SI units. J Phys Chem Ref 113:4802–4816
Data 11(Suppl 2):1–392 65. Haberfield P, Rakshit AK (1976) J Am Chem Soc 98:4393–4394
36. Finch A, Gardner PJ, Hameed A (1970) J Inorg Nucl Chem 66. Haberfield P (1980) J Chem Educ 57:346–347
32:2869–2874 67. Wilson B, Georgiadis R, Bartmess JE (1991) J Am Chem Soc
37. Thorn RP Jr, Stief LJ, Kuo SC, Klemm RB (1996) J Phys Chem 113:1762–1766
100:14178–14183 68. Appelman EH, Dunkelberg O, Kol M (1992) J Fluor Chem
38. Thorn RP Jr, Monks PS, Stief LJ, Kuo SC, Zhang Z, Klemm RB 56:199–213
(1996) J Phys Chem 100:12199–12203 69. Eliot TS (2014) The little gidding. In: Four quartets. http://www.
39. Grant DJ, Garner EB, Matus MH, Nguyen MT, Peterson KA, columbia.edu/itc/history/winter/w3206/edit/tseliotlittlegidding.html.
Francisco JS, Dixon DA (2010) J Phys Chem A 114:4254–4265 Accessed Aug 2014
40. Blondel C, Delsart C, Goldfarb F (2001) J Phys B Atom Mol Opt
34:L281–L288
41. Berzinsh U, Gustafsson M, Hanstorp D, Klinkmuller A, Ljung-
blad U, Martenssonpendrill AM (1995) Phys Rev A 51:231–238

123

You might also like