ArticlePDF Available

Complexes of BX 3 with EMe 2 (X = F, Cl, Br, I; E = Se or Te): Synthesis, multinuclear NMR spectroscopic and structural studies

Authors:

Abstract and Figures

The [BX3(EMe2)] (X = Cl, Br, I; E = Se or Te) have been prepared by reaction of BX3 with the EMe2 in hexane under anhydrous conditions. The X-ray crystal structures of [BX3(TeMe2)] (X = Cl, Br, I) and [BX3(SeMe2)] (X = Cl, Br) have been determined; all are pseudo-tetrahedral monomers and show d(B−E) decreases with halogen, Cl > Br > I. Multinuclear NMR data (¹H, ¹¹B, ⁷⁷Se and ¹²⁵Te) are reported and compared with data on the corresponding [BX3(SMe2)], and the trends discussed. The unstable [BF3(SeMe2)], prepared from BF3 and SeMe2 in the absence of a solvent, has been similarly characterised by multinuclear NMR spectroscopy, and evidence for the existence of unstable [BF3(TeMe2)] obtained for the first time, although it could not be obtained pure. The results are discussed in the light of recent theoretical modelling of boron halide adducts.
Content may be subject to copyright.
Accepted Manuscript
Complexes of BX3 with EMe2 (X=F, Cl, Br, I; E=Se or Te): Synthesis, multinuclear
NMR spectroscopic and structural studies
Coco K.Y.A. Okio, William Levason, Francesco M. Monzittu, Gillian Reid
PII: S0022-328X(17)30484-9
DOI: 10.1016/j.jorganchem.2017.08.004
Reference: JOM 20057
To appear in: Journal of Organometallic Chemistry
Received Date: 16 June 2017
Revised Date: 27 July 2017
Accepted Date: 4 August 2017
Please cite this article as: C.K.Y.A. Okio, W. Levason, F.M. Monzittu, G. Reid, Complexes of BX3 with
EMe2 (X=F, Cl, Br, I; E=Se or Te): Synthesis, multinuclear NMR spectroscopic and structural studies,
Journal of Organometallic Chemistry (2017), doi: 10.1016/j.jorganchem.2017.08.004.
This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to
our customers we are providing this early version of the manuscript. The manuscript will undergo
copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please
note that during the production process errors may be discovered which could affect the content, and all
legal disclaimers that apply to the journal pertain.
MANUS CRIP T
ACCEP TED
ACCEPTED MANUSCRIPT
MANUS CRIP T
ACCEP TED
ACCEPTED MANUSCRIPT
1
Complexes of BX
3
with EMe
2
(X= F, Cl, Br, I; E = Se or Te): synthesis,
multinuclear NMR spectroscopic and structural studies.
Coco K.Y.A. Okio
a
, William Levason
,b
, Francesco M. Monzittu
b
and Gillian Reid
b
a. Departamento de Química, Facultad de Ciencias, Universidad Nacional de Colombia,
Bogotá, Colombia.
b. Chemistry, University of Southampton, Southampton SO17 1BJ, UK; email: wxl@soton.ac.uk
Abstract
The [BX
3
(EMe
2
)] (X = Cl, Br, I; E = Se or Te) have been prepared by reaction of BX
3
with the EMe
2
in hexane under anhydrous conditions. The X-ray crystal structures of [BX
3
(TeMe
2
)] (X = Cl, Br, I)
and [BX
3
(SeMe
2
)] ( X = Cl, Br) have been determined; all are pseudo-tetrahedral monomers and show
d(BE) decreases with halogen, Cl > Br > I. Multinuclear NMR data (
1
H,
11
B,
77
Se and
125
Te) are
reported and compared with data on the corresponding [BX
3
(SMe
2
)], and the trends discussed. The
unstable [BF
3
(SeMe
2
)], prepared from BF
3
and SeMe
2
in the absence of a solvent, has been similarly
characterised by multinuclear NMR spectroscopy, and evidence for the existence of unstable
[BF
3
(TeMe
2
)] obtained for the first time, although it could not be obtained pure. The results are
discussed in the light of recent theoretical modelling of boron halide adducts.
Keywords: boron halide, selenoether, telluroether, boron-11 NMR, crystal structures
1. Introduction
The study of boron Lewis acids remains a very active area of research and has been discussed in
several recent reviews [1-6].
The boron(III) halides are good Lewis acids whose acidity generally
increases BF
3
< BCl
3
< BBr
3
< BI
3
, which is counter-intuitive on electronegativity grounds, and was
for many years attributed to significant XB π bonding which decreased down Group 17. Much
computational work over the past fifteen years has been devoted to exploring the factors involved,
almost all focussed on N- or O-donor ligands. The results of the DFT calculations may differ in fine
details, however it is generally agreed that the order of Lewis acidity results from the varying strength
of the σ-interactions, and the π-bonding explanation has been discounted [5,7-9]. It is important to
note that the properties of both the Lewis acid and Lewis base in a complex must be taken in
consideration, and that calculations deal with gas phase species, meaning solvation or solid state
effects may mask Lewis acidity trends in solution or in the solid state. Due to their innate Lewis
acidity, boron(III) halides have found widespread use in organic syntheses and catalysis [6,10]. There
Corresponding author.
MANUS CRIP T
ACCEP TED
ACCEPTED MANUSCRIPT
2
is also much current interest in the chemistry of boron-based frustrated Lewis pairs [11],
and the use
of boron compounds in fluoride sensing and PET imaging [12].
We recently reported a systematic study of the coordination chemistry of the neutral diphosphines,
R
2
P(CH
2
)
2
PR
2
(R = Me or Et) and o-C
6
H
4
(PR’
2
)
2
(R’ = Me or Ph), and the diarsine, o-C
6
H
4
(AsMe
2
)
2
towards BX
3
(X = F, Cl, Br and I). The studies revealed that whilst flexible ligands produced [X
3
B(µ-
LL)BX
3
] complexes, with o-phenylene linked diphosphines and diarsines, very rare dihaloboronium
cations, [BX
2
{o-C
6
H
4
(EMe
2
)
2
}]
+
(E = P, As) were obtained [13]. In these complexes and in the
corresponding [BX
3
(PMe
3
)] [14]
and [BX
3
(AsMe
3
)] [15] the d(BP) and d(BAs) bond lengths
follow the expected order with X (F < Cl < Br <I), although data on fluoride complexes are rather
limited.
Boron trihalide complexes of thio- and seleno-ethers were reported many years ago and are all of type
[BX
3
(ER
2
)] (X = F, Cl, Br or I; E = S or Se, R most often Me, sometimes Et or
i
Pr, R
2
= c-(CH
2
)
n
)
Typical syntheses involved direct reaction of the BX
3
and ER
2
in the absence of a solvent, whilst
others used alkanes, CCl
4
, CH
2
Cl
2
or CS
2
as solvents [16-25]. Generally, the complexes were
characterised by microanalysis,
1
H and sometimes
11
B NMR and IR spectroscopy. X-Ray structural
data are surprisingly rare with only the crystal structures of [BX
3
(tht)] (X = Cl, Br or I; tht =
tetrahydrothiophene) reported [26]. In these, the d(BS) bond length increased with halide Br > Cl >
I, the anomalous position of the bromide was suggested by the authors to be due either to solid state
effects or a consequence of the disorder of the tht molecule [26].
The unit cell data for several
[BX
3
(EMe
2
)] (E = S, Se) were reported in a conference paper [27], but the full structures have never
appeared. Very much less is known about TeMe
2
adducts [22,23], and no complex was reported to
form between BF
3
and TeMe
2
at ambient temperatures [28].
Here, we report a detailed study of the complexes [BX
3
(EMe
2
)] (E = Se or Te) focussing on
multinuclear NMR data and the X-ray crystal structures of [BX
3
(TeMe
2
)] (X = Cl, Br, I) and
[BX
3
(SeMe
2
)] ( X = Cl, Br). Multinuclear NMR spectroscopic data are also reported for [BX
3
(SMe
2
)]
for comparison. Comparisons between the boron complexes and chalcogenoether complexes with
other Group 13 halides are also described.
2. Experimental
Infrared spectra were recorded as Nujol mulls between CsI plates using a Perkin-Elmer
Spectrum 100 spectrometer over the range 4000–200 cm
−1
.
1
H,
11
B,
19
F{
1
H},
77
Se{
1
H}, and
125
Te{
1
H} NMR
spectra were recorded from CH
2
Cl
2
/CD
2
Cl
2
solutions using a Bruker AV400
spectrometer and
referenced to the residual solvent resonance, external [BF
3
(OEt
2
)], CFCl
3
,
neat SeMe
2
and neat TeMe
2
respectively. Microanalyses were undertaken by Medac Ltd.
Hexane was dried prior to use by distillation from sodium and CH
2
Cl
2
from CaH
2
, and all
preparations were carried out under rigorously anhydrous conditions via a dry dinitrogen
MANUS CRIP T
ACCEP TED
ACCEPTED MANUSCRIPT
3
atmosphere and standard Schlenk and glovebox techniques. Boron trifluoride was obtained
from Fluorochem. Other boron trihalides were obtained from Sigma-Aldrich and used as
received. SMe
2
and SeMe
2
(Sigma-Aldrich) were stored over molecular sieves. TeMe
2
was
made as described [29]. [BF
3
(SMe
2
)] and [BCl
3
(SMe
2
)] were commercial samples (Sigma-
Aldrich), and [BX
3
(SMe
2
)] (X = Br, I) were made as described, by reaction of the constituents
in n-hexane [23].
2.1 [BCl
3
(TeMe
2
)]
TeMe
2
(0.14 g, 0.9 mmol.) was dissolved in n-hexane (10 mL), the solution cooled in an ice bath, and
a slow stream of BCl
3
passed in, producing a white precipitate. The mixture was stirred for 30 min.
and then the solid filtered off and dried in vacuo. Yield: 0.185 g, 61%. Anal. Required for C
2
H
6
BCl
3
Te
(274.8): C, 8.74; H, 2.20%. Found: C, 9.19; H, 2.13%.
1
H NMR (CDCl
3
, 295 K): δ 2.10 (s).
11
B NMR
(CH
2
Cl
2
/CD
2
Cl
2
, 295 K): δ 20.01 (s),
125
Te NMR (CH
2
Cl
2
/CD
2
Cl
2
, 295 K): δ +242.9 (s). IR (Nujol):
ν = 721 (br) (BCl) cm
1
. Colourless crystals were grown by slow evaporation of a CH
2
Cl
2
solution of
the product.
2.2 [BBr
3
(TeMe
2
)]: To an n-hexane solution (10 mL) of BBr
3
(0.20 g, 0.8 mmol) in an ice bath, was
added dropwise TeMe
2
(0.126 g, 0.8 mmol), leading to the formation of an immediate white
precipitate. After stirring for 30 min, the white powder was isolated by filtration and dried in vacuo.
Yield: 0.230 g, 70%. Anal. Required for C
2
H
6
BBr
3
Te (408.2): C, 5.99; H, 1.48 %. Found: C, 6.94; H,
1.81%.
1
H NMR (CDCl
3
, 295 K): δ 2.10 (s).
11
B NMR (CH
2
Cl
2
/CD
2
Cl
2
, 295 K): δ −20.01 (s),
125
Te
NMR (CH
2
Cl
2
/CD
2
Cl
2
, 295 K): δ +242.9 (s). IR (Nujol): ν = 622 (s) (BBr) cm
1
. Small colourless
crystals were grown by slow evaporation of a CH
2
Cl
2
solution of the product.
2.3 [BI
3
(TeMe
2
)]
Powdered BI
3
(0.20 g, 0.5 mmol) was added to n-hexane (15 mL) and TeMe
2
(0.08g, 0.5 mmol) was
added dropwise to the stirred solution. A pale cream solid formed immediately and after stirring for
20 min. the white solid was filtered off and dried in vacuo. Yield: 0.19 g, 69%. C
2
H
6
BI
3
Te (549.2): C,
4.37; H, 1.10%. Found: C, 5.14; H, 1.68%.
1
H NMR (CDCl
3
, 295 K): δ 1.91 (s).
11
B NMR
(CH
2
Cl
2
/CD
2
Cl
2
, 295 K): δ −86.80 (s).
125
Te NMR (CH
2
Cl
2
/CD
2
Cl
2
, 295 K): δ +290.2 (q,
1
J
BTe
= 121
Hz). IR (Nujol): ν = 560(m), 543(s) (BI) cm
1
.
2.4 [BCl
3
(SeMe
2
)]
SeMe
2
(0.16 g, 1.5 mmol) was dispersed in stirred n-hexane (15 mL), and BCl
3
gas slowly bubbled
into the solution for 5 min, resulting in the rapid formation of a white powdery precipitate. The BCl
3
was stopped, and the mixture stirred for 30 min, after which the white precipitate was isolated by
filtration and dried to a white powder in vacuo. Yield: 0.120 g, 36%. C
2
H
6
BCl
3
Se (226.2): C, 10.62; H,
MANUS CRIP T
ACCEP TED
ACCEPTED MANUSCRIPT
4
2.67%. Found: C, 10.61; H, 2.65%.
1
H NMR (CDCl
3
, 295 K): δ 2.36 (s).
11
B NMR (CH
2
Cl
2
/CD
2
Cl
2
,
295 K): δ +7.36 (s).
77
Se NMR (CH
2
Cl
2
/CD
2
Cl
2
, 295 K): δ +170.5 (s). IR (Nujol): ν = 765 (m), 723
(s) (BCl) cm
1
.
2.5 [BBr
3
(SeMe
2
)]
To a solution of SeMe
2
(0.087 g, 0.8 mmol) in n-hexane (8 mL) was added dropwise BBr
3
(0.20 g, 0.8
mmol) which immediately led to the precipitation of a white solid. The reaction was stirred for 30
min, and then the white powder was isolated by filtration and dried in vacuo. Yield: 0.134 g, 47%.
C
2
H
6
BBr
3
Se (359.6): C, 6.68; H, 1.68%. Found: C, 6.76; H, 2.60%.
1
H NMR (CDCl
3
, 295 K): δ 2.33
(s).
11
B NMR (CH
2
Cl
2
/CD
2
Cl
2
, 295 K): δ −12.94 (s).
77
Se NMR (CH
2
Cl
2
/CD
2
Cl
2
, 295 K): δ +194.4
(s). IR (Nujol): ν = 680(m), 640 (s) (BBr) cm
1
.
2.6 [BI
3
(SeMe
2
)]
To a n-hexane (10 ml) solution of BI
3
(0.05 g, 0.13 mmol) was added dropwise Me
2
Se (0.014 g, 0.13
mmol) leading to the immediate formation of a white precipitate. After stirring the mixture for a
further 30 min. the white powder was filtrated and dried in vacuo. Yield: 0.022 g (34%). C
2
H
6
BI
3
Se
(500.6): C, 4.80; H, 1.21%. Found: C, 4.87; H, 1.33%.
1
H NMR (CDCl
3
, 295 K): δ 2.30 (s).
11
B NMR
(CH
2
Cl
2
/CD
2
Cl
2
, 295 K): δ −73.96 (s),
77
Se NMR (CH
2
Cl
2
/CD
2
Cl
2
, 295 K): δ +203.7 (q,
1
J
BSe
= 52
Hz). IR (Nujol): ν = 560 (m), 550 (s) (BI) cm
1
.
2.7 [BF
3
(SeMe
2
)]
Neat SeMe
2
(0.22 g, 2.0 mmol) was cooled in an ice-bath and a slow stream of BF
3
bubbled in for 5
min. The product was a clear straw coloured liquid which fumes in air and is hydrolysed by trace
moisture. The liquid was stored under a nitrogen atmosphere and measurements made on freshly
prepared samples. The complex has a significant vapour pressure of BF
3
at room temperature and the
microanalysis cannot be obtained.
1
H NMR (CDCl
3
, 295 K): δ 2.04 (s).
11
B NMR (CH
2
Cl
2
/CD
2
Cl
2
,
295 K): δ +5.31 (s); (183 K): δ +3.25 (s).
19
F{
1
H) NMR (CH
2
Cl
2
/CD
2
Cl
2
, 295 K): δ −132.8 (s); (183
K): δ −134.2 (s).
77
Se NMR (CH
2
Cl
2
/CD
2
Cl
2
, 295 K): δ +1.03 (s); (183 K): δ +0.53 (s).
2.8 [BF
3
(TeMe
2
)]
Treatment of TeMe
2
with BF
3
gas either at ambient temperature or in an ice bath gave a yellow oil
with some orange-red solid. Similarly, adding BF
3
gas to a solution of TeMe
2
in CH
2
Cl
2
gave a yellow
solution and some red precipitate. Both the neat liquid and the solution decompose in a few hours at
room temperature, turning orange and then dark red. It was not possible to produce a pure sample, and
the freshly made yellow CH
2
Cl
2
solution was used for the spectroscopic measurements.
1
H NMR
(CD
2
Cl
2
, 295 K): δ 1.98 (s); (183 K): 1.86 (s).
11
B NMR (CH
2
Cl
2
/CD
2
Cl
2
, 295 K): δ +5.31 (s); (183
MANUS CRIP T
ACCEP TED
ACCEPTED MANUSCRIPT
5
K): δ −1.52 (s).
19
F{
1
H) NMR (CH
2
Cl
2
/CD
2
Cl
2
, 295 K): δ −144.5 (s); (183 K): δ −148.3 (s).
125
Te
NMR (CH
2
Cl
2
/CD
2
Cl
2
, 295 K): δ −4.80 (s); (183 K): δ −29.2 (s).
After 12 h at room temperature the CH
2
Cl
2
solution had become deep orange with much orange-red
precipitate and had singlet
1
H NMR resonances at 1.92, 2.68 and 5.41, tentatively assigned to free
TeMe
2
and [Me
2
TeCH
2
Cl]
+
(or [Me
2
Te(CH
2
Cl)Cl]), with corresponding
125
Te{
1
H} resonances at 13
and +418. The
19
F{
1
H} NMR showed [BF
4
]
as the major species, with some smaller amounts of F
.
2.8 X-ray experimental
Crystals of the complexes were grown from CH
2
Cl
2
solutions of the complexes allowed to evaporate
slowly in the glove box. Data collections used a Rigaku AFC12 goniometer equipped with an
enhanced sensitivity (HG) Saturn724+ detector mounted at the window of an FR-E+ SuperBright
molybdenum (λ = 0.71073 Å) rotating anode generator with VHF Varimax optics (70 micron focus)
with the crystal held at 100 K. Structure solution and refinement were performed using
SHELX(S/L)97, SHELX-2013 or SHELX-2014/7 [30]. H atoms bonded to C were placed in
calculated positions using the default C−H distance, and refined using a riding model. Details of the
crystallographic parameters are given in Table 1. CCDC reference numbers in cif format are
[BCl
3
(TeMe
2
)]: CCDC 1554690; [BBr
3
(TeMe
2
)]: CCDC 1554688; [BI
3
(TeMe
2
)]: CCDC 1554689;
[BCl
3
(SeMe
2
)]: CCDC 1554686; [BBr
3
(SeMe
2
)]: CCDC 1554687. These data can be obtained free of
charge via http://www.ccdc.cam.ac.uk/conts/retrieving.html, or from the Cambridge Crystallographic
Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: (+44) 1223-336-033; or e-mail:
deposit@ccdc.cam.ac.uk.
MANUS CRIP T
ACCEP TED
ACCEPTED MANUSCRIPT
6
Table 1. X-ray crystallographic data.
a
Compound [BCl
3
(TeMe
2
)] [BBr
3
(TeMe
2
)] [BI
3
(TeMe
2
)] [BCl
3
(SeMe
2
)] [BBr
3
(SeMe
2
)]
Formula C
2
H
6
BCl
3
Te C
2
H
6
BBr
3
Te C
2
H
6
BI
3
Te C
2
H
6
BCl
3
Se C
2
H
6
BBr
3
Se
M 274.85 408.21 549.18 226.19 359.57
crystal system monoclinic monoclinic monoclinic monoclinic monoclinic
Space group (No) P2
1
/m (11) P2
1
/m (11) P2
1
/m (11) P2
1
/m (11) P2
1
/m (11)
a [Å] 5.76470(10) 6.08639(17) 6.6604(2) 5.78171(18) 6.1384(3)
b [Å] 10.8546(2) 11.0548(3) 11.3864(9) 10.6641(2) 10.8664(4)
c [Å] 6.5636(2) 6.69386(18) 6.96470(10) 6.4752(3) 6.6455(3)
α
90 90 90 90 90
β ° 104.676(3) 104.876(3) 105.488(2) 107.398(4) 108.280(5)
γ
°
90 90 90 90 90
U [Å
3
] 397.308(17) 435.29(2) 509.01(4) 380.97(2) 420.90(3)
Z 2 2 2 2 2
µ
(Mo Kα) [mm
1
] 4.645 17.089 11.938 5.868 18.600
total no. reflns 4289 3420 11377 216 324
unique reflns 822 904 468 8451 9534
R
int
0.0163 0.020 0.039 0.035 0.055
no. of params, restraints 38, 0 38, 0 38, 0 38, 0 38, 0
F(000) 252 360 1059 781 876
GOF 1.089 1.074 1.165 1.140 1.104
Largest peak and hole e
3
0.379, -0.235 0.923, -0.623 0.566, -0.741
0.523, -0.249 0.798, -0.498
R
1
b
[I
o
> 2
σ
(I
o
)] 0.010 0.017 0.014 0.016 0.021
R
1
[all data] 0.027 0.044 0.014 0.016 0.023
wR
2
b
[I
o
> 2
σ
(I
o
)] 0.011 0.017 0.035 0.039 0.051
wR
2
[all data] 0.027 0.044 0.035 0.039 0.052
a. Common items: temperature = 100 K; wavelength (Mo-K
α
) = 0.71073 Å; θ(max) = 27.5°.b R
1
= ΣF
o
| F
c
/ ΣF
o
; wR
2
= [Σw(F
o2
- F
c2
)
2
/Σ wF
o4
]
1/2
MANUS CRIP T
ACCEP TED
ACCEPTED MANUSCRIPT
7
3. Results and Discussion.
3.1 [BX
3
(TeMe
2
)] (X = Cl, Br, I)
Although [BX
3
(TeMe
2
)] (X = Cl, Br, I) have been briefly mentioned in larger studies of boron halide
complexes [22,23], they lack detailed characterisation. All three complexes were made by reaction of
the appropriate BX
3
and TeMe
2
in dry n-hexane at 0°C, when they separated as white powders
(Scheme 1).
BX3hexane, 0oC
X = Cl, Br, I
E = S, Se
E*= E, Te
TeMe2
EMe2
hexane
[BX3(EMe2)]
[BF3(E*Me2)] E*Me2
X = F [BX3(TeMe2)]
Scheme 1. Synthesis of the complexes.
The same compounds are produced using excess ligand. The solids slowly darken and decompose at
room temperature and are significantly decomposed after 24 h in solution in chlorocarbons.
Colourless crystals of [BX
3
(TeMe
2
)] (X = Cl, Br, I) were grown by evaporation of CH
2
Cl
2
solutions
in the glove box and the structures are shown in Fig. 1.
MANUS CRIP T
ACCEP TED
ACCEPTED MANUSCRIPT
8
Fig. 1 The structure of [BX
3
(TeMe
2
)] (X = Cl, Br, I) showing the numbering scheme. Ellipsoids are drawn at the
50% probability level and H atoms are omitted for clarity. Symmetry operation: i = x, ½ -y, z. Selected bond
lengths (Å) and angles (°):
X = Cl: Te1−B1 = 2.298(2), Cl1−B1 = 1.8302(13), Cl2−B1 = 1.834(2), C1−Te1−C1
i
= 95.00(9), Cl1−B1−Te1
= 106.29(8), Cl2−B1−Te1 = 108.87(10), Cl2−B1−Cl1 = 111.18(8), Cl2−B1−Cl2
i
= 112.73(11).
X = Br: Te1−B1 = 2.266(4), Br2−B1= 1.986(4), Br1−B1 = 1.996(2), C1−Te1−C1
i
= 95.08(16), Br2−B1−Te1 =
110.60(17), Br2−B1−Br1 = 110.92(12), Br1−B1−Te1 = 106.29(12), Br1−B1−Br1
i
= 111.63(18).
X = I: I1−B1= 2.215(4), I2−B1 = 2.227(2), Te1−B1 = 2.262(4), C1−Te1−C1
i
= 95.84(17), I1−BI−I2 =
111.15(12), I1−B1−Te1 = 111.66(18), I2−B1−I2
i
= 111.12(18), I2−B1−Te1 = 105.75(13).
The crystals are isomorphous and contain tetrahedrally coordinated boron centres with <XBX
slightly greater than the idealised tetrahedral angle and one of the <XBTe slightly smaller. The
d(BX) (Table 2) are very similar to those in the corresponding PMe
3
or AsMe
3
adducts [14,15]. The
d(BE) bonds increase I ~ Br < Cl. The B-X bonds are not significantly affected by the
chalcogenoether present. The <CTeC angles are ~95°, consistent with a large tellurium p-character
in the CTe bonds [31]. Examination of the packing diagrams shows no Te
···
X interactions between
neighbouring molecules within the sum of the van der Waals radii, and hence no hypervalent
interactions at Te.
Table 2 Bond length comparisons.
[BCl
3
(TeMe
2
)]
[BBr
3
(TeMe
2
)]
[BI
3
(TeMe
2
)]
[BCl
3
(SeMe
2
)]
[BBr
3
(SeMe
2
)]
B-E/ Å
2.298(2)
2.266(4)
2.262(4)
2.106(3)
2.088(4)
B-X /Å
1.8302(13)
1.834(2)
1.986(4)
1.996(2)
2.215(4)
2.227(2)
1.8287(15)
1.825(3)
1.987(5)
1.993(3)
The
11
B NMR resonances (
11
B, 80%, I = 3/2, Ξ = 32.084 MHz, Q = 3.55 x 10
30
m
2
, R
c
= 754) move
to high frequency with halide, I < Br < Cl, the same trend as found with Group 15 donor ligands
[13,32]. The
11
B chemical shifts of the telluroether complexes are all much lower frequency than
those of the parent trihalides, reflecting both the change in coordination number and the different
electronic environment at B (Table 3).
Table 3. Multinuclear NMR Data on [BX
3
(ER
2
)]
a,b
.
1
H
11
B
c
19
F
d
77
Se/
125
Te
[BF
3
(SMe
2
)] 2.25 +3.52 −138.8 -
[BF
3
(SeMe
2
)] 2.01 +5.34 −132.8 +1.04
[BF
3
(TeMe
2
)] 1.93 +1.90 −144.2 −4.8
[BCl
3
(SMe
2
)] 2.49 +8.49 - -
MANUS CRIP T
ACCEP TED
ACCEPTED MANUSCRIPT
9
[BCl
3
(SeMe
2
)] 2.36 +7.36 - +170.5
[BCl
3
(TeMe
2
)] 2.10 +4.78 - +182.9
[BBr
3
(SMe
2
)] 2.57 −9.84 - -
[BBr
3
(SeMe
2
)] 2.33 −12.94 - +194.4
[BBr
3
(TeMe
2
)] 2.10 −20.01 - +243.2
[BI
3
(SMe
2
)] 2.56 −67.55 - -
[BI
3
(SeMe
2
)] 2.30 −73.96 - +203.7 (q)
[BI
3
(TeMe
2
)] 2.05 −86.80 - +290.2 (q)
a. all data from CH
2
Cl
2
/CD
2
Cl
2
solution at 295 K. b. BF
3
:
19
F NMR (CD
2
Cl
2
): −127.8;
11
B NMR (CD
2
Cl
2
): BF
3
+11.03; BCl
3
+41.9, BBr
3
+39.5, BI
3
−5.5 from Ref. 34. c. relative to external [BF
3
(OEt
2
)]. d. relative to
external CFCl
3
. e. relative to neat SeMe
2
or TeMe
2
singlets except q = four line pattern. Since the zero
references are SeMe
2
and TeMe
2
, the coordination shifts () in these cases are numerically the same as the
observed chemical shifts.
The
125
Te{
1
H} NMR resonances observed for X = Cl or Br, are singlets with no resolved coupling to
11
B; this can be attributed to a combination of fast ligand exchange in solution and quadrupolar
relaxation of the
11
B in the significant electric field gradients about the boron centre [20,22]. Cooling
the samples stepwise down to 190 K resulted in small changes in the line width and small drifts in
chemical shifts, but did not resolve couplings. In contrast, the
125
Te{
1
H} NMR spectrum of
[BI
3
(TeMe
2
)] at 295 K is a four line pattern with the
11
B coupling clearly resolved (Fig. 2). Couplings
to the
10
B (I = 3, Ξ = 10.75 MHz, Q = 7.4 x 10
30
m
2
) were not resolved due to its larger quadrupole
moment, and account for the broad absorption underlying the four line pattern. The coordination shifts
are large and increase Cl < Br < I, suggesting increasingly strong BTe interaction in this order,
although care should be taken in interpreting such changes as due to a single effect [32,33].
Fig. 2 The
125
Te NMR spectrum of [BI
3
(TeMe
2
)] in CH
2
Cl
2
at 295K showing the
11
B
125
Te coupling,
1
J = 120
Hz. The broad feature under the resonance is due to unresolved
10
B couplings.
Chemical Shift (ppm)305 300 295 290 285
MANUS CRIP T
ACCEP TED
ACCEPTED MANUSCRIPT
10
3.2 [BX
3
(SeMe
2
)] (X= Cl, Br, I)
The dimethylselenide complexes are white powders and unlike the telluroether analogues (Scheme 1),
are stable for many weeks in the solid state and show no decomposition in chlorocarbon solution after
several days. X-ray structures were determined for [BX
3
(SeMe
2
)] (X = Cl or Br) (Fig. 3) and again
are isomorphous. The trends in bond length with X and E are as seen in the telluroether complexes.
Fig. 3
The structure of [BX
3
(SeMe
2
)] (X = Cl, Br) showing the numbering scheme. Ellipsoids are drawn at the
50% probability level and H atoms are omitted for clarity. Symmetry operation: i = x, ½ -y, z. Selected bond
lengths (Å) and angles (°):
X = Cl: B1−Se1 = 2.106(3), B1−Cl2 = 1.8287(15), B1−Cl1 = 1.825(3), Cl2−B1−Se1 = 105.30(10),
Cl2−B1−Cl2
i
= 112.58(14), Cl1−B1−Se1 = 109.57(13), Cl1−B1−Cl2 = 111.82(10), C1−Se1−C1
i
97.93(11).
X = Br: B1−Se1 = 2.088(4), B1−Br2 = 1.987(5), B1−Br1 = 1.993(3), Br2−B1−Se1 = 110.7(2), Br1−B1−Br1
i
=
112.1(2), Br1−B1−Se1 = 105.14(14), Br1−B1−Br2 = 111.66(14), C1−Se1−C1
i
98.3(2).
The
11
B NMR chemical shifts occur at slightly higher frequency than those found in the telluroether
complexes for common X (Table 3) and the trend to higher frequency Cl > Br > I is observed. The
77
Se NMR spectra also show coordination shifts with X (Cl < Br < I). At 295 K the
77
Se NMR
resonances of [BX
3
(SeMe
2
)] (X = Cl or Br) are broad singlets, but that of [BI
3
(SeMe
2
)] (Fig. 4) shows
a sharp four line pattern due to coupling to
11
B,
1
J
BSe
= 52 Hz. Cooling a CH
2
Cl
2
solution of
[BBr
3
(SeMe
2
)] to 183 K caused the singlet resonance to significantly broaden, but even at this
temperature
11
B
77
Se coupling was not resolved.
MANUS CRIP T
ACCEP TED
ACCEPTED MANUSCRIPT
11
Fig 4. The
77
Se NMR spectrum of [BI
3
(SeMe
2
)] in CH
2
Cl
2
at 295 K showing the
11
B
77
Se coupling,
1
J = 52 Hz.
3.3 [BX
3
(SMe
2
)] (X= Cl, Br, I)
Multinuclear NMR data for these three complexes are given in Table 3. The data are generally in
good agreement with literature data [18,21,22,23] when allowance is made for the use of different
solvents and sometimes different boron zero references. The trends in chemical shift with X mirror
those described above.
3.4 [BF
3
(EMe
2
)] (E = Te, Se, S)
The BF
3
adducts are very different to those formed by the other boron halides. The commercially
available [BF
3
(SMe
2
)], which is a convenient source of BF
3
in organic synthesis,
is a colourless oil,
which fumes in air and has a significant vapour pressure at 298 K [16,18]. The
1
H,
11
B and
19
F NMR
data (Table 3) obtained from CH
2
Cl
2
solution at 295 K are in good agreement with literature data [20].
Cooling the solution results in small low temperature drifts of the chemical shifts and below ~ 200 K
the resonances broaden, but even at 183 K,
11
B-
19
F coupling is not resolved. [BF
3
(SeMe
2
)] was
obtained as a very pale yellow liquid by saturating neat SeMe
2
in an ice-bath with BF
3
. It is more
volatile and more extensively dissociated than the SMe
2
analogue [16], and has a significant vapour
pressure of BF
3
at ambient temperature. The
1
H,
11
B and
19
F NMR data (Table 3) obtained from
CH
2
Cl
2
solution at 295 K are similar to those observed for [BF
3
(SMe
2
)]. The
77
Se chemical shift of
[BF
3
(SeMe
2
)] in CH
2
Cl
2
solution is very different to those found with the other three boron halides.
At ambient temperatures it is a broad singlet at ~ +1 ppm, which is little changed on cooling the
solution to 183 K. It is worth pointing out here that the zero reference is neat SeMe
2
(δ = 0), and that
SeMe
2
in CH
2
Cl
2
solution has (δ ~ −7) [34], so the observed value is some ~ 8 ppm to high frequency
of the free selenoether in this solvent; nonetheless a very small coordination shift. The most obvious
explanation is that the SeMe
2
is very weakly bound to the BF
3
centre and fast neutral ligand exchange
may be taking place even at the lowest temperatures. The [BF
3
(SeMe
2
)] fumes in air and is very
MANUS CRIP T
ACCEP TED
ACCEPTED MANUSCRIPT
12
sensitive to moisture which generates [BF
4
]
and a species with δ (
1
H) = 2.59 and δ (
77
Se) = 247,
which is tentatively identified as [Me
3
Se]
+
(lit. δ (
77
Se) = 253 (H
2
O) [34], 259 (acetone) [35]).
The only report [28] found of the reaction of BF
3
with TeMe
2
described “immediately upon addition
of BF
3
the solution assumed a brilliant red hue and a volatile orange coloured sublimate quickly
formed”. We observed that on passing BF
3
gas into TeMe
2
in an ice bath, a yellow liquid with small
amounts of orange-red solid formed. A yellow solution and some red solid formed on conducting the
reaction in anhydrous CH
2
Cl
2
solution. The amount of red solid increases over time, and both the
liquid and the CH
2
Cl
2
solution seem extensively decomposed after 12 h at room temperature. The
multinuclear NMR data (Table 3) suggest the yellow liquid is [BF
3
(TeMe
2
)], but this has not been
obtained pure, always seeming to contain some orange impurity. The
11
B and
19
F NMR spectra are not
dissimilar to those obtained from the SMe
2
and SeMe
2
adducts and are little affected by cooling the
solution to 183 K. The
125
Te{
1
H} NMR spectrum in CH
2
Cl
2
at 295 K is a sharp resonance at δ = −4.8,
which shifts to δ = −29 at 183 K. Like the case of the selenium analogue discussed above, the
chemical shift of free TeMe
2
is solvent and temperature sensitive [34] and in CH
2
Cl
2
solution at 295
K has δ = −15 relative to the zero reference of neat TeMe
2
. A CH
2
Cl
2
solution which had become
deep orange with much orange precipitate after ~20 h at room temperature showed new resonances
identified as free TeMe
2
, [BF
4
]
, and possibly [Me
2
Te(CH
2
Cl)Cl] (see Section 2.9). The
[Me
2
Te(CH
2
Cl)Cl] (or possibly[Me
2
Te(CH
2
Cl)][BF
4
] is formed by the quaternisation of the
telluroether by the solvent, promoted by the strong Lewis acidic BF
3.
We have observed similar Lewis
acid promoted quaternisation of chalcogenoether ligands in gallium and aluminium systems [36,37].
3.5 Comparisons and Conclusions
The X-ray crystallographic data described for the [BX
3
(EMe
2
)] (E = Se or Te; X = Cl, Br, I) above
(Table 2) shows systematic trends with d(BE) falling with X, Cl > Br > I, the same order as found
with most other neutral donor systems, and consistent with BI
3
being the strongest Lewis acid [3,5].
Although the X-ray data are less complete, the heavier elements of Group 13 (Al, Ga and In) show the
reverse trends, with the metal chloride forming the shortest ME bonds [37-40]. The multinuclear
NMR data (Table 3) also show some consistent trends, for example the
11
B chemical shifts for a fixed
chalcogenoether move to low frequency Cl Br I, whilst for a fixed halide the trend to low
frequency is S Se Te. For complexes of these three boron halides with SeMe
2
the
77
Se
coordination shifts ( = δ
complex
δ
ligand
) are also large and positive (to high frequency) with I > Br
> Cl. The
125
Te coordination shifts follow the same order. In many series of organoselenium and
organotellurium compounds the ratio δ(Te)/δ(Se) ~ 1.8 is observed [33,34,41], and this empirical
observation also holds for some d-block metal complexes in medium oxidation states [33]. However,
in low valent organometallic or carbonyl complexes the ratio of the coordination shifts are much
larger, δ(Te)/δ(Se) > 2.3, rationalised as due to greater R
2
TeM donation in the soft metal centres
MANUS CRIP T
ACCEP TED
ACCEPTED MANUSCRIPT
13
and with expanded metal d-orbitals resulting from the low oxidation states providing good orbital
overlap with Te centres [33,42-44]. In the case of the boron halide adducts the opposite trend is
observed with δ(Te)/δ(Se) ~ 1.11.4, consistent with the stronger interaction being BSeMe
2
.
The chalcogenoether adducts of BF
3
are clearly different to those with the other boron halides, and
whilst this work has shown that [BF
3
(TeMe
2
)] does exist, it is unstable and could not be obtained
pure. The
11
B NMR spectra show the chemical shifts move to low frequency Cl > F > Br > I for a
fixed EMe
2
, the anomalous position of the fluoride also being observed with Group 15 donor ligands
[3,13]. The very small coordination shifts observed in the
77
Se and
125
Te NMR spectra of the fluoride
complexes also suggest weak interactions. Overall the data confirm the trends in Lewis acidity of the
boron halides predicted by recent theoretical studies [5-9].
3.5 Acknowledgements
C.K.Y.A.O. thanks the Universidad Nacional de Colombia for authorising sabbatical leave. We also
thank the EPSRC (EP/ L505651/1, EP/K039466/1) and the Newton Fund for support.
Supplementary Information.
ESI for this work including X-ray crystallographic data, and multinuclear NMR spectra may be found
at DOI:…..
MANUS CRIP T
ACCEP TED
ACCEPTED MANUSCRIPT
14
References
[1] A. Staubitz, A. P. M. Robertson, I. Manners, Chem. Rev. 110 (2010) 4079.
[2] A. Staubitz, A. P. M. Robertson, M. E. Sloan, I. Manners, Chem. Rev. 110 (2010) 4023.
[3] J. Burt, W. Levason, G. Reid, Coord. Chem. Rev. 260 (2014) 65.
[4] N. A. Young, Coord. Chem. Rev. 257 (2013) 956.
[5] I. B. Sivaev, V. I. Bregadze, Coord. Chem. Rev. 270-271 (2014) 75.
[6] N. S. Hosmane (Ed), Boron Science: New technologies and applications, CRC Press, Florida, 2012.
[7] F. Bessac, G. Frenking, Inorg. Chem. 42 (2003) 7990.
[8] F. Bessac, G. Frenking, Inorg. Chem. 45 (2006) 6956.
[9] J. A. Plumley, J. D. Evanseck, J. Phys. Chem. A, 113 (2009) 5985.
[10] A. M. McKinty, C. Lund, D. W. Stephan, Organometallics, 32 (2013) 4730.
[11] D. W. Stephen, Nat. Chem. 6 (2014) 952.
[12] K. Chansaenpak, B. Vabre, F. P. Gabbaï, Chem. Soc. Rev. 45 (2016) 954.
[13] J. Burt, J. W. Emsley, W. Levason, G. Reid, I.S. Tinkler, Inorg. Chem. 55 (2016) 8852.
[14] D. L. Black, R. C. Taylor, Acta Cryst. Sect. B, 31 (1975) 1116.
[15] R. K. Chadha, J. M. Chehayber, J. E. Drake, J. Cryst. Spectros. Res. 15 (1985) 53.
[16] W.A.G. Graham, F. G. A. Stone, J. Inorg. Nucl. Chem. 3 (1956) 164.
[17] H. L. Morris, N. I. Kulevsky, M. Tamres, S. Searles Jr., Inorg. Chem. 5 (1966) 124.
[18] H. C. Brown, N. Ravindran, Inorg. Chem. 16 (1977) 2938.
[19] K. Kinberger, W. Siebert, Z. Naturforsch. Teil B. 30 (1975) 55.
[20] M. J. Bula and J. S. Hartman, J. Chem. Soc., Dalton Trans. (1973) 1047.
[21] H. Fuβstetter, H.Nöth, B. Wrackmeyer, W. McFarlane, Chem. Ber. 110 (1977) 3172.
[22] R. Conrady, W. Müller-Warmuth, B. Krebs, G. Schwetlik, M. Wienkenhöver, Ber. Bunsenges. Phys.
Chem. 91 (1987) 1322.
[23] M. Schmidt, H. D. Block, Chem. Ber. 103 (1970) 3705.
[24] J. Le Calve, J. Lascombe, Spectrochim. Acta 24A (1968) 737.
[25] P. Labarbe, M. T. Forel, Spectrochim. Acta 31A (1975) 525.
[26] B. Krebs, G. Schwetlik, M. Wienkenhöver, Acta Crystallogr. B45 (1989) 257.
[27] M. Wienkenhover, A. Vahrenhorst, M. Tatermann, B. Krebs, Eur. Cryst. Meeting 9 (1985) 154. CCDC-
codes DIMCEH, DIMCIL, DIMCOR, DIMCAO, DIMBUW.
MANUS CRIP T
ACCEP TED
ACCEPTED MANUSCRIPT
15
[28] A. A. Palko, J. S. Drury, J. Chem. Phys. 46 (1967) 2297.
[29] N. Kuhn, P. Faupel, E. Zauder, J. Organomet. Chem. 302 (1986) C4.
[30] G. M. Sheldrick, Acta Crystallogr. Sect. A 64 (2008) 112.
[31] W. Levason, G. Reid, in Comprehensive Coordination Chemistry II, J. A. McCleverty, T.J. Meyer (eds),
Elsevier Oxford 1 (2004) 391.
[32] H. Nöth, B. Wrackmeyer, Nuclear magnetic resonance spectroscopy of boron compounds, Springer-Verlag,
N.Y. 1978.
[33] W. Levason, S. D. Orchard, G. Reid, Coord. Chem. Rev. 235 (2002) 159.
[34] N. P. Luthra, J. D. Odom, in The chemistry of organic selenium and tellurium compounds, S. Patai and Z.
Rappoport (Eds), Wiley, NY. 1, (1986) chapter 6.
[35] W. Levason, L. P. Ollivere, G. Reid, M. Webster, J. Organometal. Chem. 695 (2010) 1346.
[36] K. George, M. Jura, W. Levason, M E. Light, L. P. Ollivere, G. Reid
.
Inorg. Chem. 51 (2012) 2231.
[37] K. George, M. Jura, W. Levason, M. E. Light, G. Reid, Dalton Trans. 43 (2014) 3637.
[38] C. Gurnani, W. Levason, R. Ratnani, G. Reid, M. Webster, Dalton Trans. (2008) 6274.
[39] S. Mishra, E. Jeanneau, S. Daniele, Polyhedron 29 (2010) 500.
[40] C. Gurnani, M. Jura, W. Levason, R. Ratnani, G. Reid, M. Webster, Dalton Trans. (2009) 1611.
[41] T. Kemmitt, W. Levason, Organometallics 8 (1989) 1303.
[42] N. Kuhn, H. Schumann, E. Zauder, J. Organometal. Chem. 327 (1987) 17.
[43] A. J. Barton, W. Levason, G. Reid, J. Organometal. Chem. 579 (1999) 235.
[44] W. Levason, S. D. Orchard, G. Reid, J. M. Street, J. Chem. Soc., Dalton Trans. (2000) 2537.
Highlights
Synthesis of boron halide-dimethyltelluride complexes
X-ray structures of TeMe
2
and SeMe
2
complexes of boron halides
Multinuclear NMR data on boron halide-chalcogenoethers
Comparisons of S, Se and Te ligand complexes with boron halides
MANUS CRIP T
ACCEP TED
ACCEPTED MANUSCRIPT
16
MANUS CRIP T
ACCEP TED
ACCEPTED MANUSCRIPT
Highlights
Synthesis of boron halide-dimethyltelluride complexes
X-ray structures of TeMe
2
and SeMe
2
complexes of boron halides
Multinuclear NMR data on boron halide-chalcogenoethers
Comparisons of S, Se and Te ligand complexes with boron halides
... It should be noted that this approach could also be used to assess the Lewis acidity of other boron compounds, for example, boron halides [60], where it gives results comparable to the widely used Gutmann-Beckett method [37]. ...
Article
Experimentally observed Lewis acidity order of the boron trihalides (BI3 > BBr3 > BCl3 > BF3) towards majority of the bases barring the very weak type (CO, HCN, CH3CN, CH3F etc.) is an abiding theme of chemical interest over several decades. The mainstream explanation exclusively invoked in the chemistry textbooks, is the π-type back-donation from filled pπ orbital of the halogen into the vacant pπ orbital of boron. A plethora of computational studies have been performed at various levels of sophistication (LCAO MO to DFT) from the late 1950s till date, in order to obtain a theoretical insight into the experimentally observed Lewis acidity order. Theoretical investigations uniformly suggest an insignificant role of the π-type back donation, rather the reorganisation energy of boron halides appears to be the parameter of profound impact. The pliability of the molecular acids and subsequent boron-halogen σ-bond elongation upon Lewis base coordination, have been authenticated as the prime factors in influencing the reorganisation energy. However, out-of-plane deformation of the ground state structure of boron halide to the geometry resembling the adduct, primarily accounts for the reorganisation energy. The origin and magnitude of the reorganisation energy have been probed from the standpoint of energy partitioning analysis (EPA). Quantitative Lewis acidity scales developed so far have been also discussed on the basis of NMR spectral features and quantum-chemically derived ion affinity calculations. Recently Parr's electrophilicity index for the boron halides have been computed, which can serve as a base-free metric to address the inherent Lewis acidity. The electrophilicity index and various ion affinity parameters have been found to bear well-behaved linear relations. Eluding the conventional spectral characterisation data related to the much explored N-donor ligands, we have herein purposely reported the X-ray structural features of the boron trihalide adducts with the trialkyl substituted phospines, arsines and stibines (π-acceptor ligands) as well as with the seleno and telluro ethers (π-donor ligands). The documented donor-boron bond length values in all the complexes have been found to attest the established acidity order unambiguously. Tetrahedral character (THC) of all those complexes have been calculated with the formula proposed by Hopfl and it reflects substantial covalent character in the boron-donor (P/As/Sb/Se/Te) bonds. This work primarily encompasses the theoretical investigations almost exclusively in a chronological way to provide a comprehensive understanding of underlying principles of the observed acidity order. After a rigorous computational research period of more than sixty years, it can be unequivocally inferred that the use of widely advocated π-type back-donation concept has been relegated over the passage of time. Instead, benchmark computational studies indicate that the electron accepting property should be interpreted in terms of polarisability, electron affinity and energy of LUMO of the boron halides. The subtle interplay of electronegativity and radius of the bonded halogen atom also contributes to the Lewis acidic order. Several theoretical models, such as, ligand close packing (LCP), minimum electrophilicity principle (MEP) and activation strain model (ASM), have emerged to treat the ruling factors of Lewis acidity in a quantitative way. So, the boron-halogen σ-bonding effects have practically outweighed the long-standing concept of π-type back-donation and nowadays the latter is discounted.
Article
In wake of emerging applications of organotellurium compounds in biological and material science avenues, the current review describes their key synthetic methodologies while focusing the synthesis of organotellurium compounds through five ligand-to-metal linkages including carbon; carbon-oxygen; carbon-nitrogen; carbon-metal; carbon-sulfur to tellurium. In all of these linkages whether tellurium links with ligands through a complicated or simple pathways, it is often governed through electrophilic substitution reactions. The present study encompasses these major synthetic routes so as to acquire comprehensive understanding of synthetic organotellurium compounds.
Article
Full-text available
Density functional theory calculations have been performed to elucidate the electronic structure and bonding scenario in various carbene‐borane (LBX₃) and carbene‐alane (LAlX₃) compounds (X = ‐H, ‐Me, ‐Cl, ‐Ph, ‐C6F5). We have performed extended transition state (ETS) analysis to reveal the nature of the donor–acceptor bonds (Ccarb–E; E = B, Al) and also for the assessment of the intrinsic donor–acceptor strength in this class of compounds. Our computations suggest that the Ccarb–Al bonds in all the LAlX₃ adducts have substantially higher electrostatic nature than covalent character. Conversely, the nature of the Ccarb–B bonds in LBX₃ have a strong dependence on the electronic nature of both carbene and borane. Moreover, unlike alanes in LAlX₃, the intrinsic Lewis acid strength of the boranes in LBX₃ has a strong dependence on the electronic nature of the carbenes. We have also explored the correlation of the interaction energies (ΔEint) with various bonding parameters i.e., geometrical and Natural bond orbital (NBO) parameters of LEH₃. Furthermore, natural orbital for chemical valence (NOCV) calculations are performed to have a qualitative understanding of the relative σ‐donating and π‐accepting abilities of the carbenes in LEH₃. Additionally, we have investigated the abilities of carbenes to activate the B–H bonds in BH3 and pinacolborane. ETS analysis shows a strong dependence of the B–H activation barriers on the distortion energies of both carbene and borane fragments.
Article
Full-text available
Most high and medium oxidation state (O.S. ≥ 3) metal and non-metal fluorides and oxide fluorides have Lewis acidic properties although detailed exploration of their chemistry with neutral ligands, which differs significantly from that with chloride, bromide or iodide co-ligands, has only been undertaken in recent years. The previous review (Benjamin et. al. Chem. Soc. Rev. 42 (2013) 1460) covered work published up to ~ 2011, and the present article covers new work up to early 2019, a period which has seen many new contributions to the field. This article describes work on the coordination chemistry of d, f and p-block fluorides and oxide fluorides with neutral ligands containing donor atoms drawn from Groups 15 (N, P, As or Sb) and 16 (O, S, Se or Te) and including N-heterocyclic carbenes. The incorporation of the radionuclide 18F into neutral metal complexes and their use in medical diagnosis via positron emission tomography (PET) is described, along with briefer coverage of other potential applications.
Article
Silylenes have become an indispensable tool for molecular bond activation. Their use for the construction of silicon-boron bonds is uncommon in comparison to the numerous studies on silylene-derived silicon-element bond formations. Herein we investigate the reactivity of the pyramidal NHC-coordinated hydrosilylene tBu3SiSi(H)LMe4 (1; NHC = N-heterocyclic carbene, LMe4 = 1,3,4,5-tetramethylimidazolin-2-ylidene) with various boron-centered electrophiles. The reaction of 1 with THF·BH3 or H3N→BH3 afforded the silylene complex 1→BH3 or the product of insertion of the silicon(II) atom into an N–H bond with concomitant dehydrogenation along the HN–BH moiety (2). The respective conversion of 1 with BPh3 yields 1→BPh3 which readily reacts with excess LMe4 to form the more stable complex LMe4→BPh3 with release of 1. Treatment of 1 with the haloboranes Et2O→BF3, BCl3, BBr3 and Me2S→BBr3 resulted in the formation of the Lewis acid-base adducts 1→BX3 (X = F, Cl, Br) and an equilibrium with their auto-ionization products [12BX2]+[BX4]− slowly develops. The ratio of 1→BX3 significantly increases with rising atomic number of the halide, thus 1→BF3 majorly transforms within hours while 1→BBr3 is near-quantitatively retained over time. Accordingly, the complex 1→BPhBr2 was isolated after conversion of 1 with PhBBr2.
Article
The stability and properties of donor-acceptor adducts of square-planar Pt(II) and Ir(I) complexes (designated as PtX, IrX, or generally MX complexes) with trihydrides and trihalides of group 13 elements of general formula YZ3 (Y = B, Al, Ga; Z = H, F, Cl, Br) were studied theoretically using DFT methodology in the gas phase. MX complexes were represented by wide range of the ligand environment which included model complexes [Ir(NH3)3X]0 and cis-[Pt(NH3)2X2]0 (X = H, CH3, F, Cl, Br) and isoelectronic complexes [Ir(NNN)(CH3)]0 and [Pt(NCN)(CH3)]0 with tridentate NNN and NCN pincer ligands. MX complexes acted as the Lewis bases donating electron density from the doubly occupied 5d z2 atomic orbital of the metal M atom to the empty valence p z orbital of Y whose evidence was clearly provided by the natural atomic orbital (NAO) analysis. This charge transfer led to the formation of pentacoordinated square pyramidal MX·(YZ3) adducts with M·Y dative bond. Binding energies were -44.7 and -75.2 kcal/mol for interaction of GaF3 as the strongest acid with PtNCN and IrNNN pincer ligands complexes. Only M·B bonds had covalent character although MX·BZ3 adducts were the least stable due to large values of Pauli repulsion and deformation energies. The highest degree of covalent character was found for adducts of BH3 in all series of structures studied. Al and Ga adducts showed remarkably similar behavior with respect to geometry and binding energies.
Article
The reaction of BX3 (X = Cl, Br, I) with SbR3 (R = Et,ⁱPr) in anhydrous hexane or toluene produced moisture-sensitive [BX3(SbR3)], whose stability increases with halide: Cl < Br < I. Unstable [BF3(SbR3)] species were also characterized spectroscopically but lose BF3 very readily. The [MX3(SbR3)] (M = Al, Ga, In) complexes were isolated from hexane solution; the Ga and In complexes can be recrystallized unchanged from CH2Cl2, but [AlX′3(SbR3)] (X′ = Br, I) undergoes X′/Cl exchange with the solvent. The new complexes were characterized by IR and multinuclear NMR spectroscopy (¹H, ¹³C{¹H}, ¹¹B, ²⁷Al, ⁷¹Ga, ¹¹⁵In, as appropriate). X-ray crystal structures are reported for [BX3(SbR3)] (X = Br, I; R = Et, ⁱPr; X = Cl, R = ⁱPr), [AlI3(SbⁱPr3)], [GaX3(SbR3)] (X = Cl, I; R = ⁱPr), [GaBr3(SbEt3)], and [InX3(SbR3)] (X = Cl, I; R = Et, ⁱPr); all contain pseudotetrahedral geometries at both the group 13 and antimony atoms, and comparisons with phosphine and arsine analogues are discussed. The donor–acceptor bond lengths are unexceptional, but coordination of the stibine is accompanied by a significant widening of the C–Sb–C angles and a small reduction in C–Sb distances. DFT calculations confirm these results from significant changes in the antimony 5s and 5p contributions to the Sb–C bonds, with corresponding increases in the 5p character of the antimony-based “lone pair” on coordination. Intermolecular hypervalent Cl···Sb interactions are present in the two [InCl3(SbR3)] complexes but absent in the others.
Article
Full-text available
The complexes [BX3(μ-L−L)BX3] (X = Cl, Br, I; L−L = EtS(CH2)2SEt, MeSe(CH2)2SeMe), [BX′3{o-C6H4(SMe)2}] (X’ = Cl, I), [(BBr3)2{o-C6H4(SMe)2}], [(BBr3)2{o-C6H4(SeMe)2}] and [BI3{o-C6H4(SeMe)2}] have been prepared as moisture-sensitive pale solids by reaction of the appropriate BX3 with the dichalcogenoether in anhydrous n-hexane solution, and characterised by microanalysis, IR and multinuclear (¹H, ¹¹B, ⁷⁷Se{¹H}) NMR spectroscopy. In contrast, the [BF3(μ-L−L)BF3], [(BF3)2{o-C6H4(SMe)2}] and [(BF3)2{o-C6H4(SeMe)2}], made from BF3 and the neat ligands, are viscous oils which have a significant vapour pressure of BF3 at ambient temperatures. X-ray crystal structures are reported for [BX3{μ-EtS(CH2)2SEt}BX3] (X = Cl, Br, I), [BBr3{μ-MeSe(CH2)2SeMe}BBr3], [BCl3{o-C6H4(SMe)2}] and [(BBr3)2{μ-o-C6H4(SeMe)2}]. The complexes [(BX3)2{MeTe(CH2)3TeMe}] (X = F, Cl, Br) have been identified in solution by multinuclear NMR spectroscopy, but decompose rapidly, whilst o-C6H4(TeMe)2 decomposes immediately on contact with BBr3 or BCl3. Dealkylation of some of the chalcogenoether ligands at room temperature by BI3, to yield complexes including [BI2{o-C6H4S(SMe)] and [BI2{o-C6H4Se(SeMe)], has been identified and the X-ray structure of [BI2{o-C6H4Se(SeMe)] determined. The trends in behaviour along the series of boron halides and with the various chalcogenoethers are described and compared with the behaviour of BX3 with neutral phosphorus and arsenic donor ligands (Burt et al., Inorg. Chem., 2016, 55, 8852) and with [BX3(EMe2)] (E = S, Se, Te) (Okio et al., J. Organometal. Chem., 2017, 848, 232).
Article
Full-text available
The reaction of AlCl3 with Me2E (E = S, Se or Te) or (n)Bu2E (E = Se or Te) in CH2Cl2 under rigorously anhydrous conditions gave the pseudo-tetrahedral complexes [AlCl3(R2E)]. The [AlX3(Me2E)] (X = Br or I, E = S; X = Br, E = Te) were made from toluene solution since attempted syntheses in CH2Cl2 resulted in substantial chloride incorporation. The synthesis of [(AlCl3)2{o-C6H4(CH2SEt)2}], in which the ligand bridges two tetrahedral aluminium centres, and of the six-coordinate trans-[AlX2{MeE(CH2)2EMe}2][AlX4] (X = Cl or Br, E = S, and X = Cl, E = Se) and cis-[AlI2{MeS(CH2)2SMe}2][AlI4] are reported. The tripodal thioether forms [AlCl3{MeC(CH2SMe)3}], which is a chain polymer with κ(2)-coordinated ligand and a tbp arrangement at Al(iii). Chalcogenoether macrocycle complexes [AlCl3([9]aneS3)], [AlCl2([14]aneS4)][AlCl4] and [AlCl2([16]aneSe4)] [AlCl4] are also described. All complexes were characterised by microanalysis, IR and multinuclear NMR ((1)H, (27)Al, (77)Se or (125)Te) spectroscopy as appropriate. In CH2Cl2 solution [AlCl3(Me2S)] with added Me2S forms [AlCl3(Me2S)2], and the [AlX2{MeS(CH2)2SMe}2][AlX4] exist as mixtures of cis and trans isomers which undergo rapid exchange at ambient temperatures. X-Ray crystal structures are reported for [AlCl3(Me2Se], [AlX3(Me2Te)] (X = Cl or Br), trans-[AlCl2{MeE(CH2)2EMe}2][AlCl4] (E = S or Se), cis-[AlI2{MeS(CH2)2SMe}2][AlI4], [AlCl3{MeC(CH2SMe)3}], and for the sulfonium salt [Me2SH][AlCl4]. The aluminium halide chalcogenoether chemistry is compared with the corresponding gallium and indium systems, and the relative Lewis acidities of the metals discussed. Attempts to use [AlCl3((n)Bu2E)] (E = Se or Te) as LPCVD reagents to form aluminium chalcogenide films were unsuccessful.
Article
Full-text available
The coordination chemistry of copper, gallium and indium halides with the simplest possible selenoether i.e. Me2Se was investigated with the aim to use the resulting complexes as precursors for selenium-containing chalcopyrite semiconducting materials. An optimized general procedure for the high yield synthesis is described and the influence of a halide ion on the structure and solubility of these metal halide dimethyl selenide complexes are discussed. These complexes were characterized by the elemental analysis, FT-IR and 1H NMR spectroscopy as well as single crystal X-ray structures, the later study showing them to be monomeric for gallium halides, mono- or dimeric for indium halides and either an ion-pairs or 2-D extended structure in the case of copper halides.
Article
The coordination chemistry of the neutral diphosphines, R2P(CH2)2PR2 (R = Me or Et) and o-C6H4(PR'2)2 (R' = Me or Ph), and the diarsine, o-C6H4(AsMe2)2, toward the Lewis acidic BX3 (X = F, Cl, Br, and I) fragments is reported, including several rare complexes incorporating BF3 and BF2(+). The studies have revealed that the flexible dimethylene linked diphosphines form [(BX3)2{μ-R2P(CH2)2PR2}] exclusively, confirmed by multinuclear NMR ((1)H, (11)B, (19)F{(1)H}, and (31)P{(1)H}) and IR spectroscopy and microanalytical data. Crystallographic determinations of the four BX3 complexes with Et2P(CH2)2PEt2 confirm the 2:1 stoichiometry and, taken together with the spectroscopic data, reveal that the Lewis acid behavior of the BX3 fragment toward phosphine ligands increases in the order F ≪ Cl ∼ Br < I. The first diphosphine- and diarsine-coordinated dihaloboronium cations, [BX2{o-C6H4(EMe2)2}](+) (E = P, As), are obtained using the rigid, preorganized o-phenylene linkages. These complexes are characterized similarly, and the data indicate that the complexes with o-C6H4(AsMe2)2 are much more labile and readily decomposed than the phosphine analogues. X-ray crystallographic studies on [BX2{o-C6H4(PMe2)2}][BX4] (X = Cl, Br), [BI2{o-C6H4(PMe2)2}][I3], and [BCl2{o-C6H4(AsMe2)2}][BCl4] confirm the presence of distorted tetrahedral coordination at boron through a chelating diphosphine or diarsine and two X ligands, with d(B-P) revealing a similar increase in Lewis acidity down group 17. Comparison of d(B-P) and d(B-As) reveals an increase of ca. 0.08 Å from P to As. Reaction of BCl3 with the diphosphine dioxide Ph2P(O)CH2P(O)Ph2 gives the ligand-bridged dimer [(BCl3)2{Ph2P(O)CH2P(O)Ph2}], while using either BF3 gas or [BF3(SMe2)] gives a mixture containing both [(BF3)2{μ-Ph2P(O)CH2P(O)Ph2}] and the unexpected difluoroboronium salt, [BF2{Ph2P(O)CH2P(O)Ph2}][B2F7] containing a chelating phosphine oxide. The structure of the latter was confirmed crystallographically.
Article
The field of (18)F chemistry is rapidly expanding because of the use of this radionuclide in radiotracers for positron emission tomography (PET). Until recently, most [(18)F]-radiotracers were generated by the direct attachment of (18)F to a carbon in the organic backbone of the radiotracer. The past decade has witnessed the emergence of a new strategy based on the formation of an (18)F-group 13 element bond. This approach, which is rooted in the field of fluoride anion complexation/coordination chemistry, has led to the development of a remarkable family of boron, aluminium and gallium [(18)F]-fluoride anion complexing agents which can be conjugated with peptides and small molecules to generate disease specific PET radiotracers. This review is dedicated to the chemistry of these group 13 [(18)F]-fluorides anion complexing agents and their use in PET. Some of the key fluoride-binding motifs covered in this review include the trifluoroborate unit bound to neutral or cationic electron deficient backbones, the BF2 unit of BODIPY dyes, and AlF or GaF3 units coordinated to multidentate Lewis basic ligands. In addition to describing how these moieties can be converted into their [(18)F]-analogs, this review also dicusses their incorporation into bioconjugates for application in PET.
Chapter
Considering the large number of classes of compounds a number of approaches are possible for organizing the material covered in this book. The following problems had to be solved (i) to list 11 B NMR data for more than 3000 compounds (ii) to relate the contents of the tables as closely as possible with the discussion in the preceding chapters (iii) to provide an easy access to the NMR information of certain classes of compounds related by their structure and substituents at boron. The approach finally chosen is not systematic in a strict sense, but we believe it is practical both for the preparative chemist as well as for those scientists who wish to obtain information about 11 B chemical shifts and structures of boron compounds. Thus we provide structural symbols at the head of each table as a guideline. These symbols show the atoms to which the boron atom is directly connected. In addition the headlines of each table indicate the classes of compounds treated. As a consequence compounds with the same structural unit, e.g., CBN2, will be found not only in one table, since this structural unit will be met in bis(amino)organylboranes, borazines, tetraazadiborines etc. Moreover it was necessary to subdivide several tables in order to keep their lengths within a manageable size. A subdivision according to different substituents at boron and different structures (cyclic, noncyclic) was also necessary. The term “substituted” in the headlines of the tables is used to indicate the presence of an organometallic or complex organic group bound either to boron or to the atom next to boron, e.g., carbon, oxygen, nitrogen; however, this term has been omitted when the overall number of compounds per table was small.
Article
The Ru dithiolate alkylidene species (SIMes)Ru(CHPh)(O(CH2CH2S)2) (1) reacts with 1 and 2 equiv of BCl3 to give (SIMes)RuCl(CHPh)(Cl2B(SCH2CH2)2O) (2) and [(SIMes)Ru(CHPh)(Cl2B(SCH2CH2)2O)][BCl4] (3), respectively. Compounds 1 and 2 are inactive in ROMP, RCM, and CM reactions, whereas 3 is an active catalyst for these metatheses.
Article
The effects of different electronic and structural factors in determining the Lewis acidity of boron compounds are analyzed. Scales of Lewis acidity for boron Lewis acids based on the Gutmann–Beckett and Childs methods have been constructed using data available in the literature. The Lewis acidities of transient boron Lewis acids have been estimated and their high Lewis acidity has been confirmed.
Article
Complexes of Group 2, 12, 13, 14, 15 and 16 elements with mono-, bi-, and poly-dentate phosphine and arsine ligands (and including the very few examples of stibine and bismuthine donor ligands) are described. Polydentate ligand complexes containing neutral or charged N, O, C, or S donor groups in addition to phosphino or arsino donor groups are included, but charged P or As (phosphides, arsenides, phosphinomethanides etc.) ligands are excluded. Emphasis is placed upon the X-ray structures, multinuclear NMR data and reactions. The major differences of this class of complexes compared to the familiar d-block phosphine/arsine complexes are discussed and rationalised in terms of the E–M bonding models. Literature coverage is focussed on the last 20 years, although key older work is also included where necessary for comparison purposes, and the article includes work published up to early 2013.