Bromine Monofluoride

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 5

ARTICLE

DOI: 10.1002/zaac.201200293

Bromine Monofluoride
Thomas Drews[a] and Konrad Seppelt*[a]
Dedicated to Professor Rudolf Hoppe on the Occasion of His 90th Birthday

Keywords: Bromine monofluoride; NMR spectroscopy; Raman spectroscopy; Bromine; Fluorine; Crystal structure

Abstract. BrF can be prepared by reaction of Br2 and F2 in various The ordered structure is made of Br–Cl···Br–Cl chains. Computations
solvents below –100 °C. In solution it is identified by 19F NMR and suggest that neat BrF crystallizes in a dipol-dipole type polymeric
Raman spectroscopy. From CH3Cl solution a single crystal structure structure rather than a halogen bonded one as in ClF. A possible inter-
determination of a BrF/CH3Cl adduct was obtained. The preparation mediate along the decomposition pathway into BrF3 and Br2 is Br–
from BrF3 and Br2 at 600 °C also delivers BrF in small amounts. The BrF2.
crystal structure of BrCl comes in an ordered and a disordered variety.

Introduction Results and Discussion


Simple, binary compounds have a principle importance, e.g. We obtained BrF in two ways, first by com-proportionation
for the test of theoretical models. A historic and famous exam- of BrF3 and Br2 at 600 °C, as suggested previously.[14] This
ple is the detection of noble gas fluorides in 1962,[1,2,3] that mixture is heated in a monel autoclave and released into a
came at a time when theoretical modelation has already been vacuum at –196 °C. The yellow product is dissolved in C2H5F
possible. The related halogen fluorides are known much at –100 °C. The NMR spectrum showed the presence of BrF,
longer, however. It is the merit of Otto Ruff to have prepared but the major products are still Br2 and BrF3, as shown by
many of them for the first time.[4] The only latecomers here Raman spectroscopy.
are ClF5[5] and IF3.[6] The diatomic inter halogen compounds Alternatively Br2 was treated in solution with elemental flu-
X–Y have been the starting point of the electronegativity scale orine (20 % in nitrogen) between –95 °C and –110 °C in vari-
by L. Pauling.[7] The prediction there has been that their sta- ous solvents (see Table 1). The same method was tried before
bility against decomposition into the elements would be largest at –78 °C in CFCl3, but the only product obtained was very
if the electronegativity differences between X and Y are largest. pure BrF3.[15] The choice of solvents is limited. CFCl3 and
IF and BrF should be particularly stable. This is certainly true, other chlorofluorocarbons do not dissolve Br2 enough at
but both disproportionate rapidly into IF5 and I2 and BrF3 and –100 °C. Propionitrile works well but BrF reacts with it slowly
Br2. even at this temperature. The solvents CH3Cl, C2H5F, and
Of course the molecule BrF exists in the gas phase, it was CH3F worked best, but due to their low boiling point the pro-
detected by spectroscopic methods[8,9] or in matrices.[10] Also cedure is tricky. CH2Cl2 works also fine, however Br2 is not
it is an obvious intermediate in addition reactions: soluble enough in CH2F2. So we thought that CH2ClF might
SF4 + BrF3 / Br2 / CsF 씮 SF5Br[11] be a good compromise. Strangely enough CH2ClF reacts ex-
CF3–CF=CF2 + BrF3/Br2 씮 CF3–CBrF–CF3[12]
In the following we show that BrF can indeed be obtained Table 1. 19F NMR chemical shifts and Raman lines of BrF in various
in substance, but it is very unstable. Without diminishing the solvents.
merits of O. Ruff it must be stated that his description of BrF 19
F NMR /ppm Raman /cm–1
as an orange compound, melting around –35 °C and boiling
C2H5CN –288.2 543
under decomposition around 20 °C, cannot be correct.[13] If CH3CN/CH2Cl2 –285.7
BrF would be similarly weakly associated in the condensed CH3Cl –349.9 585
phases as ClF we extrapolated a boiling point of about –65 °C! CH2Cl2 –371.6 569
CH2ClFa) –388.2
C2H5F –452.0 633
* Prof. Dr. K. Seppelt CH3F –453.9 637
E-Mail: Seppelt@chemie.fu-berlin.de Matrixb) 661
[a] Institut für Chemie, Anorganische und Analytische Chemie Gas IRc) 662.3 / 660.7 79/81
BrF
Freie Universität Berlin
Fabeckstr. 34–36 a) This value is obtained from a reaction in CH3F and subsequent
14195 Berlin, Germany addition of CH2ClF. b) Ref. [10]. c) Reference[9].

2106 © 2012 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Z. Anorg. Allg. Chem. 2012, 638, (12-13), 2106–2110
Bromine Monofluoride

plosively with 20 % F2 in nitrogen even at –100 °C, while in


CH2Cl2 and CH2F2 only traces of fluorination products are
detected at –100 °C.[16]
Usually the reaction is done in 5 mm NMR tubes so that 19F
NMR spectra can be taken directly. In those cases, in which
Br2 does not dissolve completely at –100 °C it is dissolved at
higher temperatures and the solution is shock frozen to
–196 °C to obtain very finely divided Br2. During the fluorina-
tion the yellow-orange color of the solution lightens up, and
un-dissolved Br2 goes into solution. In Table 1 the NMR and
Raman data are collected.
The 19F NMR spectra show clearly the advent of a new
compound with an extreme high field shift, besides traces of
Figure 2. Raman spectrum of BrF in CH3Cl at –100 °C. Raman lines
solvent fluorination products that usually can be identified, see of CH3Cl with higher wave numbers are cut off.
Figure 1. Since ClF (–419.4 ppm[17]) and related low valent
main group fluorides like CF3–S–F (–351 ppm[18]) or R–Se–F
tial linear array Cl···Br–F (179.0°) is a clear indication. The
(–312.8 to –371.0 ppm[19]) have similar high field shift reso-
interaction between the two adduct components is considered
nances, we were certain that BrF has been formed. BrF3 ap-
to be weak: The Br–F bond length [182.2(2) pm] is only a
pears only after excess fluorination or after warming with a
little longer than the bond length in the gaseous state, which
broad signal around –22 ppm. Another characteristic of the
was obtained from the microwave spectrum of 79Br–F
BrF NMR signal is the extreme solvent shift. It is obvious that
[175.9(10) pm[20]] The C–Cl bond length in the adduct has
increasing donor strengths of the solvents shifts the resonance
increased to 179.8(3) pm as compared to gaseous CH3Cl
dramatically to lower field (Table 1). It is safe to assume that
[177.60(3) pm)[21]], whereas the Br···Cl distance is
pure BrF would have shifts close to those observed in CH3F
264.0(2) pm. The bond length of the single bond in BrCl is
and C2H5F solutions since these solvents can be considered as
217.9(2) pm, see below.
extremely low basic.

Figure 3. Structure of BrF·CH3Cl in the crystal, Ortep representation,


50 % displacement ellipsoids.

We were not able to crystallize BrF without solvent. For this


case the extreme low basic solvents C2H5F and CH3F were
Figure 1. 19F NMR spectrum of BrF in CH3Cl at –100 °C. The tiny tested. Indeed, BrF precipitates below –130 °C as a fine, yel-
signals between –20 and –260 ppm are BrF3 and fluorination products low powder that can be re-dissolved at –100 °C. Evaporation
of CH3Cl: –28.3 ppm, broad, BrF3; –73.0 ppm doublet, JHF = 63.6 Hz,
of the solvent to dryness at –130 °C induced decomposition
CHF3; –143.1 ppm, triplet, JHF = 98.3 Hz, CH2F2; –170.0 ppm, triplet,
JHF = 96.8 Hz, CH2ClF; –267.3 ppm, quartet, JHF = 138.7 Hz, CH3F. into BrF3 and Br2. It seems that BrF is only stable in the pres-
ence of a donor solvent, even if these are such a weak donors
The Raman spectra in solution show the BrF vibration be- like CH3F and C2H5F.
tween 543 and 637 cm–1. These values also have a strong sol- So the crystal structure of neat BrF remains unknown. This
vent shift that parallels the 19F NMR shifts: Higher basicity of would be of interest in the light of the crystal structures of ClF
the solvent decreases the Br–F frequency. This is also interpre- and BrCl. In ClF the dominating intermolecular interactions
ted as a donor acceptor complex formation solvent Br–F in are between the chlorine atoms resulting in a zig-zag chain
solution (Figure 2). with “terminal” fluorine atoms.[22] This is a very simple exam-
ple of so called halogen bonding.
The crystal structure of BrCl is to the best of our knowledge
Crystal Structure of BrF···CH3Cl
unknown. Crystals of yellow BrCl are easily obtained at
Once formed BrF is very soluble in the named solvents. –78 °C, they come in two modifications. The first modification
Only in case of CH3Cl we were able to crystallize it (see Fig- is nothing but a rotational disordered Cl2 or Br2 type structure,
ure 3). It appears as adduct with CH3Cl. Obviously the chlo- where bromine and chlorine are not discriminated. Lattice con-
rine atom acts as donor to the bromine atom in BrF. The essen- stants, positional parameters, and bond length are very close

Z. Anorg. Allg. Chem. 2012, 2106–2110 © 2012 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.zaac.wiley-vch.de 2107
ARTICLE T. Drews, K. Seppelt

Table 2. X-ray results of the disordered BrCl modifications, all spaces groups are Cmca.
a /pm b /pm c /pm V /106pm3 T /°C
a)
Br2 665.67(4) 455.4(3) 870.68(5) 263.95 –103 °C Br–Br = 230.1(2) pm
Cl2a) 622.35(2) 445.61(1) 817.85(2) 226.8 –173 °C Cl–Cl = 199.4(2) pm
BrCl 645.1(2) 447.1(2) 846.5(2) 244.15 –160 °C Br–Cl = 220.5(1) pm
a) B. M. Powell, K. M. Heal, B. H. Torrie, Mol. Phys. 1984, 53, 929–939.

to the averaged numbers of the known Br2 and Cl2 structures In the Br–F system BrF3/Br2 is the most stable composition,
within the same space group Cmca (see Table 2). when the strong interaction of BrF3 molecules is taken into
In the ordered structure the bromine and chlorine positions account, which is also experimentally established. Dimeriza-
are separated, although the lattice parameters have hardly tion of BrF molecules results in a slight preference for the
changed. The Br–Cl bond length is 217.8(2) pm. This com- FBr···FBr combination over the FBr···BrF configuration by
pares to 214 pm in the free molecule.[23] 0.38 kcal·mol–1. In order to arrive at a more realistic model for
In the crystal the Br–Cl molecules are arranged in a zig-zag the condensed phase we calculated a tetramer (FBr···BrF)2 and
chain with close Br···Cl contacts is found (see Figure 4), quite a hexamer (FBr···FBr)3. These cyclic oligomers were chosen
similar to the long known structure of ICl.[24] since they have realistic bridge angles which are also expected
in polymer chain arrangements. Herein also a preference of the
FBr···FBr interaction is observed over the FBr···BrF interaction
(0.6 kcal·mol–1) (however (FI···IF)2 seems to be more stable
than (FI···FI)3!).
The dimerization of BrF into Br–BrF2, FBr···BrF, and
FBr···FBr has already been calculated before by various DFT
Figure 4. The ordered crystal structure of Br–Cl, showing a similar methods.[25] The relative energies vary somewhat, but overall
chain as in ICl. Ortep representation, 50 % probability, displacement similar qualitative similar results are obtained.
ellipsoids. Our calculations also show that the stabilization of BrF by
For BrF the question is whether the larger atom size in BrF CH3F and more so by CH3Cl can prevent the disproportiona-
versus ClF would favor a halogen bond structure as in ClF, or tion into BrF3 + Br2. Some of the calculated structures of the
whether the larger dipole momentum in BrF versus ClF would BrF system are shown in Figure 6.
favor a BrCl or ICl like structure. Experimentally we cannot One interesting local minimum is the T-shaped compound
answer this question. Br–BrF2. This has been possibly found in a matrix isolation
In order to arrive at least at a guess we have computed the study[10] but identified only by one absorption in the IR (at
BrF system (see Figure 5), also some calculations of the ClF 555 cm–1, νasBrF2). We have calculated the whole vibrational
and IF system are included. spectrum, and the most characteristic line should be the Br–Br
These MP2 calculations reproduce some long known experi- stretching mode at around 300 cm–1 (see Table 3). But this is
mental facts: ClF and ClF3/Cl2 are close to equilibrium, while very close to the Br–Br stretching mode in Br2 (316 cm–1 dis-
the IF5/I2 combination is by far the most stable composition solved, 296 cm–1 solid), which is extremely intense in the Ra-
for “IF”. man spectrum, so that any Br–BrF2 might be obscured in our

Figure 5. Energetics of the BrF system, as compared to the ClF and IF system. Molecular ClF, BrF, and IF are set arbitrarily to 0 kcal·mol–1.
Energies are normalized to the composition of X–F in each case for comparison.

2108 www.zaac.wiley-vch.de © 2012 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Z. Anorg. Allg. Chem. 2012, 2106–2110
Bromine Monofluoride

Physical Measurements: JEOL JNM-LA 400MHz multinuclear NMR


spectrometer, 1403 Spex industries Raman spectrometer with
1064 nm / 300 mW laser excitation, Bruker SMART CCD 1000 TK
diffractometer with Mo-Kα irradiation and graphite monochromator,
ω-scan method.

Bromine Fluoride: Usually Br2 (10–20 mg) was dissolved in solvent


(2 mL) in a 5 mm NMR tube and cooled to –100 °C. C2H5–CN,
CH2Cl2, CH3Cl dissolve Br2 completely, also the CH3–CN/CH2Cl2
mixture. Suspensions of finely divided Br2 in C2H5F and CH3F were
obtained by dissolving similar amounts at –30 °C (C2H5F) of –60 °C
(CH3F) in closed tubes (caution, high pressure!) and shock frozen to
–196 °C. 20 % F2 in N2 was passed through a NaF column in order to
free it from residual HF and afterwards bubbled into the solutions by
a PFA capillary (PFA = perfluorovinyl ether tetrafluoro ethylene co-
polymerisate). The orange color of the bromine solutions lightened up
to yellow, and in those cases where the Br2 was only partially dis-
solved, it slowly went into solution. In these cases the fluorination was
stopped when a small residue of solid bromine was left in order to
avoid over-fluorination. Usually the fluorination lasted 45 min at a
speed of 1 gas bubble per second through a 1 mm thick PFA capillary.
During the fluorination the temperature was held between –95 °C and
–110 °C. The samples were transferred without warming into the NMR
and Raman spectrometer.
Figure 6. Calculated structures /pm,° and energetics (a.u., zero point
energy corrected, MP2 approximation) of some Br–F systems.
Crystal Growth and X-ray Measurement: A solution of BrF in
CH3Cl was prepared as described. The solvent was reduced by evapo-
spectra. The Br–F stretching region is always disturbed by the ration at –120 °C to about 10 % of its volume. Subsequently, CH2ClF
presence of BrF3, whose Raman absorptions vary very much (1 mL) was added. This mixture was slowly cooled to –140 °C when
with concentration, temperature, and solvent. The three defor- orange crystals appeared. With a special device modified[27] one crys-
mation modes of Br–BrF2 may be characteristic but are pre- tal was mounted on a glass capillary and glued on it by a perfluoro-
dicted to be weaker in the Raman spectrum, and none of them hexane/perfluoro vinylether oil. 1800 frames by 0.3° scans in ω were
can be located with certainty (Table 3). taken, 20 sec per frame, covering a full sphere of reflections, followed
The analogue Cl–ClF2 is predicted to be quite unstable, on by semi empirical absorption correction by equalizing symmetry
the contrary I–IF2 could be quite stable, similar to IF3 that has equivalent reflections (method SADABS). The structure was solved
and refined by the Sheldrick programs[28] (for details see Table 4).
be isolated.[6] The possible dimerization of BrF into Br–BrF2
is paralleled by the experimentally established dimerization of Bromine Chloride: Bromine (50 mg) was dissolved in CH2Cl2
SF2 into F–S–SF3.[26] (3 mL). With a glass vacuum line an excess of Cl2 was condensed into
the glass ampoule, and the ampoule was sealed. Warming to room
temperature und shaking the content, followed by slowly cooling to
Conclusions –78 °C overnight, afforded orange crystals. The crystallographic pro-
BrF can be prepared, but it is very unstable. The question cedure was done as above.
of its solid-state structure remains open. Another interesting
Computational Procedure: The ab initio and DFT calculations were
case is that of IF. A similar approach for its preparation seems performed with the Gaussian 03 programs[29] using the implemented
impossible so long as no suitable solvent for I2 at –100 °C or MP2, MP4, CCSD(T), and Becke3LYP procedures as implemented in
below is found. the program. Basis sets: Cl, F: aug-cc-pVTZ as implemented in the
program, Br: Relativistically corrected core potential for 28 electrons
and a 14s10p2d1f(3s3p2d1f) basis set, I: Relativistically corrected core
Experimental Section potential for 28 electrons and a 22s21p10d2f(6s5p4d2f) basis set.[30]
Caution: Handling 20 % F2 in N2 requires rigorous safety protections.
In the majority of the calculations we used the MP2 approximation,
Materials: 20 % F2 in N2 was provided by Solvay Co, Bad Wimpfen, since it is known that this method is able to reproduce weak interac-
Germany. All used solvents were treated before use with 20 % F2 in tions better than the more economical DFT methods, although it might
N2 at –100 °C. overestimate all bond strengths. More precise thermodynamic data

Table 3. Calculated vibrational spectra of Br–BrF2 /cm–1 (Raman intensity).


νasBrF2, B2 νsBrF2, A1 νBrBr, A1 τBrF2, B1 δsBrF2, A1 δasBrF2, B2
CCSD(T) 581 527 310 255 2001 184
MP4 564 500 308 246 184 178
MP2 593(0.9) 530(42.8) 329(7.5) 255(0.0) 202(2.7) 187(3.8)
B3LYP 560(0.9) 512(31.0) 296(8.3) 254(0.0) 180(4.0) 177(4.0)

Z. Anorg. Allg. Chem. 2012, 2106–2110 © 2012 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.zaac.wiley-vch.de 2109
ARTICLE T. Drews, K. Seppelt

Table 4. Crystallographic data of BrF·CH3Cl and BrCl (ordered modi- [11] R. Winter, R. J. Terjeson, G. L. Gard, J. Fluorine Chem. 1988,
fication). 89, 105–106.
[12] R. D. Chambers, W. K. R. Musgrave, J. Savory, J. Chem. Soc.
BrF·CH3Cl BrCl 1961, 3779; R. D. Chambers, W. K. R. Musgrave, J. Savory, Proc.
M 149.4 115.4 Chem. Soc. London 1961, 113.
[13] O. Ruff, A. Braida, Z. Anorg. Allg. Chem. 1933, 214, 81–90.
T /°C –150 –140
[14] R. K. Streunenberg, R. C. Vogel, J. Fischer, J. Am. Chem. Soc.
Space group Pnma (No 62) Cmc21 (No 36)
1957, 79, 1320–1323.
a /pm 945.9(2) 644.9(3)
[15] E. Lehmann, D. Naumann, M. Schmeisser, Z. Anorg. Allg. Chem.
b /pm 655.8(3) 454.7(2)
1972, 388, 1–3.
c /pm 692.6(1) 835.0(3)
[16] We tentatively explain the kinetic instability of CH2ClF against
V /106pm3 429.7 244.9
F2(N2(1:4) at –100 °C as follows: F2 is very insoluble in CH2Cl2,
Z 4 4
so only a slow reaction occurs. F2 is better soluble in CH2F2, but
ρ calcd 2.310 3.129
the further fluorination to CHF3 or CF4 is kinetically (higher posi-
Reflections
tive charge on the hydrogen atoms) and thermodynamically hin-
Collected 5231 1476
dered. In CH2ClF there might be higher solubility of F2 along
Independent 706 404 with a quicker and more exothermic reaction.
2θmax /° 61.04 61.00 [17] K. O. Christe, J. F. Hon, D. Philipovich, Inorg. Chem. 1973, 12,
μ /mm–1 10.0 17.44 84–89.
Rint 0.0248 0.0199 [18] F. Seel, W. Gombler, R. Budenz, Angew. Chem. 1967, 79, 686.
R refined parameters 32 14 [19] H. Poleschner, S. Ellrodt, M. Malischewski, K. Seppelt, Angew.
wR2 0.0470 0.1233 Chem. Int. Ed. 2012, 51, 419–422; H. Poleschner, K. Seppelt,
R [Fo ⬎ 4σ(Fo)] 0.0191 0.0450 Chem. Eur. J. 2004, 10, 6565–6574.
[20] D. F. Smith, M. Tidwell, D. V. P. Williams, Phys. Rev. 1956, 77,
420–421.
would be obtained by the CCSD(T) method, but these calculation ex- [21] P. Hansen, S. Brodersen, G. Guelachvili, J. Mol. Spectrosc. 1981,
ceeds our available computer resources at least for the larger systems. 88, 378.
[22] R. Boese, A. D. Boese, D. Bläser, M. Y. Antipin, A. Ellern, K.
Seppelt, Angew. Chem. Int. Ed. Engl. 1997, 36, 1489–1491.
Further details of the crystal structure investigations may be obtained
[23] W. V. F. Brooks, B. Crawford Jr., J. Chem. Phys. 1955, 23, 363;
from the Fachinformationszentrum Karlsruhe, 76344 Eggenstein- H. Stammreich, D. Bassi, O. Sala, Spectrochim. Acta 1961, 17,
Leopoldshafen, Germany (Fax: +49-7247-808-666; E-Mail: 1173; D. F. Smith, M. Tidwell, D. V. P. Williams, Phys. Rev. 1950,
crysdata@fiz-karlsruhe.de, http://www.fiz-karlsruhe.de/request for de- 79, 1007–1008.
posited data.html) on quoting the depository number CSD-424851 [24] R. Minkwitz, M. Berkei, Z. Naturforsch. B 1999, 54, 1615–1617.
(BrF–CH3Cl) and CSD-424850 (BrCl). [25] L. Gong, Q. Li, W. Xu, Y. Xie, H. F. Schaefer III, J. Chem. Phys.
A 2004, 108, 3598–3614.
[26] W. Gombler, A. Haas, H. Willner, Z. Anorg. Allg. Chem. 1980,
Acknowledgements 469, 118.
[27] H. Schumann, W. Genthe, E. Hahn, M.-B. Hossein, D. v. d. Helm,
This work was supported by the Deutsche Forschungsgemeinschaft as J. Organomet. Chem. 1986, 298–317, 2561–2567.
[28] G. M. Sheldrick, Acta Crystallogr., Sect. A 2008, 64, 112–122.
part of the Graduiertenkolleg “Fluorine as a key Element“ DFG GK
[29] Gaussian 03, Revision D.01, M. J. Frisch, G. W. Trucks, H. B.
1582/1. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, J. A.
Montgomery Jr., T. Vreven, K. N. Kudin, J. C. Burant, J. M. Mil-
lam, S. S. Iyengar, J. Tomasi, V. Barone, B. Mennucci, M. Cossi,
References G. Scalmani, N. Rega, G. A. Petersson, H. Nakatsuji, M. Hada,
M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T.
[1] N. Bartlett, Proc. Chem. Soc. London 1962, 218. Nakajima, Y. Honda, O. Kitao, H. Nakai, M. Klene, X. Li, J. E.
[2] H. H. Claassen, H. Selig, J. G. Malm, J. Am. Chem. Soc. 1962, Knox, H. P. Hratchian, J. B. Cross, V. Bakken, C. Adamo, J. Jara-
84, 3593. millo, R. Gomperts, R. E. Stratmann, O. Yazyev, A. J. Austin, R.
[3] R. Hoppe, W. Dähne, H. Mattauch, K. M. Rödder, Angew. Chem. Cammi, C. Pomelli, J. W. Ochterski, P. Y. Ayala, K. Morokuma,
1962, 74, 903. G. A. Voth, P. Salvador, J. J. Dannenberg, V. G. Zakrzewski, S.
[4] F. Kraus, O. Ruff, Z. Anorg. Allg. Chem. 2012, 638, 707–709. Dapprich, A. D. Daniels, M. C. Strain, O. Farkas, D. K. Malick,
[5] D. F. Smith, Science 1963, 141, 1039–1040. A. D. Rabuck, K. Raghavachari, J. B. Foresman, J. V. Ortiz, Q.
[6] M. Schmeißer, W. Ludovici, D. Naumann, D. Sartori, E. Scharf, Cui, A. G. Baboul, S. Clifford, J. Cioslowski, B. B. Stefanov, G.
Chem. Ber. 1968, 101, 4214–4220; D. Naumann, E. Renk, E. Liu, A. Liashenko, P. Piskorz, I. Komaromi, R. L. Martin, D. J.
Lehmann, J. Fluorine Chem. 1977, 10, 395–403; S. Hoyer, K. Fox, T. Keith, M. A. Al-Laham, C. Y. Peng, A. Nanayakkara, M.
Seppelt, Angew. Chem. Int. Ed. 2000, 39, 1448–1449. Challacombe, P. M. W. Gill, B. Johnson, W. Chen, M. W. Wong,
[7] L. Pauling, J. Am. Chem. Soc. 1932, 54, 3570–3582. C. Gonzalez, and J. A. Pople, Gaussian, Inc., Wallingford CT,
[8] R. A. Durie, Proc. R. Soc. (London) Ser. A 1951, 207, 388–395. 2004.
[9] H. Bürger, P. Schulz, E. Jacob, M. Föhnle, Z. Naturforsch. A [30] Universität Stuttgart, Theoretische Chemie (theochem.uni-
1986, 41, 1015–1020. stuttgart.de).
[10] R. R. Smardzewski, W. B. Fox, J. Fluorine Chem. 1976, 7, 453–
455; E. S. Prochaska, L. Andrews, R. Norman, G. Mamantov, Received: June 25, 2012
Inorg. Chem. 1978, 17, 970–977 (Br–BrF2!). Published Online: August 8, 2012

2110 www.zaac.wiley-vch.de © 2012 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Z. Anorg. Allg. Chem. 2012, 2106–2110

You might also like