Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

FULL PAPER

DOI: 10.1002/chem.201102960

Halogenated Benzene Cation Radicals

Matthias J. Molski,[a] Doreen Mollenhauer,[a] Sebastian Gohr,[a] Beate Paulus,[a]


Monther A. Khanfar,[b] Hashem Shorafa,[a] Steven H. Strauss,[c] and Konrad Seppelt*[a]

Abstract: The halogenated benzenes cation radicals [C6HF5]+ and 2,3,5,6- In the case of C6F6+Sb2F11 , two low-
C6HF5, 2,4,6-C6H3F3, 2,3,5,6-C6H2F4, [C6H2F4]+ had structures with the same symmetry geometries were observed in
C6F6, C6Cl6, C6Br6, and C6I6 were con- geometry as C6HF5 and 2,3,5,6- the same crystal. Interestingly, the
verted into their corresponding cation [C6H2F4]. In contrast, the cation radi- structures of the cation radicals 2,4,6-
radicals by using various strong oxi- cals [C6F6]+, [C6Cl6]+, and possibly also [C6H3F3]+ and C6I6+ did not exhibit
dants. The cation-radical salts were iso- [C6Br6]+ exhibited Jahn–Teller-distort- Jahn–Teller distortions. DFT calcula-
lated and characterized by electron ed geometries in the crystalline state. tions showed that the explanation for
paramagnetic resonance (EPR) spec- the lack of distortion of these cations
troscopy and by single-crystal X-ray from the D3h or D6h symmetry of the
Keywords: benzene · cations · crys-
diffraction. The thermal stability of the neutral benzene precursor was differ-
tal structures · Jahn–Teller distor-
cation radicals increased with decreas- ent for 2,4,6-[C6H3F3]+ than for [C6I6]+.
tion · X-ray diffraction
ing hydrogen content. As expected, the

Introduction zenes, see the Supporting Information, Table SI-1). By anal-


ogy to other high-IP molecules, such as O2 (12.07 eV), Cl2
Benzene, C6H6, is one of the archetypal organic molecules (11.51 eV), Br2 (10.51 eV), and I2 (9.22 eV), all of which
owing to its D6h symmetry, aromaticity, ring current, and sta- have been successfully oxidized to produce cation radicals
bility. Because its exceptional stability is related to the de- that were isolable without further reaction (i.e., [O2]+,[4]
localization of six p electrons, it is expected that variation of [Cl4]+,[5] [Br2]+,[6] and [I2]+,[7] respectively), we reasoned that
the electron count, such as in [C6H6]+, [C6H6]2+, and some benzene cation radicals should be obtainable.
[C6H6] , would result in highly reactive species. Herein, we The cation radical [C6H6]+ has been observed spectroscop-
focused on the cation radicals C6X6+, with X = F, Cl, Br, and ically in the gas phase.[8, 9] However, in a condensed phase, it
I (hereafter, to avoid confusion, the generic term “benzene” may be the most-difficult benzene cation radical to study.
is used to refer to C6H6 and all of its possible monocyclic de- Although C6H6 has a lower IP than the other benzenes stud-
rivatives, the formula “C6H6” is used to refer specifically to ied herein, side-reactions following oxidation, such as the
the parent molecule). In the gas phase, C6H6 has a first ioni- loss of a proton, may be very rapid and therefore unavoida-
zation potential (IP) of 9.25(2) eV,[1] and the substitution of ble. Nevertheless, EPR spectra have been observed. The ox-
hydrogen by halogen atoms causes this to increase in a sys- idation of benzene by persulfate in concentrated sulfuric
tematic manner[1–3] (for a list of IP values for relevant ben- acid gave a poorly resolved seven-line spectrum, and a reac-
tion mixture containing C6H6 and NbF5 has been interpreted
as that of the ion-pair [C6H6+Nb2F11 ].[10] On the other hand,
[a] M. J. Molski, D. Mollenhauer, S. Gohr, Prof. Dr. B. Paulus,
C6F6, with an IP of 9.88(5) eV,[1] should be the most-difficult
Dr. H. Shorafa, Prof. Dr. K. Seppelt benzene to oxidize, but, once obtained, its cation radical
Freie Universitt Berlin might be more stable than all other benzene cation radicals.
Institut fr Chemie und Biochemie As we will show below, the stability of [C6H6-nFn]+ cation
Fabeckstrasse 34-36, 14195 Berlin (Germany)
radicals increases as the value of n increases.
Fax: (+ 49) 30-838-54289
E-mail: seppelt@zedat.fu-berlin.de In addition to demonstrating that benzene cation radicals
[b] Prof. M. A. Khanfar can be isolated in a condensed phase, the central focus of
Chemistry Department, The University of Jordan this work was the determination of their structures. For neu-
Amman (Jordan) tral precursors with D6h and D3h symmetry, Jahn–Teller dis-
[c] Prof. S. H. Strauss tortions are expected following oxidation into the cation
Department of Chemistry, Colorado State University radicals. But what particular lower-symmetry structure
Fort Collins, CO 80523 (USA)
should be present for a given cation radical in a given con-
Supporting information for this article, including additional informa-
tion on selected structures, calculated orbital energies, and EPR spec-
densed-phase environment is an open question. It has been
tra, is available on the WWW under http://dx.doi.org/10.1002/ shown theoretically that gas-phase [C6H6]+, [C6F6]+, and
chem.201102960. 2,4,6-[C6H3F3]+ should distort into D2h ([C6H6]+, [C6F6]+) or

Chem. Eur. J. 2012, 00, 0 – 0  2012 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim &1&
These are not the final page numbers! ÞÞ
C2v structures (2,4,6-[C6H3F3]+).[11–21] However, in each case, Table 2. Figures 1–6 show the structures of most of the ben-
two alternative structures were predicted that were very zene cation radicals and, only where relevant, details of
close in energy: a quinoid structure and a bis-allyl-like struc- cation–anion interactions.
ture (hereinafter referred to as the bis-allyl structure or ge- DFT-predicted relative energies for three limiting geome-
ometry). The quinoid cations were predicted to have two tries for the isolated [C6X6]+ cation radicals are listed in
short C C bonds and four longer ones, whereas the bis-allyl Table 3 (X = F, Cl, Br, I). DFT-predicted C C bond lengths
cations were predicted to have four short and two longer C were compared with their corresponding lengths as deter-
C bonds. mined by X-ray crystallography for neutral molecules C6X6
All benzene cation radicals can be unstable at room tem- and for the [C6X6]+ cation radicals in Table 4. Similar com-
perature and hence susceptible to secondary reactions. parisons were made between the [C6HF5]+, 2,4,6-[C6H3F3]+,
Therefore, our plan was to first detect these species qualita- and 2,3,5,6-[C6H2F4]+ cation radicals in Table 5, which also
tively by their often intense color, then to study them in so- includes the DFT-predicted C C bond lengths for the
lution by EPR spectroscopy, and finally to isolate them as 2,4,5,6- and 3,4,5,6-[C6H2F4]+ cation radicals.
their crystalline salts and determine their structures by X-
ray diffraction. The latter part of the plan was deemed nec-
essary because the Jahn–Teller distortion of benzene cation Discussion
radicals, which were made from a D6h-symmetry- or D3h-
symmetry-neutral benzene precursor, should not be observa- ACHTUNGRE[C6F6]+: This cation radical may have been first observed by
ble in isotropic EPR spectra owing to the rapid interconver- EPR spectroscopy in a mixture of C6F6 and SbF5, which ex-
sion of equivalent Jahn–Teller-distorted isomers in solution. hibited hyperfine splittings that were attributable to six
However, the observation of Jahn–Teller distortions by crys- equivalent F atoms.[22] Later, the yellow compound [C6F6]+-
tallography could be hampered by positional disorder of the ACHTUNGRE[AsF6] was obtained by oxidizing C6F6 with [O2]+ACHTUNGRE[AsF6] ,
pseudo-hexagonal cations. In an attempt to avoid this prob- but information about the structure of the cation radical
lem, special emphasis was placed on growing single crystals could not be obtained.[23, 24] It has also been reported that
very slowly at low temperatures. CrF5 can oxidize C6F6 into [C6F6]+.[25]
In 2009, we reported that OsF6/SbF5 or [O2]+ACHTUNGRE[Sb2F11] oxi-
dized C6F6 into [C6F6]+ in anhydrous HF solution. Our EPR
Results spectrum,[26] the parameters of which are listed in Table 1,
was identical to the previously published spectrum.[22] Two
EPR g values and hyperfine coupling constants (aF and aH) salts, [C6F6]+ACHTUNGRE[Os2F11] and [C6F6]+ACHTUNGRE[Sb2F11] , were isolated
for solutions of [C6HF5]+ACHTUNGRE[AsF6] and 2,3,5,6-[C6H2F4]+- and structurally characterized (they were found to be iso-
ACHTUNGRE[SbF6] in SO2, [C6F6]+ACHTUNGRE[Sb2F11] and [C6Cl6]+ACHTUNGRE[Sb2F11] in structural).[26] The X-ray structures revealed two pleasant
SO2ClF, and microcrystalline [C6I6]+ACHTUNGRE[AsF6] and [C6I6]+- surprises: 1) the two independent, Jahn–Teller-distorted
ACHTUNGRE[SbF6] are listed in Table 1. The EPR spectra of [C6F6]+- [C6F6]+ cation radicals in the asymmetric unit were not dis-
ACHTUNGRE[Sb2F11] , [C6HF5]+, and [C6H2F4]+ are shown in the Sup- ordered, and 2) they had different distorted geometries
porting Information, and matched those reported in the lit- (Figure 1). In addition, both structures were determined
erature.[22] very precisely. The R1 value and the standard error for indi-
vidual C C bond lengths for [C6F6]+ACHTUNGRE[Sb2F11] were 0.015 and
either 0.3 or 0.4 pm, respectively. The two “bond-stretch iso-
Table 1. EPR spectroscopic parameters.[a]
mers”[27] discovered in the same unit cell were indeed the
Compound Solvent T g Hyperfine coupling
[8C] constants [G]
quinoid and bis-allyl isomers predicted by DFT and ab initio
calculations.[21, 26]
ACHTUNGRE[C6F6]+ACHTUNGRE[Sb2F11] [b] SO2ClF 20 2.0022 aF = 13.65
ACHTUNGRE[C6HF5]+ACHTUNGRE[AsF6] [c] SO2 70 2.004 aF(1,5)–aF(2,4) = 26.9 The isomeric [C6F6]+ cations were in distinctly different
aF(3) = 4.8, lattice environments, and that electrostatic cation···anion in-
aH = 1.4 teractions caused the relative stabilization of one geometry
2,3,5,6-[C6H2F4]+- SO2 70 2.004 aF = 25.6, or the other. In our analysis, only C F atomic distances
ACHTUNGRE[SbF6] [d] aH = 1.0
shorter than about 326 pm were considered because the van
ACHTUNGRE[C6Cl6]+ACHTUNGRE[Sb2F11] SO2ClF 25 2.010 broad envelope
ACHTUNGRE[C6I6]+ACHTUNGRE[AsF6] , none 25, very broad signal der Waals radii of F and C atoms are 147–150 and 176–
ACHTUNGRE[C6I6]+ACHTUNGRE[SbF6] (solid) 70, 177 pm, respectively. The bis-allyl structure, with four short
196 and two longer C C bonds, should have more positive
[a] All data are from this work unless otherwise noted. [b] Data taken charge localized on the C2 and C3 atoms than on the C1
from reference [26]. [c] Literature values from reference [22]: g = 2.0036, and C4 atoms. In harmony with this prediction, the closest
aF(1,5)–aF(2,4) = 25.8ACHTUNGRE(0.2) G, aF(3) = (4.8  0.2) G. [d] Literature values from
C1···F and C4···F contacts were 325.5(3) ( 2) and
reference [22]: g = 2.0039, aF = (25.8  0.3) G.
318.2(3) pm ( 2), respectively, whilst the closest C2···F con-
tacts were 272.4(3) and 275.7(3) pm, respectively, and the
Selected crystal data and structure-refinement parameters closest C3···F contacts were 272.2(3) and 281.9(3) pm, re-
for the X-ray structures reported herein are listed in spectively, about 50 pm shorter. The situation for the qui-

&2& www.chemeurj.org  2012 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Chem. Eur. J. 0000, 00, 0 – 0

ÝÝ These are not the final page numbers!


Halogenated Benzene Cation Radicals
FULL PAPER
Table 2. Crystal data and structure-refinement parameters.
Compound ACHTUNGRE[C6HF5]+ACHTUNGRE[AsF6] ACHTUNGRE[C6H2F4]+ACHTUNGRE[SbF6] ·2 HF ACHTUNGRE[C6H3F3]+ACHTUNGRE[AsF6] ACHTUNGRE[C6H3F3]+ACHTUNGRE[SbF6 ] ACHTUNGRE[C6Cl6]+ACHTUNGRE[Sb2F11]
formula C6HF11As C12F21H5Sb2 C6F9H3As C6F9H3Sb C6Cl6F11Sb2
Mr 356.99 791.66 321.00 367.83 737.27
crystal system monoclinic triclinic hexagonal hexagonal monoclinic
space group P21/n P1̄ R3̄ R3̄c P21/m
a [pm] 790.74(8) 805.3(2) 973.6(1) 991.66(7) 789.9(2)
b [pm] 2054.7(2) 1020.2(3) 973.6(1) 991.66(7) 1530.2(2)
c [pm] 1170.91(1) 1273.3(4) 1579.3(2) 1605.5(2) 830.3(2)
a [8] 90 99.33(1) 90 90 90
ß [8] 100.461(3) 98.28(1) 90 90 116.80(1)
g [8] 90 93.27(1) 120 120 90
V,  106 [pm3] 1870.9(3) 1018.0(5) 1296.5(3) 1367.3(2) 895.9(3)
Z 8 2 6 6 2
1calcd [Mg m 3] 2.535 2.583 2.467 2.680 2.733
m [mm 1] 3.787 2.841 4.051 3.139 4.006
crystal size [mm] 0.3  0.1  0.1 0.15  0.14  0.05 0.5  0.01  0.01 0.2  0.05  0.05 0.4  0.2  0.01
color yellow green red red brown blue
independent data 3290 10086 443 468 2752
Rint 0.045 0.0251 0.0230 0.0181 0.0277
parameters 336 336 29 30 115
R1, wR2 0.0328, 0.0901 0.026, 0.0543 0.0188, 0.0477 0.0111, 0.0304 0.0237, 0.0513
ACHTUNGRE[C6I6]+ACHTUNGRE[AsF6] ACHTUNGRE[C6I6]+ACHTUNGRE[SbF6] ACHTUNGRE[C6I6]+[HACHTUNGRE(SO3CF3)2] 2,3,5,6-C6H2F4
formula C6F6I6As C6F6I6Sb C8F6HI6O6S2 C6F4H2
Mr 1022.38 1069.21 1132.61 150.08
crystal system monoclinic monoclinic triclinic monoclinic
space group C2/c C2/c P1̄ P21/c
a [pm] 1820.2(14) 1838.9(2) 945.2(7) 448.0(1)
b [pm] 669.6(6) 673.54(7) 986.2(5) 1025.1(2)
c [pm] 1456.2(13) 1484.2(2) 1274.2(9) 630.1(2)
a [8] 90 90 75.39(2) 90
ß [8] 109.36(2) 109.243(5) 89.76(2) 107.99(3)
g [8] 90 90 81.92(3) 90
V,  106 [pm3] 1674(3) 1735.5(3) 1137.3(13) 275.3(1)
Z 4 4 2 2
1calcd [Mg m 3] 4.056 4.092 3.307 1.811
m [mm 1] 13.133 12.303 8.448 0.201
crystal size [mm] 0.15  0.04  0.01 0.3  0.05  0.01 0.3  0.05  0.01 0.2  0.2  0.2
color black black black colorless
independent data 2559 2627 4935 1948
Rint 0.0291 0.0283 0.0387 0.0177
parameters 87 87 256 50
R1, wR2 0.0248, 0.0496 0.0292, 0.0475 0.0321, 0.0772 0.0537, 0.1310

Table 3. B3LYP-predicted relative energies [kJ mol 1] for the [C6X6]+


cation radicals. partially fluorinated benzenes. The impurities, mostly
Cation radical Quinoid Bis-allyl D6h symmetry
C6HF5, were as high as 2 mol %. If a large excess of C6F6
structure structure point was used relative to the oxidant, the resulting product would
ACHTUNGRE[C6F6]+ 0 0.4 12[a] contain [C6HF5]+ as well as [C6F6]+, because hydrogen-con-
ACHTUNGRE[C6Cl6]+ 0 0.3 6.7[a] taining fluorobenzenes have lower IPs than C6F6. This result
ACHTUNGRE[C6Br6]+ 0 0.3 4.2[a] could only be avoided by using C6F6 with a purity of 99.9 %
ACHTUNGRE[C6I6]+ 0 or above, and the molar excess used in oxidation reactions
[a] Jahn–Teller stabilization energy at the intrinsic linearly-interpolated should be as low as practically possible (preferably less than
coordinates. a two-fold excess of C6F6).

ACHTUNGRE[C6HF5]+: When one of the six F atoms in C6F6 was replaced


noid [C6F6] cation in [C6F6] ACHTUNGRE[Sb2F11] was the complete op-
+ +
by a H atom, calculations predicted that the separation of
posite, although the differences were only about 10 pm in- the quinoid and bis-allyl isomer energies increased to about
stead of 50 pm. A more-detailed analysis of the cationic···a- 34 kJ mol 1,[20] with the bis-allyl structure becoming the
nionic interactions is given in the Supporting Information. ground state. The compound [C6HF5]+ACHTUNGRE[AsF6] prepared
One experimental problem that should be mentioned herein was a yellow crystalline solid; its EPR spectrum had
before discussing the other benzene cation radicals is that the expected hyperfine structure and resembled the spec-
commercially available C6F6 contained various amounts of trum reported more than 40 years ago.[22] In crystals of this

Chem. Eur. J. 2012, 00, 0 – 0  2012 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.chemeurj.org &3&
These are not the final page numbers! ÞÞ
K. Seppelt et al.

Table 4. X-ray and B3LYP-predicted average C C bond lengths for sets of real or idealized symmetry-related compound, there were two
bonds in neutral C6X6 and [C6X6]+ cation radicals [pm].[a] unique [C6F5H]+ cations and
X Neutral C6X6, ACHTUNGRE[C6X6]+ quinoid ACHTUNGRE[C6X6]+ bis-allyl-like ACHTUNGRE[C6X6]+ D6h two unique AsF6 anions in the
D6h geometry geometry geometry geometry asymmetric unit, and the struc-
X-ray DFT X-ray DFT X-ray DFT X-ray DFT
tures of the two cation radicals
F 137.8(2) 138.9 136.8(4)  2, 137.0  2, 137.7(4)  4’ 138.8  4,
were very similar: both cations
141.0(3)  4[b,c] 142.7  4 143.4(3)  2[b,c] 144.8  2
Cl 139.5(5)[d] 139.8 138.2  2, 140.2(3)  4, 139.9  4, had the C2v bis-allyl geometry
144.8  4 144.1(5)  2[b] 144.7  2 (Figure 2). As was the case with
Br 140.1(5)[d] 140.0 138.8  2, 140.0  4, the [C6F6]+ ions, the long and
142.8  4 144.3  2 short C C atomic distances
I 139.9(7)[e] 140.8 139.6(5)  6[f] 139.7  6
were clearly statistically differ-
[a] All data are from this work unless otherwise noted (T = 140 8C). For the X-ray structures, the standard ent, even though the standard
error shown is the standard error for an individual C C distance, not the standard deviation of the average.
[b] Reference [26] (T = 140 8C). [c] [Sb2F11] salt. [d] Reference [29] (T = 173 8C). [e] Reference [37] (T =
errors were higher: the differ-
173 8C). [f] [SbF6] salt. ence between the average of
one set of short C C bond
lengths, 136.8(7) pm, and the
Table 5. X-ray and B3LYP-predicted average C C bond lengths for sets of real or idealized symmetry-related
bonds in [C6HnFm]+ cation radicals [pm].[a] average of the long bond
Compound ACHTUNGRE[C6HnFm]+ quinoid ACHTUNGRE[C6HnFm]+ bis-allyl-like ACHTUNGRE[C6HnFm]+ D3h
lengths, 143.7(7) pm, was about
geometry geometry geometry 10 times the standard error for
X-ray DFT X-ray DFT X-ray an individual C C bond length.
ACHTUNGRE[C6HF5]+ACHTUNGRE[AsF6] 137.1(7)  2, 137.9  2, Furthermore, as observed
143.7(7)  2, 144.8  2, above, the DFT-predicted C C
136.8(7)  2 138.9  2 bond lengths closely matched
2,3,5,6-[C6H2F4]+ACHTUNGRE[SbF6] ·0.5 HF 137.2(3)  4, 137.9  4
the X-ray-determined bond
143.9(3)  2[b] 144.9  2
2,4,6-[C6H3F3]+ACHTUNGRE[SbF6] 142.6  2 144.2  2 140.7(2)  6[c] lengths (Table 5).
141.6  2 138.0  2
136.3  2 138.4  2 2,3,5,6-[C6H2F4]+: Tetrafluoro-
3,4,5,6-[C6H2F4]+ 142.0  2
benzene 2,3,5,6-C6H2F4 was oxi-
136.2  2,
142.1  2 dized to produce crystals of
2,4,5,6-[C6H2F4]+ 141.4  2 both 2,3,5,6-[C6H2F4]+-
136.2  2, ACHTUNGRE[SbF6] ·0.25 HF and 2,3,5,6-
143.5  2 [C6H2F4]+ACHTUNGRE[Sb2F11] . The EPR
[a] All data are from this work (T = 140 8C). For the X-ray structures, the standard error shown is the stan- spectrum of the 2,3,5,6-
dard error for an individual C C distance, not the standard deviation of the average. [b] The three correspond- [C6H2F4]+ cation radical was
ing distances for neutral 2,3,5,6-C6H2F4 (T = 140 8C) are 138.2(1), 138.3(1), and 138.3(1) pm. [c] The B3LYP-
predicted difference in energies between the quinoid and bis-allyl (ground state) structures of 2,4,6-[C6H3F3]+
a straightforward quintet (aF =
cation is 1.2 kJ mol 1. 25.6 G) of triplets (aH = 1.0 G).
A single crystal of 2,3,5,6-
[C6H2F4]+ACHTUNGRE[SbF6] gave the more
precise X-ray structure, so only these data will be discussed

Figure 1. The two [C6F6]+ cation radicals in the crystal of Figure 2. The [C6HF5]+ ion in [C6HF5]+ [AsF6] ; ellipsoids set at 50 %
[C6F6]+ [Sb2F11] , ellipsoids set at 50 % probability (see reference [26]). probability, except for the H atom, which is shown as a sphere of arbitra-
Left: the quinoid cation. Right: the bis-allylic cation. Only the unique C ry size. The two distances [pm] given for each C C bond refer to the two
C bond lengths [pm] are shown. The average C C distance in neutral crystallographically independent cations in the asymmetric unit. The
C6F6 is 137.8(2) pm (average of nine values; see the Supporting Informa- numbers in italics are the DFT-predicted C C bond lengths for the gas-
tion). phase cation radical.

&4& www.chemeurj.org  2012 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Chem. Eur. J. 0000, 00, 0 – 0

ÝÝ These are not the final page numbers!


Halogenated Benzene Cation Radicals
FULL PAPER
herein (although the cations in both structures had very sim-
ilar geometries).
There were four independent 2,3,5,6-[C6H2F4]+ cations,
four [SbF6] anions, and two HF molecules in the asymmet-
ric unit. All of the cations possessed the bis-allyl geometry
(Figure 3), which agreed with the predicted geometry
(Table 5, also see the Supporting Information). The C C
bond lengths in the cations were compared with neutral
2,3,5,6-C6H2F4, which was readily obtained as single crystals
owing to its relatively high melting point. The C C bond
lengths were all statistically the same in the structure of this
neutral tetrafluorobenzene (138.23(9)–138.33(9) pm; see the
Supporting Information, Figure SI-4). Figure 4. Left: The formula unit in the structure of 2,4,6-
[C6H3F3]+ [SbF6] ; ellipsoids set at 50 % probability, except for the H
atoms, which are shown as spheres of arbitrary size. The rigorously
planar cation is in the plane of the page and the SbF6 anion is below the
plane of the page. Only the unique C C, C H, and C F distances are
shown. The corresponding distances in the structure of 2,4,6-[C6H3F3]+-
ACHTUNGRE[AsF6] are 140.5(2), 84(4), and 131.3(3) pm, respectively. Right: The cat-
ionbanion packing along the threefold crystallographic axis in 2,4,6-
[C6H3F3]+ACHTUNGRE[SbF6] . All of the C1···F2 distances shown are 292.7(1) pm. Se-
lected interatomic distance [pm] and angles [8]:Sb–F 187.90(8);
F2···C1···F2’ 173.88(8), F-Sb-F 90.41(4), 89.59(4), and 180.00(4).

ally symmetric and efficient cation···anion packing in the


+ crystal. Figure 4 shows the columnar arrangement of alter-
Figure 3. The 2,3,5,6-[C6H2F4] cation radical in
[C6H2F4]+ [SbF6] ·0.25 HF; ellipsoids set at 50 % probability, except for nating cations and anions along the hexagonal axis of the
the H atoms, which are shown as spheres of arbitrary size. The first four rhombohedral crystal. There were relatively short contacts
numbers are the C C bond lengths [pm] for the four crystallographically (292.7(1) pm) between the three benzene C(F) atoms and
independent cations in the lattice. Only the unique C C bond lengths are
the anion F atoms; this was a sensible result because the
shown. The first number in italics is the DFT-predicted C C distance for
the gas-phase cation radical; the second number in italics is the DFT-pre- C(F) atoms were expected to carry most of the positive
dicted C C distance for the neutral precursor 2,3,5,6-C6H2F4 (also see the charge of the cation. The substantial charge-neutralization
Supporting Information). and concomitant large lattice energy caused by the efficient
ion packing sterically enforced the three-fold symmetry on
the 2,4,6-[C6H3F3]+ ion, despite the electronically-preferred
2,4,6-[C6H3F3]+: Owing to the D3h symmetry of the parent distorted geometry. This result also explains why only the
benzene, 2,4,6-C6H3F3, the corresponding cation should ex- SbF6 compound could be isolated: the low solubility of
hibit a Jahn–Teller distortion. Again, the two possible iso- 2,4,6-[C6H3F3]+ACHTUNGRE[SbF6] drove the equilibrium [Sb2F11] Q
mers, quinoid and bis-allyl, should be very close in energy, [SbF6] +SbF5 to the right, even when the oxidation with
with the quinoid form possibly being the minimum [O2]+ACHTUNGRE[Sb2F11] was carried out with a large excess of SbF5.
(Table 3). Oxidation of 2,4,6-C6H3F3 with [O2]+ACHTUNGRE[Sb2F11] and We were not able to isolate a salt of the 2,4,6-[C6H3F3]+
[O2]+ACHTUNGRE[AsF6] in liquid HF produced 2,4,6-[C6H3F3]+SbF6 cation with a lower-symmetric anion, either by starting with
and 2,4,6-[C6H3F3]+ACHTUNGRE[AsF6] , respectively, as ruby-red, an alternative oxidant or through an anion-exchange reac-
needle-shaped crystals in almost quantitative yield. Interest- tion following the oxidation.
ingly, the use of [O2]+ACHTUNGRE[Sb2F11] as the oxidant yielded the Our current interpretation of the structure of this com-
[SbF6] salt, even in the presence of a large excess of SbF5. pound is that the electrostatic interaction provides enough
Both salts resulted in very precise crystal-structure deter- energy to hold the cation in the observed symmetric geome-
minations. The crystals were isomorphous, and the space try. Our calculations predicted that the energy needed to
group R3̄c resulted in the cations and anions being in special overcome the Jahn–Teller distortion was only about
positions. This structure required rigorous D3d symmetry for 12 kJ mol 1.
the anions and, more importantly, rigorous D3h symmetry
for the 2,4,6-[C6H3F3]+ cation radical. Although distorted ACHTUNGRE[C6Cl6]+ and [C6Br6]+: As expected, C6Cl6 and C6Br6 were
cations that were three-fold disordered could not be exclud- easier to oxidize than the fluorinated benzenes (IP: 9.13(3)
ed, the small and relatively symmetric thermal displacement and 8.80 eV, respectively; see the Supporting Information,
parameters (at 140 8C) did not give any indication of disor- Table SI-1). Neat SbF5 alone could also act as the oxidant
der. We believe that the lack of an observable Jahn–Teller for these hexahalobenzenes. Recrystallization from SO2FCl
distortion in the cation, despite the prediction that the afforded crystals of [C6Cl6]+ACHTUNGRE[Sb2F11] and [C6Br6]+ACHTUNGRE[Sb2F11] .
cation should be distorted, could be explained by the unusu- The crystal structure of [C6Cl6]+ACHTUNGRE[Sb2F11] was precise enough

Chem. Eur. J. 2012, 00, 0 – 0  2012 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.chemeurj.org &5&
These are not the final page numbers! ÞÞ
K. Seppelt et al.

to draw definite conclusions about the geometry of the Hypothetically, if the [C6Cl6]+ cation radical had the quinoid
[C6Cl6]+ cation radical. However, the crystallographic char- geometry, we would expect the C2 atoms to have closer
acteristics of [C6Br6]+ACHTUNGRE[Sb2F11] were poor. The [C6Br6]+ and/or a greater number of contacts with the F atoms of the
cation seemed to have a quinoid structure, as predicted by anions relative to the C1 and C3 atoms.
computation, but better structural data are needed.
However, the [C6Cl6]+ cation had a bis-allyl geometry CHTUNGRE[A C6I6]+: C6I6 was the benzene with the lowest ionization po-
(Figure 5), with two long C C bonds (144.5(5) and tential studied herein (IP = 7.9 eV; see the Supporting Infor-
mation, Table SI-1). Several attempts to oxidize C6I6 have
been reported, and black products were always observed in
H2O2/CF3SO3H solutions. These products have been de-
scribed as [C6I6]2+ salts.[31, 32] The crystal structure of a com-
pound with the formula [C6I62+]ACHTUNGRE[(CF3SO3 )2] was described
in a PhD dissertation,[33] but, to the best of our knowledge,
such a structure has not been published elsewhere. Herein,
we oxidized C6I6 by using SbF5 or AsF5 in HF, which indeed
resulted in such black precipitates. However, single crystals
were obtained with the compositions [C6I6]+ACHTUNGRE[SbF6] and
[C6I6]+ACHTUNGRE[AsF6] .[34] These two salts were isomorphous.
The [C6I6]+ cation radicals in both structures had undis-
torted hexagonal geometries (Figure 6). Because the packing
arrangements of the ions were quite irregular (unlike the sit-
uation in the 2,4,6-[C6H3F3]+ salts), the symmetric geome-
tries observed for the [C6I6]+ cation radicals in [C6I6]+
[SbF6] and [C6I6]+ [AsF6] were quite clearly not a result of
differential cation···anion interactions. In fact, a symmetric
geometry was predicted by our calculations. The HOMO of
C6I6 was a non-bonding A1g iodine-centered orbital that was
Figure 5. The structure of [C6Cl6]+ [Sb2F11] , which shows two symmetry- represented by a combination of the six iodine p-orbitals
related anions and one cation radical; ellipsoids set at 50 % probability.
Selected interatomic distances [pm]: C1–C2 139.9(3), C1–C1A 144.5(5),
lying in the molecular plane in an antibonding manner.
C2–C3 140.5(3), C3–C3A 144.1(5), C1···F4 317.3(3), C1···F5B 300.4(3), Therefore, one-electron oxidation of C6I6 removed an elec-
C2···F5B 312.5(3), C3···F3 326.6(3), C3···F2B 319.1(3). tron largely from the I atoms, not from the C atoms (i.e.,
not from a degenerate C C p orbital), and should not result
144.1(5) pm) and four shorter C C bonds (in the range in a Jahn–Teller distortion of the molecule (unlike the other
139.9(3)–140.5(3) pm). The three unique C C distances in halogens, the ionization potential of an I atom, 10.45 eV, is
neutral C6Cl6 were statistically the same and ranged from lower, not higher, than that of a C atom, 11.26 eV). Further-
138.6(5)–140.0(5) pm at 173 8C[29] and from 139.1(3)– more, oxidation of the [C6I6]+ cation radical should result in
[30]
139.5(3) pm at 24(1) 8C. Although this result contradicted a [C6I6]2+ dication that also has a hexagonal geometry. It has
the theoretical prediction that the ground state of the been suggested that removal of the two A1g electrons from
[C6Cl6]+ cation should adopt the quinoid geometry, the cal- C6I6 would result in a s-aromatic ring-current in the [C6I6]2+
culated difference in energy was only 0.24 kJ mol 1 between dication,[31, 32, 35, 36] in contrast to the situation for [C6Cl6]2+.[37]
the quinoid ground state and
the bis-allyl transition state.
There were two C···F(Sb) inter-
actions for the C1 atom and
two for the C3 atom, which
ranged from 300.4(3) to
326.6(3) pm, but only one
C2···F(Sb) interaction
(312.5(3) pm), as expected for
a bis-allyl structure. In fact, it
can be argued that the C2···F5B
contact was really a conse-
quence of the proximity of the
F5B atom to the C1 atom
Figure 6. Left: The [C6I6]+ cation radical in [C6I6]+ [AsF6] and [C6I6]+ [SbF6] , which are isostructural; ellip-
(C1···F5b = 300.4(3) pm), which soids set at 50 % probability. The top number refers to the [AsF6] salt and the bottom number to the [SbF6]
required that the F5B atom salt. Only the unique C C distances [pm] are shown. Right: The crystal structure of [C6I6]+ [HACHTUNGRE(SO3CF3)2] .
must be close to the C2 atom. The anion has a short O(H)···O distance of 243.9 pm.

&6& www.chemeurj.org  2012 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Chem. Eur. J. 0000, 00, 0 – 0

ÝÝ These are not the final page numbers!


Halogenated Benzene Cation Radicals
FULL PAPER
The average intramolecular I···I distance for the [C6I6]+ zenes occasionally resulted in the formation of salts of the
cation radical (347.8 pm) was only marginally shorter than perhalogenated benzenium cations [C6Cl7]+, [C6Cl5F2]+, and
the corresponding average distance in neutral C6I6 [C6Br5F2]+. These compounds will be reported in a future
[37]
(351.3 pm). This was a sensible result, because the as- report.
sumed half-bond between the I atoms in [C6I6]+, spread as
over six intramolecular I···I interactions, would not be ex- Computational results: Perhalogenated benzenes: Our DFT-
pected to result in a significant amount of shortening. Fur- predicted and X-ray structural parameters for the neutral
thermore, the shortening was not due to differences in the benzenes C6X6 (D6h symmetry; X = F, Cl, Br, I) and the
displacements of the I atoms from the C6 planes of C6I6 and cation radicals [C6X6]+ are listed in Table 4. At the level of
+
[C6I6] : in both cases, the mean displacement was only about theory used in this work, the C X bond lengths increased
4 pm. from 1.331  in C6F6 to 2.123  in C6I6, as did the intramo-
We believed that the compound previously reported to be lecular X···X distance, from 2.270  to 3.531 . Interesting-
[C6I6]2+ACHTUNGRE[(CF3SO3 )2] was probably [C6I6]+[HACHTUNGRE(CF3SO3)2] . ly, the aromatic-ring C C bond lengths also increased as the
Dissolution of [C6I6]+ACHTUNGRE[AsF6] in CF3SO3H resulted in the size of the halogen atoms increased, from 1.389  in C6F6 to
formation of small amounts of crystalline [C6I6+][H- 1.408  in C6I6. The two-fold degenerate HOMO in C6F6
ACHTUNGRE(CF3SO3)2] , which we analyzed by X-ray crystallography. was a pair of C C p orbitals, labeled as B (a bis-allyl-like or-
Our compound and the compound previously reported as bital) and Q (a quinoid-like orbital) in Figure 7. The bis-
[C6I6]2+ACHTUNGRE[(SO3CF3 )2][33] were
crystallographically identical in
all respects. As in the structures
of [C6I6]+ACHTUNGRE[SbF6] and [C6I6]+-
ACHTUNGRE[AsF6] , the monovalent cation
radical [C6I6]+ in [C6I6+][H-
ACHTUNGRE(CF3SO3)2] had a symmetric
hexagonal geometry. The over-
looked hydrogen atom in the
previously reported crystal
structure was not only detected
in this work by a difference
Fourier analysis, its presence
was also obvious because of the
short O···O distance (243.9 pm),
which is the hallmark of
a strong O H···O hydrogen
bond.[38] Furthermore the two
S O bonds that participated in Figure 7. The HOMO energies for C6X6 (X = F, Cl, Br, and I) are shown in a partial MO diagram. B = bis-allyl
the hydrogen bond were elon- orbital; Q = quinoid orbital; XX3* = totally symmetric antibonding in-plane halogen-atom p-orbital with three
gated by 5–8 pm relative to the nodal planes. Orbital plots for C6F6 are shown as an example (B3LYP/tzvpp).
other S O bonds.
The absence of an observable
EPR spectrum was previously cited as evidence for the exis- allyl orbital had a perpendicular nodal plane that bisected
tence of the diamagnetic [C6I6]2+ dication.[33] However, in opposite C C bonds, whilst the quinoid orbital had a perpen-
our hands, all of the C6I6+ salts exhibited very broad EPR dicular nodal plane that included two para-C atoms. The
signals at 196 8C. Taken together, our crystallographic and third C C p orbital was HOMO-1. Below the energy of
spectroscopic results led us to conclude that the [C6I6]2+ di- HOMO-1 in C6F6 were six orbitals that were composed of
cation remains an unknown species. Nevertheless, if it is linear combinations of in-plane X-atom p orbitals (X = F).
ever isolated, it will probably be found to have a symmetric One of these six orbitals (XX3*, Figure 7) was important be-
hexagonal geometry like C6I6 and [C6I6]+, as predicted by cause its energy was especially sensitive to the nature of the
calculations. halogen atom. The XX3* energy was 0.175 below the B/Q
energy for C6F6, but this difference decreased to 0.039 for
Halogenated benzenium ion salts: Salts of benzenium cat- C6Cl6 and to 0.003 for C6Br6. For C6I6, XX3* was the
ions are intermediates in electrophilic aromatic substitution HOMO and the degenerate B/Q orbital was HOMO-1.
reactions and therefore have been studied extensively. In The molecular orbital diagrams indicated that Jahn–
many cases, they can be isolated, particularly if electron-do- Teller-distorted geometries should be found for [C6X6]+ with
nating substituents are present.[39] The parent cation [C6H7]+ X = F, Cl, and Br but not for X = I. Two distorted cations,
was recently isolated by using a weakly-coordinating carbor- a compressed one with the quinoid orbital as a singly occu-
ane anion.[39] Our oxidation reactions of halogenated ben- pied orbital and an elongated one with the bis-allyl orbital

Chem. Eur. J. 2012, 00, 0 – 0  2012 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.chemeurj.org &7&
These are not the final page numbers! ÞÞ
K. Seppelt et al.

as a singly occupied orbital, were found for [C6X6]+, with


X = F, Cl, and Br. The compressed (quinoid) cation geome-
tries were classified as minima, with the elongated (bis-allyl)
geometries as transition states. The [C6I6]+ cation retained
the D6h symmetry of neutral C6I6. The behavior of the orbi-
tals was as expected from the molecular orbital diagram of
the neutral molecule. The difference between the quinoid
and bis-allylic orbitals was about 0.13–0.18 for the cations
(X = F, Cl, Br), and the XX3* orbitals became closer in
energy to the quinoid and bis-allylic orbitals. Significantly,
the barrier between the distorted ground-state geometries
and the hexagonal geometry decreased in the order
[C6F6]+ > [C6Cl6]+ > [C6Br6]+.
Linearly interpolated intrinsic coordinates (LIICs) for the
Figure 8. The quinoid- and bis-allyl-distorted cation-radical structures are
cation radicals [C6F6]+, [C6Cl6]+, and [C6Br6]+ were generat- connected by two paths. The path through the D6h symmetry point can
ed to calculate the Jahn–Teller stabilization energy, eJT, as be described by calculating the LIICs; the other path, which leads from
the difference EACHTUNGRE(D6h) EACHTUNGRE(D2h), where EACHTUNGRE(D2h) was the energy the bis-allyl cation as a transition state to the minimum-energy quinoid
of the bis-allyl cation radical. For C6F6+ this difference was cation, can be calculated by using IRCs. The combined paths can be used
to construct a potential energy surface for the system, which has the ap-
about 12 kJ mol 1, which decreased to 6.7 kJ mol 1 for
pearance of a “Mexican hat” (see reference [26]).
[C6Cl6]+ and to 5.4 kJ mol 1 for [C6Br6]+. The bis-allyl and
quinoid structures were nearly isoenergetic. In [C6F6]+, their
energy difference was only 0.4 kJ mol 1, and the difference because the maximum difference in the C C bond lengths
was even smaller for [C6Cl6]+ and [C6Br6]+ (Table 3). These was smaller than 1 pm. Inspection of the molecular orbital
values were certainly within the uncertainty of the computa- diagrams for these molecules showed that the bis-allyl and
tional methods used, but the calculated trend was independ- quinoid p orbitals were no longer degenerate. In C6HF5 and
ent of the basis set and the method used. A calculation 2,3,5,6-C6H2F4, the quinoid orbital was the HOMO, and so
using intrinsic reaction coordinates (IRCs) offered a possible removing an electron would lead to a bis-allyl-distorted
reaction path from the bis-allylic cation to the quinoid cation radical. In 2,4,5,6-C6H2F4 and 3,4,5,6-C6H2F4, the bis-
cation. In both paths, the LIICs and the IRCs connected the allyl orbital was the HOMO and a quinoid-distorted cation
two distorted structures. These are the two typical paths for radical would be expected (Figure 9).
the interconversion of geometries in such systems,[19] and
they can be combined to calculate a potential energy surface
(Figure 8).
The Jahn–Teller distortions in the [C6X6]+ cation radicals
were quantified geometrically by comparing the difference
between the elongated and compressed C C bond lengths;
we found that this difference was predicted to decrease
from 6.0 pm for both types of [C6F6]+ cations to 4.8 and
4.6 pm for the bis-allyl and quinoid [C6Cl6]+ cations, respec-
tively, to 4.3 and 4.0 pm for the bis-allyl and quinoid
[C6Br6]+ cations, respectively. Experimentally, the observed
differences were 6.0(5) and 5.4(4) pm for the bis-allyl
[C6F6]+ cation in [C6F6]+ACHTUNGRE[Sb2F11] and 4.6(6) and 3.6(6) pm
for the bis-allyl [C6Cl6]+ cation in [C6Cl6]+ACHTUNGRE[Sb2F11] .
The average Mulliken charge on the F atoms in [C6F6]+
was predicted to be slightly negative ( 0.1 e). In contrast, it
was slightly positive (+0.1 e) on the Cl and Br atoms in
Figure 9. A partial molecular orbital diagram for fluorinated benzenes
[C6Cl6]+ and [C6Br6]+ and +0.25 e on the I atoms in [C6I6]+. showing the high-energy occupied orbitals, including the bis-allyl (B) and
Both the Mulliken charge and the occupancy of the orbitals quionid (Q) p orbitals. Orbital plots for C6F6 are shown as an example
indicated a significant positive charge on the I atoms in the (B3LYP/tzvpp).
cation radical.

Partially fluorinated benzene derivatives: The molecular The explanation of these differences is as follows: In C6F6,
symmetry of the partially-fluorinated neutral molecules the bis-allyl and quinoid p orbitals contained an antibonding
C6HF5, 2,3,5,6-C6H2F4, and 2,4,5,6-C6H2F4 was D2h, and for contribution from the in-plane F atom p-orbitals, thus rais-
3,4,5,6-C6H2F4 it was C2v. Nevertheless, the hexagonal C C ing their energy. For the bis-allyl p orbital, two large and
framework in all four molecules had effective D6h symmetry four small out-of-plane F atom p-orbital contributions were

&8& www.chemeurj.org  2012 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Chem. Eur. J. 0000, 00, 0 – 0

ÝÝ These are not the final page numbers!


Halogenated Benzene Cation Radicals
FULL PAPER
present, whilst for the quinoid p orbital, four large out-of- [C6F6]+ [Os2F11] and [C6F6]+ [Sb2F11] were prepared according to litera-
plane contributions were present. In C6HF5 and 2,3,5,6- ture procedures.[26]
C6H2F4, one and two of the large out-of-plane contributions Physical methods: X-band EPR spectra were recorded on a Bruker ER
200 D-SRC spectrometer equipped with an ER 4111 variable-tempera-
in C6F6 were missing due to the presence of the H atoms, re-
ture unit that allowed for low-temperature measurements. Except for the
spectively. This result lowered the energy of the bis-allyl or- [C6I6]+ salts, all samples were freshly prepared in either SO2 or SO2ClF
bital, thereby making the quinoid orbital the HOMO. In (Table 1) and sealed in 4 mm O.D. PFA tubes. Samples of [C6I6]+ [AsF6]
3,4,5,6-C6H2F4 and 2,4,5,6-C6H2F4, both orbitals were stabi- and [C6I6]+ [SbF6] were weighed into PFA tubes, then anhydrous
lized as follows: for the bis-allylic orbital, two small anti- CF3SO3H was added, and the tubes were sealed under an Ar atmosphere
(PFA = poly(perfluorether-tetrafluorethylene)). Single crystals suitable
bonding F atom p-orbital contributions were missing, and
for X-ray diffraction were mounted on a Bruker Smart 2000 diffractome-
for the quinoid orbital two large antibonding F atom p-orbi- ter (MoKa radiation) under oxygen- and moisture-free conditions at tem-
tal contributions were missing. Thus, the quinoid p orbital peratures below 100 8C by using a special device of local design. All
was stabilized to a greater extent, and the bis-allyl orbital data collections were performed at 140 8C. After semiempirical absorp-
became the HOMO. tion corrections, the structures were solved and refined by using the pro-
gram SHELXL.[44] Non-hydrogen atoms were refined anisotropically; H
After removing an electron from the neutral molecules, atoms were refined isotropically. Relevant data collection and refinement
structural optimizations were performed, starting with the parameters are listed in Table 2.
same symmetry. In each case, a distorted cation was found. Synthesis of new compounds: General procedures: Moisture-sensitive
As expected, [C6HF5]+ and 2,3,5,6-[C6H2F4]+ had bis-allyl nonvolatile materials were handled in a glovebox in an inert atmosphere
geometries and 2,4,5,6-[C6H2F4]+ and 3,4,5,6-[C6H2F4]+ had of less than 0.1 ppm H2O and O2. Volatile materials were handled in
quinoid geometries. The difference for the bis-allylic and glass or stainless steel vacuum lines as appropriate. The oxidation reac-
tions were generally carried out in 8 mm PFA tubes that were connected
quinoid orbitals increased from the neutral compounds to a vacuum line. Weighed portions of the nonvolatile oxidants were
(average: 0.016) to the cation radicals (average: 0.197), but transferred into the tube in the glovebox, after which the solvent and the
the order was the same. These results were in full agreement benzene reagent were added by using the vacuum line. The oxidation re-
with previous calculations reported for such systems.[20] Fol- actions frequently occurred rapidly even at low temperatures, and were
lowing ionization, the molecules were predicted to exhibit signaled by the formation of an intense color. The cation-radical-salt
products often decomposed at 25 8C, with the concomitant loss of color.
distorted bis-allyl geometries, with maximum differences in
ACHTUNGRE[C6HF5]+ [AsF6] : A sample of C6HF5 (600 mg, 3.57 mmol) was cooled to
C C distance of 6.4–7.3 pm. The differences in experimental 48 8C, the melting point of this compound. Small portions of [O2]+-
C C bond lengths for the [C6HF5]+ and 2,3,5,6-[C6H2F4]+ ACHTUNGRE[AsF6] (total 100 mg, 0.453 mmol) were added under an Ar atmosphere;
cations agreed with the calculated values to within ( 2) and the mixture was shaken carefully after each portion (about 5 mg) was
( 1) pm, respectively (these two cation radicals did have added. The resulting green suspension was warmed to 35 8C and all of
bis-allyl geometries in the solid state). the volatile compounds were removed under vacuum. A yellow powder
was obtained in 70 mg (43 % based on [O2]+ACHTUNGRE[AsF6] ). The dry product
Finally, all of our attempts to find other relatively stable decomposed within minutes at 25 8C. Recrystallization from anhydrous
distorted geometries for the cation radicals were unsuccess- HF (4 mL) at 78 8C afforded large yellow needles after 24 h.
ful, including LIICs between other assumed geometries and 2,3,5,6-[C6H2F4]+ ACHTUNGRE[SbF6] : A mixture of anhydrous C6F14 (4 mL) and
for geometries distorted along symmetry-reducing coordi- 2,3,5,6-C6H2F4 (60 mg, 0.4 mmol) was cooled to 78 8C and small por-
nates. tions of [O2]+ACHTUNGRE[Sb2F12] (total 200 mg, 0.413 mmol) were added under an
Ar atmosphere; the mixture was shaken carefully after each portion
(about 5 mg) was added. The resulting yellow–green suspension was
warmed to 45 8C and all of the volatile compounds were removed
under vacuum. The product, as a yellow powder, decomposed immediate-
Experimental Section ly at 25 8C and was recrystallized from a green solution in anhydrous HF
(2 mL) as green crystals after 1 week at 78 8C.
Caution: Handling anhydrous HF, dioxygenyl salts, SbF5, and AsF5 re-
2,4,6-[C6H3F3]+ ACHTUNGRE[AsF6] : [O2]+ACHTUNGRE[AsF6] (100 mg, 0.453 mmol) and AsF5
quires eye and skin protection.
(about 2 mL, 4.3 g, 25 mmol) were mixed in a PFA tube. The volatile
Reagents and solvents: The following reagents were purchased from the compound 2,4,6-C6H3F3 (54 mg, 0.41 mmol) was added to this mixture at
indicated vendor: C6HF5, 2,4,6-C6H3F3 and 2,3,5,6-C6H2F4 (ABCR 196 8C. Warming the mixture to 25 8C afforded a clear, wine-red solu-
GmbH), C6Cl6 and C6Br6 (Sigma–Aldrich Chemie GmbH), elemental As tion (note that the compound did not decompose rapidly at 25 8C).
(ABCR GmbH), SO2Cl2 (Merck Schuchardt), SbF5 (Fluorochem Ltd.), Slowly cooling this solution to 78 8C yielded red–brown crystals. Anhy-
NH4F (E. Merck), KI (Acros Organics), HIO4 (H5IO6 ; E. Merck), drous HF and SO2ClF were also used as solvents with similar results.
CF3COOH (Merck Schuchardt), CF3SO3H (Merck Schuchardt; distilled
2,4,6-[C6H3F3]+ ACHTUNGRE[SbF6] : [O2]+ACHTUNGRE[Sb2F11] (100 mg, 0.206 mmol) was dis-
from P2O5 (Sigma Aldrich GmbH)). The five benzenes were checked for
purity by NMR spectroscopy and were redistilled or resublimed as neces- solved in anhydrous HF (2 mL) at 25 8C in a PFA tube. The mixture was
sary. C6I6 was prepared by treating benzene with periodic acid in the cooled to 196 8C and 2,4,6-C6H3F3 (26 mg, 0.20 mmol) was added by
presence of potassium iodide.[40] Arsenic pentafluoride was prepared by vacuum transfer. The mixture was slowly warmed to 25 8C, and produced
passing F2 gas over elemental arsenic.[41] [O2]+ [AsF6] and [O2]+ [Sb2F11] the same wine-red solution as described above. Slow cooling to 78 8C
were prepared from O2, F2, and either AsF5 or SbF5 by irradiation with yielded dark-brown needle-shaped crystals.
UV light.[42] Solvent SO2ClF was prepared by treating a mixture of ACHTUNGRE[C6Cl6]+ ACHTUNGRE[Sb2F11] : C6Cl6 (50 mg, 0.18 mmol) was dissolved in SbF5
SO2Cl2 and NH4F with CF3COOH.[43] Perfluorohexane (ABCR GmbH) (100 mg, 0.461 mmol) at 25 8C in a PFA tube, thereby forming a dark-
was dried and purified by treatment with KF followed by blue solution after 15–30 min. The solution was frozen at 196 8C and an-
[XeF]+ [Sb2F11] . Anhydrous HF was vacuum-distilled into a stainless hydrous HF (2 mL) was added by vacuum transfer. The mixture was
steel cylinder containing BiF5 to remove any residual H2O. SbF5 was slowly warmed to 0 8C, and afforded a clear, dark-blue solution. Slow
vacuum-distilled twice by using a glass vacuum line with a 30 8C trap. cooling of this solution to 78 8C resulted in the formation of dark-blue
The resulting liquid was clear, colorless, and highly viscous. crystals.

Chem. Eur. J. 2012, 00, 0 – 0  2012 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.chemeurj.org &9&
These are not the final page numbers! ÞÞ
K. Seppelt et al.

ACHTUNGRE[C6Br6]+ ACHTUNGRE[Sb2F11] : C6Br6 (50 mg, 0.091 mmol) was dissolved in SbF5 Acknowledgements
(100 mg, 0.461 mmol) at 25 8C in a PFA tube, thereby forming a dark-
green solution after 2–5 min. The solution was frozen at 196 8C and This work was made possible by the Deutsche Forschungsgemeinschaft
SO2ClF (2 mL) was added by vacuum transfer. The mixture was slowly (Graduiertenkolleg “Fluorine as a Key Element”), the Fonds der Chemi-
warmed to 0 8C and afforded a clear, dark-green solution. Slow cooling schen Industrie, and a DFG grant (to M.K.). Research internship stu-
of this solution to 78 8C resulted in the formation of black crystals. a = dents C. Noufele and K. Schulz performed preliminary calculations.
802.9(5), b = 855.3(5), c = 874.0(5) pm; a = 63.12(1), b = 76.13(1), g =
63.088; V = 476.9(5)  106 pm3 ; Z = 1; disordered [Sb2F11] anion; R1 =
0.10; wR2 = 0.3202.
ACHTUNGRE[C6I6]+ ACHTUNGRE[SbF6] : SbF5 (0.42 g, 1.94 mmol), C6I6 (0.5 g, 0.60 mmol), and an- [1] I. D. Clark, D. C. Frost, J. Am. Chem. Soc. 1967, 89, 244 – 247.
hydrous HF (2.0 mL) were mixed in a PFA tube at 25 8C until a deep- [2] R. Bralsford, P. V. Harris, W. C. Price, Proc. R. Soc. London Ser. A
blue precipitate had formed. All volatile compounds were removed 1960, 258, 459 – 469.
under vacuum to afford a deep-blue powder in quantitative yield. A por- [3] N. Sato, K. Seki, H. Inokuchi, J. Chem. Soc. Faraday Trans. 1981, 77,
tion (20 mg) of the powder was dissolved in anhydrous CF3SO3H (2 mL) 1621 – 1633.
and heated to 200 8C for 5 min. Slow cooling of the hot solution to 25 8C [4] N. Bartlett, D. H. Lohmann, Proc. Chem. Soc. 1962, 115 – 116.
resulted in the formation of deep-blue needle-shaped crystals. The com- [5] S. Seidel, K. Seppelt, Angew. Chem. 2000, 112, 4072 – 4074; Angew.
pound was stable at 25 8C under an inert atmosphere, but it decomposed Chem. Int. Ed. 2000, 39, 3923 – 3925.
within 10 min when exposed to air. [6] A. J. Edwards, G. R. Jones, R. J. C. Sills, J. Chem. Soc. Chem.
ACHTUNGRE[C6I6]+ ACHTUNGRE[AsF6] : AsF5 (4.00 g, 23.5 mmol), C6I6 (1.00 g, 1.20 mmol), and Commun. 1968, 1527 – 1528.
anhydrous HF (1.3 mL) were mixed in a PFA tube at 78 8C until [7] C. G. Davies, R. J. Gillespie, R. J. Ireland, J. M. Sowa, Can. J. Chem.
a deep-blue precipitate had formed. All volatile compounds were re- 1974, 52, 2048 – 2052.
moved under vacuum to afford a deep-blue powder in quantitative yield. [8] R. Lindner, H. Sekiya, B. Beyl, K. Mller-Dethlefs, Angew. Chem.
A portion (20 mg) of the powder was dissolved in anhydrous CF3SO3H 1993, 105, 631 – 634; Angew. Chem. Int. Ed. Engl. 1993, 32, 603 – 606.
(2 mL) and heated to 200 8C for 5 min. Slow cooling of the hot solution [9] M. Iwasaki, K. Toriyama, K. Nunone, J. Chem. Soc. Chem.
to 25 8C resulted in the formation of deep-blue needle-shaped crystals. Commun. 1983, 320 – 322.
The compound was stable at 25 8C in an inert atmosphere, but it decom- [10] a) J. R. Bolton, A. Carrington, Proc. Chem. Soc. 1961, 174 – 175;
posed within 10 min when exposed to air. b) F. Marchetti, C. Pinzino, S. Zachini, G. Pampaloni, Angew. Chem.
2010, 122, 5396 – 5400; Angew. Chem. Int. Ed. 2010, 49, 5268 – 5272.
ACHTUNGRE[C6I6]+ ACHTUNGRE[CF3SO3] : When [C6I6]+ACHTUNGRE[AsF6] (120 mg) was heated to 145 8C [11] a) T. J. Sears, T. A. Miller, V. E. Bondybey, J. Chem. Phys. 1981, 74,
under high vacuum (in an attempt to sublime the salt intact), the sample 3240 – 3248; b) W. D. Hobey, A. D. McLachlau, J. Chem. Phys. 1960,
started to slowly decompose into orange C6I6, purple I2, and other un- 33, 1695 – 1703.
identified products without any indication of the sublimation of [C6I6]+- [12] C. A. Coulson, A. Golebiewski, Mol. Phys. 1962, 5, 71 – 84.
ACHTUNGRE[AsF6] . The apparatus was cooled to 25 8C after approximately half of [13] L. C. Snyder, J. Chem. Phys. 1960, 33, 619 – 621.
the sample had been converted into C6I6. To recover the remaining unde- [14] A. D. Liehr, Z. Phys. Chem. (Muenchen Ger.) 1956, 9, 338 – 354.
composed sample of [C6I6]+ACHTUNGRE[AsF6] , a small amount of anhydrous [15] K. Raghavachari, R. C. Haddon, T. A. Miller, V. E. Bondybey, J.
CF3SO3H was added under an Ar atmosphere. After 2 days at 25 8C, Chem. Phys. 1983, 79, 1387 – 1395.
deep-blue crystals of [C6I6]+ACHTUNGRE[CF3SO3] had crystallized out of this solu- [16] J. Eiding, R. Schneider, W. Domke, H. Koppel, W. von Niessen,
tion. Chem. Phys. Lett. 1991, 177, 345 – 351.
Quantum-chemical calculations: Halogenated benzenes C6X6 (X = F, Cl, [17] R. Lindner, K. Mueller-Dethlefs, E. Wedum, K. Haber, E. R. Grant,
Br, I) and partially-fluorinated benzenes C6HF5 and 2,3,5,6-, 2,4,5,6-, and Science 1996, 271, 1698 – 1702.
3,4,5,6-C6H4F2 as well as their corresponding single-positively charged [18] K. Mueller-Dethlefs, J. B. Peel, J. Chem. Phys. 1999, 111, 10 550 –
cation radicals were investigated by quantum-chemical calculations by 10 551.
using the B3LYP DFT-functional[45] and the TZVPP basis set.[46] For I [19] B. E. Applegate, T. A. Miller, J. Chem. Phys. 2002, 117, 10654 –
atoms, the Def2-ECP28 pseudo-potential[47] was applied. Test calculations 10674.
for C6F6 and [C6F6]+ with different basis sets and different functionals re- [20] V. P. Vysotsky, G. E. Salmikov, L. N. Shchgoleva, Int. J. Quantum
sulted in similar relative energies, as did the B3LYP/TZVPP and Chem. 2004, 100, 469 – 476.
CCSD(T) levels of theory.[26] The structures of the neutral molecules and [21] K. F. Hall, A. M. Tokmachev, M. J. Bearpark, M. Boggia-Pasqua, M.
the cation radicals were optimized. The two energetically similar isomers Robb, J. Chem. Phys. 2007, 127, 134111 – 1 – 134111 – 11.
with the same symmetry were calculated by starting with models that had [22] N. M. Bazhin, Yu. V. Pozdnyakovich, V. D. Shteingarts, G. G. Yakob-
4 C C bond lengths longer than 2 C C bond lengths and vice versa. Fre- son, Izv. Akad. Nauk SSSR Ser. Khim. 1969, 10, 2300 – 2302.
quency calculations were performed to determine whether the optimized [23] T. J. Richardson, N. Bartlett, J. Chem. Soc. Chem. Commun. 1974,
structures were a true minimum or a transition state. The program pack- 427 – 428.
ages used were GAUSSIAN 03[48] and TURBOMOLE5v10.[49] [24] T. J. Richardson, F. L. Tanznella, N. Bartlett, Adv. Chem. Ser 1987,
217, 169 – 176.
One reviewer brought to our attention a distantly related Jahn–Teller
[25] S. D. Brown, T. M. Loehr, G. L. Gard, J. Fluorine Chem. 1976, 7,
isomer problem: the [CuACHTUNGRE(NO2)6]4 anion has been found to exist both as
19 – 32.
an elongated and a compressed octahedron, depending on the type of
[26] H. Shorafa, D. Mollenhauer, B. Paulus, K. Seppelt, Angew. Chem.
cation. Furthermore there even exists a regular octahedral anion that is
2009, 121, 5959 – 5961; Angew. Chem. Int. Ed. 2009, 48, 5845 – 5847.
generated by a dynamic Jahn–Teller effect.[50, 51]
[27] W. D. Stohrer, R. Hoffmann, J. Am. Chem. Soc. 1972, 94, 779 – 786.
CCDC-703472 ([C6F6]+ [Os2F11] ), CCDC-703473 ([C6F6]+ [Sb2F11] ), [28] A similar “mexican hat” energy surface has been observed in C5H5
CCDC-703623 ([C6F6]), CCDC-844888 ([C6HF5]+ [AsF6] ), CCDC-844885 (see: H. J. Wçrner, F. Markt, J. Chem. Phys. 2007, 127, 034 303) and
([C6H2F4]+ [SbF6] ·0.5 HF), CCDC-844884 ([C6H2F4]), CCDC-844886 in AuX3 (X = F, Cl, Br; see: A. Schulz, M. Hargittai, Chem. Eur. J.
([C6H3F3]+ [AsF6] ), CCDC-844887 ([C6H3F3]+ [SbF6] ), CCDC-844883 2001, 7, 3657 – 3669).
([C6Cl6]+ [Sb2F11] ), CCDC-844882 ([C6Br6]+ [Sb2F11] ), CCDC-844889 [29] C. M. Reddy, M. T. Kirchner, R. C. Gundakaram, K. A. Padmanab-
([C6I6]+ [AsF6] ), CCDC-844890 ([C6I6]+ [SbF6] ), and CCDC-844891 han, G. R. Desiraju, Chem. Eur. J. 2006, 12, 2222 – 2234.
([C6I6]+ [HS2O6C2F6] ) contain the supplementary, crystallographic data [30] G. M. Brown, O. W. Strydom, Acta Crystallogr Sect. B. 1974, 30,
for this paper. These data can be obtained free of charge from The Cam- 801 – 804.
bridge Crystallographic Data Centre via www.ccdc.cam.ac.uk/data_ [31] D. J. Sagl, J. C. Martin, J. Am. Chem. Soc. 1988, 110, 5827 – 5833.
request/cif. [32] J. C. Martin, L. J. Schaad, Pure Appl. Chem. 1990, 62, 547 – 550.

&10& www.chemeurj.org  2012 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Chem. Eur. J. 0000, 00, 0 – 0

ÝÝ These are not the final page numbers!


Halogenated Benzene Cation Radicals
FULL PAPER
[33] K. M. Patel, Diss. Abstr. Int. B 1994, 54, 4679, Vanderbilt Univ., [48] Gaussian 03, Revision C.02, M. J.Frisch, G. W. Trucks, H. B. Schlegel,
Nashville, TN, USA. G. E. Scuseria, M. A. Robb, J. R. Cheeseman, J. A. Montgomery, Jr.,
[34] A crystal structure of [C6I6]+ [AsF6]+ was also presented in 1997 at T. Vreven, K. N. Kudin, J. C. Burant, J. M. Millam, S. S. Iyengar, J.
the 15th International Symposium on Fluorine Chemistry (Vancou- Tomasi, V. Barone, B. Mennucci, M. Cossi, G. Scalmani, N. Rega,
ver, CA) by J. B. Chestnut, R. D. Rogers, J. S. Thrasher, A. Water- G. A. Petersson, H. Nakatsuji, M. Hada, M. Ehara, K. Toyota, R.
feld, (personal communication from A. Waterfeld and J. S. Thrash- Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao,
er). H. Nakai, M. Klene, X. Li, J. E. Knox, H. P. Hratchian, J. B. Cross,
[35] R. W. A. Havenith, P. W. Fowler, S. Fias, B. Bultwick, Tetrahedron V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R. E. Stratmann,
Lett. 2008, 49, 1421 – 1424. O. Yazyev, A. J. Austin, R. Cammi, C. Pomelli, J. W. Ochterski, P. Y.
[36] I. Ciofini, P. P. Laine, C. Adamo, Chem. Phys. Lett. 2007, 435, 171 – Ayala, K. Morokuma, G. A. Voth, P. Salvador, J. J. Dannenberg,
175. V. G. Zakrzewski, S. Dapprich, A. D. Daniels, M. C. Strain, O.
[37] S. Ghosh, C. M. Reddy, G. R. Desiraju, Acta Crystallogr. Sect. E Farkas, D. K. Malick, A. D. Rabuck, K. Raghavachari, J. B. Fores-
2007, 63, o910-o911. man, J. V. Ortiz, Q. Cui, A. G. Baboul, S. Clifford, J. Cioslowski,
[38] S. H. Strauss, K. D. Abney, O. P. Anderson, Inorg. Chem. 1986, 25, B. B. Stefanov, G. Liu, A. Liashenko, P. Piskorz, I. Komaromi, R. L.
2806 – 2812. Martin, D. J. Fox, T. Keith, M. A. Al-Laham, C. Y. Peng, A. Na-
[39] C. A. Reed, K.-C. Kim, E. S. Stoyanov, D. Stasko, F. S. Tann, L. J. nayakkara, M. Challacombe, P. M. W. Gill, B. Johnson, W. Chen,
Mueller, P. D. W. Boyd, J. Am. Chem. Soc. 2003, 125, 1796 – 1804. M. W. Wong, C. Gonzalez, J. A. Pople, Gaussian, Inc., Wallingford
[40] D. L. Mattern, J. Org. Chem. 1983, 48, 4772 – 4773. CT, 2004.
[41] O. Ruff, A. Braida, O. Breitschneider, W. Menzel, H. Plaut, Z. [49] TURBOMOLE V5.10 2008, a development of University of Karls-
Anorg. Allg. Chem. 1932, 206, 59 – 64. ruhe and Forschungszentrum Karlsruhe GmbH, 1989 – 2007, TUR-
[42] J. Shamir, J. Binenboym, Inorg. Chim. Acta 1968, 2, 37 – 38; D. E. BOMOLE GmbH, since 2007; available from http://www.turbomole.
McKee, N. Bartlett, Inorg. Chem. 1973, 12, 2738 – 2740. com.
[43] V. P. Reddy, D. R. Bellew, G. K. S. Prakash, J. Fluorine Chem. 1992, [50] D. L. Cullen, E. C. Lingafelter, Inorg. Chem. 1971, 10, 1264 – 1268.
56, 195 – 197. [51] M. D. Joesten, S. Tagagi, P. G. Lenhart, Inorg. Chem. 1977, 16, 2680 –
[44] G. M. Sheldrick, Acta Crystallogr. Sect. A 2008, 64, 112. 2685.
[45] A. D. Becke, J. Chem. Phys. 1993, 98, 5648 – 5652.
[46] F. Weigend, R. Ahlrichs, Phys. Chem. Chem. Phys. 2005, 7, 3297 –
3305. Received: September 21, 2011
[47] K. A. Peterson, D. Figgen, E. Goll, H. Stoll, M. Dolg, J. Chem. Revised: January 18, 2012
Phys. 2003, 119, 11113 – 11123. Published online: && &&, 0000

Chem. Eur. J. 2012, 00, 0 – 0  2012 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.chemeurj.org &11&
These are not the final page numbers! ÞÞ
K. Seppelt et al.

Jahn–Teller distortion The new radicals: The [C6I6]+ radical


cation is, in contrast to all other
M. J. Molski, D. Mollenhauer, S. Gohr, [C6X6]+ radical cations presented
B. Paulus, M. A. Khanfar, H. Shorafa, herein, not Jahn–Teller distorted.
S. H. Strauss, K. Seppelt* . . . &&&&—&&&&
Halogenated Benzene Cation Radicals

&12& www.chemeurj.org  2012 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Chem. Eur. J. 0000, 00, 0 – 0

ÝÝ These are not the final page numbers!

You might also like