Equilibrium CAs and CSb bond lengths Jean Demaison , Harald Møllendal

advertisement
Journal of Molecular Structure 930 (2009) 21–25
Contents lists available at ScienceDirect
Journal of Molecular Structure
journal homepage: www.elsevier.com/locate/molstruc
Equilibrium CAs and CSb bond lengths
Jean Demaison a,*, Harald Møllendal b, Jean-Claude Guillemin c
a
b
c
Laboratoire de Physique des Lasers, Atomes, et Molécules, UMR CNRS 8523, Université de Lille I, F-59655 Villeneuve d’Ascq Cédex, France
Centre for Theoretical and Computational Chemistry (CTCC), Department of Chemistry, University of Oslo, P.O. Box 1033, Blindern, N-0315 Oslo, Norway
Sciences Chimiques de Rennes, Ecole Nationale Supérieure de Chimie de Rennes-CNRS, 35708 Rennes, France
a r t i c l e
i n f o
Article history:
Received 31 March 2009
Accepted 17 April 2009
Available online 3 May 2009
Keywords:
Equilibrium structure
Carbon–arsenic bond length
Carbon–antimony bond length
a b s t r a c t
The equilibrium structures of some small molecules containing the CAs or the CSb bond are determined
using the method of predicate observations. The input data are the semi-experimental equilibrium rotational constants (experimental ground state rotational constants corrected for the rovibrational contribution calculated from the ab initio cubic force field) and the bond lengths and bond angles calculated ab
initio using the CCSD(T) method and a relativistic effective core potential of quadruple zeta quality. It
is shown that this method, avoiding difficult isotopic substitutions, allows us to obtain reliable equilibrium structures.
Ó 2009 Elsevier B.V. All rights reserved.
1. Introduction
The equilibrium CX bond lengths are accurately known for different atoms X in many molecules where X = H [1], C [2], N [3], O
[3], F [4], and Hal (where Hal = Cl, Br, and I) [5]. On the other hand,
data for the CAs and CSb bond lengths are extremely scarse. For the
CAs bond, the microwave spectra of difluoromethylarsine, CH3AsF2
[6], ethylidynearsine, CH3CAs [7,8] and trimethylarsine, (CH3)3As
[9] have been measured and the structure of arsabenzene, cC5H5As, has been investigated by microwave spectroscopy [10]
and by electron diffraction [11]. For all these molecules, an approximate value of the CAs bond length has been estimated, but, with
the exception of CH3CAs [8], this estimation was based on many
assumptions and it is thus rather inaccurate, see Table 1. For the
CSb bond length, the data are still more scarse, only the microwave
spectra of vinylstibine, H2C@CHSbH2 [12], and stibabenzene, cC5H5Sb [13] have been measured leading to a very approximate value of the CSb bond length (see Table 1).
In addition to that, the microwave spectra of vinylarsine,
H2C@CHAsH2 [14] ethynylarsine, HC„CAsH2, and ethylarsine,
CH3CH2AsH2 [15] have also been recently measured in our laboratories allowing us to obtain very accurate ground state rotational
constants. The microwave spectrum of methylarsine, CH3AsH2,
has also been measured but never published [16]. These experimental ground state rotational constants do not permit to calculate
an equilibrium structure but they can be corrected using ab initio
rotation–vibration interaction constants (a-constants) in order to
obtain semi-experimental equilibrium rotational constants [17].
* Corresponding author. Fax: +33 3 20 33 70 20.
E-mail address: jean.demaison@univ-lille1.fr (J. Demaison).
0022-2860/$ - see front matter Ó 2009 Elsevier B.V. All rights reserved.
doi:10.1016/j.molstruc.2009.04.029
With the exception of CH3CAs, the number of rotational constants
is not large enough to determine an equilibrium structure. However, each molecule may be divided in three parts: (i) the alkyne
(hydrocarbon) group, CmHn, whose structure may be accurately
computed using standard ab initio methods; (ii) the arsine (resp.
stibine) group, AsH2 (resp. SbH2) whose ab initio structure may
be scaled using the known equilibrium structure of AsH3 [18] (resp.
SbH3 [19]); (iii) the CX (X = As, Sb) bond length and the \(CCX)
bond angle whose values may be crudely estimated using ab initio
methods. The goal of this paper is to show that it is possible to
obtain a reliable equilibrium structure for these molecules by combining the ab initio equilibrium structure and the semi-experimental rotational constants in the method of predicate observations. In
this method, the rotational constants and the ab initio structure are
used as input in a least-squares fit, each datum being weighted
according to the value of its estimated uncertainty [20,21] .
The paper is organized as follows. Section 2 describes the ab initio
calculations. Section 3 is dedicated to the equilibrium structure of
CH3CAs. Section 4 is devoted to the structure of HC„CAsH2. This section is more detailed than others because it explains the methodology used which is then applied in the subsequent sections to the
other molecules, CH2@CHAsH2, CH3CH2AsH2, and CH2@CHSbH2.
2. Ab initio calculations
Geometry optimizations have been carried out at two levels of
electronic structure theory, second-order Møller–Plesset perturbation theory (MP2) [22] and coupled-cluster (CC) method with
single and double excitations (CCSD) [23] augmented by a
perturbational estimate of the effects of connected triple
excitations [CCSD(T)] [24]. For the H and C atoms, Dunning’s
22
J. Demaison et al. / Journal of Molecular Structure 930 (2009) 21–25
Table 1
Previous determinations of the CAs or the CSb bond length (values in Å).
Molecule
Bond
Structurea
value
Ref.
(CH3)3As
CH3AsF2
c-C5H5As
CH3CAs
c-C5H5Sb
CH2@CHSbH2
C–As
r0
r0
rg
r eb
r0
rs
1.959(10)
1.92(10)
1.850(3)
1.658(2)
2.050(5)
2.13(3)
[9]
[6]
[11]
[8]
[13]
[12]
C–Sb
a
r0 = effective structure from ground state rotational constants, rg = average
internuclear distance at thermal equilibrium from electron diffraction, re = equilibrium structure, rs = substitution structure.
b
Extrapolated from the average structure (rz).
correlation-consistent polarized valence n-uple zeta basis sets, ccpVnZ [25] with n {T, Q} were employed. For the As and Sb atoms,
the correlation-consistent valence basis sets with the StuttgartDresden-Bonn relativistic effective core potential, SDB-cc-pVnZ
[26] (abbreviated as SDB-VTZ) were used. In all calculations, the
frozen core approximation was used, unless otherwise stated.
The anharmonic force field was calculated at the MP2 level of
theory using the SDB-cc-pVTZ basis set. There is indeed a large
documented evidence which shows that, in most cases, the
CCSD(T) force field only provides a negligible improvement over
the MP2 force field in determining the semi-experimental equilibrium structure [27,28]. This will be further checked here in the particular case of CH3CAs. The reference geometry was first
determined via optimization prior to computing the force field at
the same level of theory. Then, the associated harmonic force field
was evaluated analytically in Cartesian coordinates. The cubic (/ijk)
and semidiagonal quartic (/ijkk) normal coordinates force constants
were determined with the use of a finite difference procedure
involving displacements [29] along reduced normal coordinates
and the calculation of analytic second derivatives at these displaced geometries. This procedure was repeated for all isotopologues. The semi-experimental equilibrium rotational constants
needed to determine a semi-experimental equilibrium structure
are then obtained by correcting the experimental ground state
rotational constants with the a-constants, computed from the ab
initio cubic force field via the expression
Bne ¼ Bn0 þ
X
i
ani
di
2
ð1Þ
where n = a, b, c, referring to the principal axes of the molecule; the
sum is running over all normal modes and di is a degeneracy factor
(1 for non-degenerate vibrations and 2 for doubly-degenerate vibrations). The evaluation of the a-constants as well as the other spectroscopic parameters was based on second-order perturbation
theory [30].
For the calculation of the magnetic g-tensor, the Kohn-Sham
density functional theory [31] using Becke’s three-parameter
hybrid exchange functional [32] and the Lee-Yang-Parr correlation
functional [33] together denoted as B3LYP, was also employed together with the split-valence basis set 6-311G(3df,2pd), as implemented in GAUSSIAN03 (g03) [34] for the H,C and As atoms. For
molecules containing Sb, the DGDZVP basis set, as implemented
in g03, was used. The diagonal elements of the g-tensor permit
to calculate the electronic correction to the rotational constants
using the relation
Bncorr ¼
Bnexp
1 þ ðm=Mp Þg nn
ð2Þ
where gnn is expressed in units of the nuclear magneton, m is the
electron mass, and Mp the proton mass. The calculated values of
the g-tensor are given in Table S1 of the Supplementary material.
This correction was found negligible for all molecules of this work.
All computations were performed with g03.
3. Ethylidynearsine
Ethylidynearsine, CH3CAs, is the most promising molecule from
the point of view of structure determination because it has already
been thoroughly studied. The ground state rotational constants of
eight isotopologues have been determined, its infrared spectrum
has been measured, an average structure (rz) has been calculated
and an approximate equilibrium structure has been deduced [8].
The anharmonic force field, calculated at the MP2/SDB-VTZ level of theory, was used to calculate the values of the spectroscopic
parameters for which experimental values are known: quartic centrifugal distortion constants and band centers. The results are given in Table S2 of the Supplementary material. The good
agreement between the calculated and experimental parameters
indicates that the force field is reliable.
The semi-experimental (se) equilibrium structure was calculated from a weighted least-squares fit of the related equilibrium
moments of inertia and is given in Table 2. The system of normal
equations is moderately well conditioned (condition number
j = 264) [35]. This is mainly due to the fact that there is no isotopic
substitution available for As. However, the standard deviation of
the derived parameters is extremely small, indicating that the
re(se) structure should be accurate. The semi-experimental structure is in very good agreement with the previous equilibrium
structure (extrapolated from the rz structure), but it is much more
accurate. The ab initio structures are also given in Table 2. The
CCSD(T)/SDB-VQZ values with all electrons correlated are in very
good agreement with the semi-experimental structure. For the
methyl part, it was expected [36] but for the CAs bond length, it
might be accidental. The CC bond length, at 1.463 Å, is longer than
in methyl cyanide (1.459 Å) [27] indicating a smaller degree of
double bond character. Finally, it as to be noted that the CH bond
length at 1.0910(3) Å is longer than in CH3CN where the value is
1.0865(1) Å [27]. This is in agreement with the fact that the CH
stretching frequencies are larger in CH3CN than in CH3CAs. For instance, the CH3 symmetric stretch, m1, is at 2840 cm1 in CH3CAs
[8] and 2954 cm1 in CH3CN [27].
4. Ethynylarsine, HCCAsH2
The experimental ground state rotational constants of
HC„CAsH2, B0 = 3666.1654(6) MHz and C0 = 3659.6748(6) MHz
[15] are corrected with the a-constants from the MP2/SDB-VTZ
force field to give the semi-experimental equilibrium rotational
constants, Be = 3659.3(2) MHz and Ce = 3652.8(2) MHz where the
quoted uncertainty takes into account the uncertainty on the ab
initio a-constants. The ab initio structure has been optimized at
Table 2
Structure of ethylidynearsine, CH3CAs (distances in Å, angles in degrees).
Method
Basis set
a
r0
reb
MP2 fcc
CCSD(T) aec
re(se)d
a
SDB-VTZ
SDB-VQZ
r(CH)
r(CC)
r(CAs)
\(HCC)
1.105(7)
1.091(3)
1.090
1.090
1.0910(3)
1.465(2)
1.465(2)
1.466
1.465
1.4634(2)
1.660(1)
1.658(2)
1.677
1.659
1.6585(2)
110.6(2)
110.6(2)
110.71
110.71
110.575(10)
r0 = effective structure from a fit of the ground state rotational constants, Ref.
[8].
b
c
d
Extrapolated from the average structure (rz), Ref. [8].
fc = frozen core approximation, ae = all electrons correlated.
Semi-experimental equilibrium structure.
23
J. Demaison et al. / Journal of Molecular Structure 930 (2009) 21–25
Table 3
Structure of ethynylarsine, HC„CAsH2 (distances in Å, angles in degrees).
MP2a
r(H–C)
r(CC)
r(C–As)
r(As–H)
\(CCAs)
\(CAsH)
\(HAsH)
\(HCC)
a
b
CCSD(T)a
SDB-VTZ
SDB-VQZ
SDB-VTZ
SDB-VQZ
1.0623
1.2201
1.8922
1.5121
173.145
94.727
92.252
178.608
1.0624
1.2174
1.8842
1.5091
172.928
94.682
92.013
178.554
1.0642
1.2167
1.9025
1.5207
173.388
94.663
92.132
178.611
1.0643
1.214
1.8945
1.5177
173.171
94.618
91.894
178.557
Predicateb
Uncertainty
re
1.0628
1.2105
1.8945
1.5120
173.171
94.618
91.894
178.557
0.0020
0.0020
0.020
0.002
0.3
0.5
0.4
0.3
1.063
1.211
1.894
1.512
173.25
94.85
92.24
178.55
Frozen core approximation.
See text.
the MP2 and CCSD(T) levels with triple and quadruple zeta basis
sets (see Table 3). The CCSD(T)/SDB-VQZ bond lengths of the ethynyl group are corrected for the core correlation which is estimated
to be 0.0015 Å for the CH bonds and 0.0035 Å for the C„C
bonds [5]. The uncertainty of the CH and CC bond lengths was assumed to be 0.002 Å. For the AsH bond length and the \(HAsH)
bond angle, the MP2/SDB-VTZ values were preferred because, in
the case of AsH3, the MP2/SDB-VTZ structure was found to be the
closest one of the equilibrium structure (see Table 4). The uncertainty of the MP2/SDB-VTZ AsH bond length was assumed to be
0.002 Å and that of the \(HAsH) bond angle, 0.4 degree. For the
remaining parameters, the CCSD(T)/SDB-VQZ values were used
with an uncertainty of 0.020 Å for the r(CAs) bond length, 0.3 degree for the \(HCC) and \(CCAs) bond angles and 0.5 degree for
the \(CAsH) bond angle. We also made a sensitivity analysis. The
most sensitive parameters for the B and C rotational constants
are r(C-As) and \(CCAs), and to a lesser extent \(CAsH), i.e. the
parameters whose ab initio values are difficult to calculate accurately. This is a very favourable case for the use of the method of
predicate observations. The ab initio equilibrium parameters (column predicate of Table 3) were fitted together with the two
semi-experimental equilibrium rotational constants giving a structure reported in the last column (re) of Table 3.
The results of the fit are worth to be commented. The CH and CC
and AsH bond lengths remain unchanged, as expected. The CAs
bond length is also unchanged which was not expected. Only
two bond angles are significantly changed, the \(HAsH) angle
being larger by 0.35 degree and the \(CAsH) angle being also larger
by 0.23 degree but these changes are rather small indicating that
the predicate (ab initio) structure is likely to be close to the true
equilibrium structure and confirming that the fitted structure is
probably reliable. Another interesting conclusion is that the HCCAs
moiety is not linear, the \(CCAs) angle being as large as 6.8 degree
and the ethynyl H being in trans position with \(HCC) = 1.45
degree.
5. Other molecules
The structure of the other molecules, syn-vinylarsine, transethylarsine, and syn-vinylstibine is calculated in the same way as
for ethynylarsine. The experimental ground state rotational constants and the derived semi-experimental equilibrium rotational
constants are given in Table 5. The calculated structures are given
in Tables 6–8. The only difference is that, for the predicate structure of syn-vinylstibine, an offset of 0.0040 Å for the SbH bond
was estimated at the MP2/DSDB-VTZ level from the equilibrium
structure of stibine (bottom of Table 8) and the predicate value
of the CSb bond length was assumed to be equal to the MP2/
DSDB-VTZ value.
It is important to try to estimate the accuracy of the derived
structures. However, the number of input data is not large enough
to try a statistical estimation. Nevertheless, a comparison with
many other molecules indicates that the accuracy of the bond an-
Table 5
Experimental ground state and semi-experimental equilibrium rotational constants
(in MHz).
Basis set
r(As–H)
\(HAsH)
This work
VTZ-PPe
SDB-VTZ
SDB-VQZ
SDB-VTZ
SDB-VQZ
SDB-VTZ
SDB-VQZ
1.5111
1.513(2)
1.5109
1.4896
1.5106
1.5078
1.5201
1.5168
1.5200
1.5188
92.069
92.08(7)
92.189
92.3874
92.0892
91.8824
92.110
91.729
91.9694
91.881
b
re
r ec
r ed
MP2
MP2
MP2
CCSD(T) ae
CCSD(T) ae
CCSD(T)
CCSD(T)
Ref.a
B
C
3666.16536(55)
3659.67482(55)
3659.34(20)
3652.77(20)
15
syn-H2C@CHAsH2
A
B
C
36322.959(29)
3941.36086(89)
3678.56301(99)
36634.7(31)
3964.01(23)
3701.56(23)
14
trans-CH3CH2AsH2
A
B
C
24410.16(34)
3537.6052(18)
3313.1621(17)
24582.4(31)
3575.95(23)
3349.43(23)
15
syn-H2C@CH121SbH2
A
B
C
32773.1831(3)
3116.0358(1)
2944.7042(1)
33017.7(30)
3133.18(20)
2961.57(20)
12
syn-H2C@CH123SbH2
A
B
C
32766.0381(5)
3107.9617(2)
2937.4356(2)
33010.3(30)
3125.11(20)
2954.31(20)
syn-H2C@13CH121SbH2
A
B
C
32107(1)
3068.7252(5)
2897.1315(5)
32344.2(10)
3085.35(20)
2913.56(20)
syn-H2C@13CH123SbH2
A
B
C
32117(1)
3060.525(1)
2889.763(1)
32354.2(10)
3076.91(20)
2906.15(20)
a
Frozen core approximation unless otherwise specified, ae = all electrons
correlated.
b
Ref. [18].
c
Ref. [42].
d
This work, using the a-constants of Ref. [43].
e
Correlation consistent-like basis set with a small-core relativistic pseudopotential, Ref. [44].
Equilibrium
HC„CAsH2
Table 4
Structure of arsine, AsH3 (distances in Å, angles in degrees).
Methoda
Ground state
a
Reference for experimental ground state rotational constants.
24
J. Demaison et al. / Journal of Molecular Structure 930 (2009) 21–25
Table 6
Structure of syn-vinylarsine, H2C@CHAsH2 (distances in Å, angles in degrees).
Parametera
r(CC)
r(CHc)
r(CHt)
r(CAs)
r(CHg)
r(AsH)
\(CCHc)
\(CCHt)
\(CCAs)
\(HgCAs)
\(CAsH)
\(HAsH)
a
b
c
d
MP2
CCSD(T)
Predicatec
Uncertainty
re
1.3340d
1.0828
1.0823
1.9427
1.0827
1.5142
121.407
121.584
120.205
120.264
95.572
91.354
0.002
0.002
0.002
0.020
0.002
0.003
0.3
0.3
0.5
0.3
0.5
0.3
1.334
1.083
1.082
1.939
1.083
1.516
121.41
121.59
120.12
120.27
95.57
91.53
b
SDB-VTZ
SDB-VQZ
SDB-VTZ
SDB-VQZ
1.3361
1.0825
1.0824
1.9427
1.0826
1.5142
121.315
121.501
119.893
120.256
95.562
91.354
1.3338
1.0818
1.0814
1.9339
1.0821
1.5112
121.278
121.489
119.903
120.264
95.613
91.303
1.3397
1.0850
1.0848
1.9524
1.0847
1.5231
121.444
121.595
120.195
120.256
95.520
91.284
1.3374
1.0843
1.0838
1.9436
1.0842
1.5201
121.407
121.584
120.205
120.264
95.572
91.233
c = cis to As; t = trans to As; g = gem.
CCSD(T)/SDB-VTZ + MP2/SDB-VQZ – MP2/SDB-VTZ, this approximation is expected to be accurate for the CC and CH bonds lengths as well as the \(HCC) angles.
See text.
Offset, 0.0034 Å, Ref. [2].
Table 7
Structure of trans-ethylarsine, CH3CH2AsH2 (distances in Å, angles in degrees).
Parametera
r(CC)
r(CHs)
r(CHa)
r(CAs)
r(CH)b
r(AsH)
\(CCHs)
\(CCHa)
\(HaCHa)
\(CCAs)
\(HCAs)
\(HCH)b
\(CAsH)
\(HAsH)
a
b
c
d
MP2
CCSD(T)
SDB-VTZ
SDB-VQZ
SDB-VTZ
SDB-VQZc
1.5213
1.0906
1.0889
1.9724
1.0906
1.5146
111.334
110.856
107.570
115.557
105.987
106.535
94.770
91.835
1.5188
1.0894
1.0879
1.9629
1.0897
1.5115
111.349
110.842
107.546
115.400
106.073
106.512
94.891
91.742
1.5271
1.0937
1.0921
1.9827
1.0930
1.5238
111.226
110.884
107.610
115.508
106.011
106.678
94.883
91.769
1.5246
1.0925
1.0911
1.9732
1.0921
1.5207
111.241
110.871
107.586
115.351
106.097
106.655
95.004
91.677
Predicate
Uncertainty
re
1.5213d
1.0910
1.0896
1.9732
1.0906
1.5146
111.241
110.871
107.586
115.351
106.097
106.655
95.004
91.677
0.002
0.002
0.002
0.020
0.002
0.002
0.3
0.3
0.3
0.5
0.3
0.3
0.3
0.3
1.521
1.091
1.090
1.970
1.091
1.516
111.25
110.87
107.68
115.38
106.11
106.74
95.05
91.85
s = H in symmetry plane; a = H out of symmetry plane.
CH2 group.
CCSD(T)/SDB-VTZ + MP2/SDB-VQZ – MP2/SDB-VTZ, this approximation is expected to be accurate for the CC and CH bonds lengths as well as the \(HCC) angles.
Offset, 0.0033 Å, Ref. [2].
Table 8
Structure of syn-vinylstibine, H2C@CHSbH2 (distances in Å, angles in degrees)
Parametera
MP2
CCSD(T)
SDB-VTZ
SDB-VQZ
SDB-VTZ
SDB-VQZb
r(C@C)
r(C-Sb)
r(CHc)
r(CHt)
r(CHg)
r(SbH)
\(CCSb)
\(CCHc)
\(CCHt)
\(CSbH)
\(HgCSb)
\(HSbH)
1.3375
2.1403
1.0829
1.0834
1.0832
1.7079
120.5416
121.7448
121.5725
94.6510
120.1275
91.1136
1.3354
2.1323
1.0822
1.0825
1.0829
1.7025
120.6989
121.7926
121.5314
94.7083
120.1004
91.2947
1.3409
2.1505
1.0855
1.0858
1.0854
1.7189
120.7617
121.7950
121.6427
94.5655
119.8221
91.0197
1.3388
2.1425
1.0848
1.0849
1.0851
1.7135
120.919
121.843
121.602
94.623
119.795
91.201
SbH3
r(SbH)
\(HSbH)
1.7040
92.012
a
b
c
1.7153
91.818
Predicate
Uncertainty
re
1.3355c
2.1403
1.0833
1.0834
1.0836
1.7039
120.919
121.843
121.602
94.623
119.795
91.201
0.002
0.020
0.002
0.002
0.002
0.003
0.5
0.3
0.3
0.5
0.3
0.3
1.335
2.131
1.083
1.084
1.084
1.703
120.65
121.89
121.62
94.72
119.82
91.14
1.7000
91.56
c = cis to Sb; t = trans to Sb; g = gem.
CCSD(T)/SDB-VTZ + MP2/SDB-VQZ – MP2/SDB-VTZ, this approximation is expected to be accurate for the CC and CH bonds lengths as well as the \(HCC) angles.
Offset, 0.0034 Å, Ref. [2].
J. Demaison et al. / Journal of Molecular Structure 930 (2009) 21–25
gles is not worse than a few tenths of a degree [37]. Likewise, the
accuracy of the CH and CC bond lengths is probably not worse than
a few thousandths of an Angstrøm [2]. The accuracy of the CAs and
CSb bond lengths is still more difficult to estimate. However, a sensitivity analysis indicates that it is also probably not worse than a
few thousandths of an Angstrøm, a very conservative estimate
being 0.005 Å.
6. Discussion
The ethynyl CH bond length in HC„CAsH2 is only faintly longer
than in acetylene: 1.0628 Å instead of 1.0618 Å [38]. On the other
hand, the C„C bond length, at 1.211 Å is much longer than in acetylene (1.203 Å) [38]. It is also larger than in cyanoacetylene
(HC„C–C„N, 1.206 Å) [39].
The structure of the vinyl group is almost identical in
H2C@CHAsH2 and H2C@CHSbH2 (Tables 6 and 8). The C@C bond
length is slightly longer than in ethylene (1.331 Å) [2] but slightly
smaller than in vinylamine (1.336 Å) [40]. The CH bond lengths are
up to 0.005 Å longer than in vinylamine but the CH2 group is much
less asymmetric (cis and trans CH bond lengths and \(HCC) bond
angles are almost identical).
The C–C bond length in ethylarsine is not significantly different
to the value found for ethane (1.522 Å) [41] but the CH bond
lengths are again 0.001–0.002 Å longer.
As expected, the C–As bond length increases from ethynylarsine
to ethylarsine going through vinylarsine. The As–H bond length in
ethynylarsine, at 1.512 Å, is almost identical to the value found in
arsine (1.511 Å, see Table 4) but it is longer in both vinylarsine and
ethylarsine where the value is 1.516 Å. The \(HAsH) remains close
to the value found for arsine.
Acknowledgments
J.D. thanks the Barbara Mez-Starck foundation for financial support. J.-C.G. thanks the PNP (INSU-CNRS) for financial support.
Appendix A. Supplementary data
Supplementary data associated with this article can be found, in
the online version, at doi:10.1016/j.molstruc.2009.04.029.
References
[1] J. Demaison, H.D. Rudolph, J. Mol. Spectrosc. 248 (2008) 66–76.
[2] L. Margulès, J. Demaison, J.E. Boggs, Struct. Chem. 11 (2000) 145–154.
[3] J. Demaison, A.G. Császár, I. Kleiner, H. Møllendal, J. Phys. Chem. A 111 (2007)
2574–2586.
[4] L. Margulès, J. Demaison, J.E. Boggs, J. Phys. Chem. A 103 (1999) 7632–7638.
[5] J. Demaison, L. Margulès, J.E. Boggs, Struct. Chem. 14 (2003) 159–174.
[6] L.J. Nugent, C.D. Cornwell, J. Chem. Phys. 37 (1962) 523–534.
[7] J.-C. Guillemin, L. Lassalle, P. Dréan, G. Wlodarczak, J. Demaison, J. Amer. Chem.
Soc. 116 (1994) 8930–8936.
25
[8] P. Dréan, H. Bürger, J. Demaison, J.E. Boggs, J. Mol. Struct. 485–486 (1999) 51–
56.
[9] D.R. Lide, Spectrochim. Acta 15 (1959) 473–476.
[10] R.P. Lattimer, R.L. Kuczkowski, A.J. Ashe, A.L. Meinzer, J. Mol. Spectrosc. 57
(1975) 428–435.
[11] T.C. Wong, A.J. Ashe, L.S. Bartell, J. Mol. Struct. 25 (1975) 65–74.
[12] D. Petitprez, S. Kassi, G. Wlodarczak, J. Demaison, I. Merke, S.E. Chaouch, J.-C.
Guillemin, Phys. Chem. Chem. Phys. 3 (2001) 2415–2418.
[13] C.D. Fong, R.L. Kuczkowski, A.J. Ashe, J. Mol. Spectrosc. 70 (1978) 197–207.
[14] P. Dréan, D. Petitprez, J. Demaison, J.C. López, J.L. Alonso, J.-C. Guillemin, J.E.
Boggs, J. Mol. Spectrosc. 190 (1998) 365–371.
[15] P. Dréan, PhD Thesis, Lille, 1995.
[16] D.M. Levine, PhD Thesis, Pennsylvania State University, 1973.
[17] F. Pawlowski, P. Jørgensen, J. Olsen, F. Hegelund, T. Helgaker, J. Gauss, K.L. Bak,
J.F. Stanton, J. Chem. Phys. 116 (2002) 6482–6496.
[18] M. Carlotti, G. Di Lonardo, L. Fusina, J. Mol. Spectrosc. 102 (1983) 310–319.
[19] E. Canè, G. Lonardo, L. Fusina, W. Jerzembeck, H. Bürger, J. Breidung, W. Thiel, J.
Mol. Struct. 780–781 (2006) 98–110.
[20] L.S. Bartell, D.J. Romanesko, T.C. Wong, in: G.A. Sims, L.E. Sutton (Eds.),
Chemical Society Specialist Periodical Report No. 20: Molecular Structure by
Diffraction Methods, vol. 3, The Chemical Society, London, 1975, p. 72.
[21] A. Sen, M. Srivastava, Regression Analysis, Springer-Verlag, New York, 1990.
[22] C. Møller, M.S. Plesset, Phys. Rev. 46 (1934) 618–622.
[23] G.D. Purvis III, R.J. Bartlett, J. Chem. Phys. 76 (1982) 1910–1918.
[24] K. Raghavachari, G.W. Trucks, J.A. Pople, M. Head-Gordon, Chem. Phys. Lett.
157 (1989) 479–483.
[25] T.H. Dunning Jr., J. Chem. Phys. 90 (1989) 1007–1023.
[26] J.M.L. Martin, A. Sunderman, J. Chem. Phys. 114 (2001) 3408–3420.
[27] C. Puzzarini, G. Cazzoli, J. Mol. Spectrosc. 240 (2006) 260–264.
[28] P. Groner, R.D. Warren, J. Mol. Struct. 599 (2001) 323–335.
[29] W. Schneider, W. Thiel, Chem. Phys. Lett. 157 (1989) 367–373.
[30] D. Papousek, M.R. Aliev, Molecular Vibrational–Rotational Spectra, Elsevier,
Amsterdam, 1982.
[31] W. Kohn, L.J. Sham, Phys. Rev. A 140 (1965) 1133–1138.
[32] A.D. Becke, J. Chem. Phys. 98 (1993) 5648–5652.
[33] C.T. Lee, W.T. Yang, R.G. Parr, Phys. Rev. B 37 (1988) 785–789.
[34] M.J. Frisch, G.W. Trucks, H.B. Schlegel, G.E. Scuseria, M.A. Robb, J.R. Cheeseman,
J.A. Montgomery Jr., T. Vreven, K.N. Kudin, J.C. Burant, J.M. Millam, S.S. Iyengar,
J. Tomasi, V. Barone, B. Mennucci, M. Cossi, G. Scalmani, N. Rega, G.A.
Petersson, H. Nakatsuji, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa,
M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, M. Klene, X. Li, J. E. Knox, H.
P. Hratchian, J.B. Cross, C. Adamo, J. Jaramillo, R. Gomperts, R.E. Stratmann, O.
Yazyev, A.J. Austin, R. Cammi, C. Pomelli, J.W. Ochterski, P.Y. Ayala, K.
Morokuma, G.A. Voth, P. Salvador, J.J. Dannenberg, V.G. Zakrzewski, S.
Dapprich, A.D. Daniels, M.C. Strain, O. Farkas, D.K. Malick, A.D. Rabuck, K.
Raghavachari, J.B. Foresman, J.V. Ortiz, Q. Cui, A.G. Baboul, S. Clifford, J.
Cioslowski, B.B. Stefanov, G. Liu, A. Liashenko, P. Piskorz, I. Komaromi, R.L.
Martin, D.J. Fox, T. Keith, M.A. Al-Laham, C.Y. Peng, A. Nanayakkara, M.
Challacombe, P.M.W. Gill, B. Johnson, W. Chen, M.W. Wong, C. Gonzalez, J.A.
Pople, Gaussian, Inc., Wallingford CT, Revision D.01, 2004.
[35] D.A. Belsley, Conditioning Diagnostics, Wiley, New York, 1991.
[36] T. Helgaker, J. Gauss, P. Jørgensen, J. Olsen, J. Chem. Phys. 106 (1997) 6430–
6440.
[37] L. Margulès, J. Demaison, J.E. Boggs, J. Mol. Struct. (Theochem) 500 (2000) 245–
258.
[38] G. Cazzoli, C. Puzzarini, L. Fusina, F. Tamassia, J. Mol. Spectrosc. 247 (2008)
115–118.
[39] P. Botschwina, Phys. Chem. Chem. Phys. 5 (2003) 3337–3348.
[40] E. Askeland, H. Møllendal, E. Uggerud, J.-C. Guillemin, J.-R. Moreno, J.
Demaison, T.R. Huet, J. Phys. Chem. A 110 (2006) 12572–12584.
[41] C. Puzzarini, P.R. Taylor, J. Chem. Phys. 122 (2005) 054315/1–10.
[42] G.A. McRae, M.C.L. Gerry, M. Wong, I. Ozier, E.A. Cohen, J. Mol. Spectrosc. 123
(1987) 321–339.
[43] J.Y. Mandin, V. Dana, G. Tarrago, S. Klee, B.P. Winnewisser, J. Mol. Spectrosc.
172 (1995) 319–329.
[44] K.A. Peterson, J. Chem. Phys. 119 (2003) 11099–11112.
Download