Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Article

pubs.acs.org/crystal

Adducts of S/Se Donors with Dihalogens as a Source of Information


for Categorizing the Halogen Bonding
Published as part of the Crystal Growth & Design virtual special issue on Halogen Bonding in Crystal
Engineering: Fundamentals and Applications
M. Carla Aragoni, Massimiliano Arca, Francesco A. Devillanova,* Francesco Isaia, and Vito Lippolis
Dipartimento di Scienze Chimiche e Geologiche, Università degli Studi di Cagliari, S.S. 554 bivio per Sestu, 09042 Monserrato -
Cagliari, Italy

ABSTRACT: The great variety of products ensuing from the reactions


Downloaded by UNIV EASTERN FINLAND on September 8, 2015 | http://pubs.acs.org

between chalcogen donors and dihalogens and their assemblies in the crystal
lattice provides a wide experimental basis useful to disembroil the ongoing
debate on how to define the halogen bonding. In this paper we present a
critical analysis/study of the structural features retrieved from the Cambridge
Publication Date (Web): May 11, 2012 | doi: 10.1021/cg201328y

Structural Database (CSD) for linear three-body systems involving either


halogens, X−X−X (X = Br, I), or halogen(s) and chalcogen(s) atoms, E−X−
Y, X−E−Y, and E−X−E (E = S, Se; X = Y = Cl, Br, I; X = I, Y = Cl, Br, I). The
relative elongations (δ) of the two bonds in the examined three-body systems
with respect to the sum of the relevant atomic radii can be fitted by a common
nonlinear least-squares equation derived from the bond-valence model. The
similarities observed in the structural features suggest a common nature of the
chemical bond in all systems considered and indicate that the charge transfer
and the 3c-4e models can be successfully applied to all the cases considered to
explain the nature of the chemical bonding.

■ INTRODUCTION
It is well-known that the reaction between molecules containing
base nature of the reactants and the experimental conditions
used.1−12 In an attempt to evaluate whether the different
chalcogen donor atoms, E (E = S, Se), and dihalogens, XY (X = products obtainable from these reactions could ensue from a
Y = I, Br; X = I, Y = Br, Cl), can afford a great variety of common intermediate species, Husebye et al.10 proposed the
products including neutral charge-transfer (CT) adducts cation [>C−E−X]+ as a transitional species. This cation could
featuring an almost linear E−X−Y moiety, and insertion undergo a nucleophile attack on either the chalcogen or the
adducts containing a “T-shaped” X−E−Y fragment.1−12 halogen site, and, depending on the nature of the nucleophile,
According to the notation introduced by Martin and cow- the formation of the above-mentioned adducts, the formation
orkers,13,14 these adducts are respectively referred to as 10-X-2 of dications [>C−E−E−C<]+2 bearing a chalcogen-chalcogen
and 10-E-3 hypervalent compounds, indicating that the central single bond, and the formation of [>C−E−X−E−C<]+ cations
atom is formally associated with five electron pairs (10 featuring a central X+ coordinated by two donor molecules, can
electrons), among which 2 or 3, respectively, are bond be explained easily.2a,10,11 Notwithstanding the fact that the
pairs.15,16 Both types of adducts feature the central atom in a existence of this intermediate in solution has only been proven
hypervalent state, the two different three-body systems (E−X− in a few cases,17 the charge distribution calculated on the cation
Y and X−E−Y) being linear and about 10−12% longer than the [>C−E−X]+ assumed as a theoretical transitional species can
sum of the corresponding covalent radii.15 The chemical bond help in predicting the nature of the final product.2a,11b
in linear E−X−Y and X−E−Y fragments, as well as in trihalides
X3−,15 can be described by using the 3c-4e bonding scheme,14
according to which these moieties are characterized by a total
■ RESULTS AND DISCUSSION
It has been shown that the reaction between chalcogen donors
bond order of 1, with each one of the two bonds exhibiting and dihalogens can yield a plethora of products. Among these a
bond orders varying from 0 to 1 (see Figure 1 for selected recurrent structural motif is the presence of linear three-body
examples).15 However, besides the adducts containing the E− systems, recognizable either as fundamental molecular con-
X−Y and X−E−Y fragments, other different structural
archetypes were established by X-ray structural analysis and Received: October 7, 2011
vibrational spectroscopy for the products of the reactions of Revised: April 5, 2012
chalcogen donors with dihalogens, depending both on the acid/ Published: April 23, 2012

© 2012 American Chemical Society 2769 dx.doi.org/10.1021/cg201328y | Cryst. Growth Des. 2012, 12, 2769−2779
Crystal Growth & Design Article

Table 1. Number of Fragments (Structures) in Selected Linear (Angle > 165°) Three-Body Systems Found in the CSD21 with
Interatomic Distances Ranging from the Sum of the Covalent Radii up to That of Van der Waals Radii
I Br Cl
S Se S Se S Se
X−X−X 1618 (1084) 181 (127) 10 (10)
E−X−X 147 (104) 43 (32) 11 (11) 10 (7)
X−E−X 1 (1) 7 (6) 8 (7) 194 (86) 8 (7) 155 (68)
E−E−X 36 (17) 13 (7) 85 (29) 42 (18) 56 (24) 23 (13)
E−X−E 12 (11) 7 (6) 4 (3) 2 (2)
E−E−E 109 (81) 85 (58)

0
stituents or as supramolecular assemblies in competition with sAB = e(dAB− dAB) / kAB (2)
other intermolecular interactions, in the crystal lattice.1−12,18−20
In order to identify the existing sequences of chalcogen and where d0AB, calculated as a sum of listed empirical parame-
halogen atoms involved in linear three-body systems, a search ters,28,29 represents the A−B distance in isolated systems with
on the Cambridge Structural Database (CSD)21 was performed sAB = 1, and dAB is the distance in a perturbed system (sAB < 1).
on the fragments A−B−C (A, B, C = S, Se, Cl, Br, I) by kAB is a parameter which only marginally depends on the nature
imposing linearity (A−B−C angle > 165°) and bond distance of the atomic species A and B and shows values of about 0.37 Å
Downloaded by UNIV EASTERN FINLAND on September 8, 2015 | http://pubs.acs.org

values ranging between single bonds and the sum of the van der for most bonds.29a
Waals radii of the involved atoms (Table 1). Occurrences were By assuming that in A−B···A systems, the relationship sA−B +
found in the CSD for almost all the considered combinations, sB···A = 1 holds, it can be shown that23
Publication Date (Web): May 11, 2012 | doi: 10.1021/cg201328y

and Figure 1 reports the scatter plots of the two bond distances 0
(dAB vs. dBC) for some selected systems showing a considerable
0
dAB = dAB − kAB ln[1 − e−(dAB− dAB) / kAB] (3)
incidence. It is important to point out that the imposed CSD By extending this equation to A−B···C three-body systems,
search conditions do not take in consideration the charge borne the following equation can be written:
by the fragment, so that, for example, the structural data relative
0
to the tri-iodine systems (Figure 1a) not only include discrete 0
dAB = dAB − kAB ln[1 − e−(dBC− dBC) / kBC] (4)
I3− anions but also all the sequences of three iodine atoms
following the imposed conditions. If d0AB and d0BC are approximated to the sum of covalent
It is important to stress that among the considered three- radii, rA + rB and rB + rC, respectively,30 and eq 1 is introduced
body systems, those featuring a central halogen atom are in eq 4, eq 5 can be obtained:
currently classified as cases in which a halogen bonding kAB 0
occurs.22 Thus, information obtained from the analysis of the δAB = − 0
ln[1 − e−δBCdBC/ kBC]
data in the CSD can help obtain a deeper insight into the dAB (5)
nature of the halogen bond in its widest definition within an On the condition that
A−X−B system (A, B elements, X halogen). The scatter plots
shown in Figure 1 indicate that the two bond distances in the kAB kBC
= =k
considered three-body systems are correlated and follow similar 0
dAB 0
dBC (6)
trends, although the bond values directly depend on the nature
of the atoms involved. In order to quantitatively compare all which implies that the relationship δAB = f(δBC) should be
retrieved data, it is possible to consider for each couple of A−B symmetric with respect to the straight line of equation δAB =
and B−C distances (dAB and dBC, respectively) in an A−B−C δBC (the symmetry of δAB = f(δBC) with respect to the bisector
three-body system their normalized elongations with respect to of the I quadrant is required because in a A−B···C three-body
the sum of the covalent radii of the atomic species involved (rA, system A and C can be swapped). Eq 5 can be simplified in eq
rB, and rC): 7:

d − (rA + rB) d − (rB + rC) δAB = −k ln[1 − e−δBC/ k] (7)


δAB = AB ; δ BC = BC
rA + rB rB + rC (1) Equation 7 provides a direct simplified relationship between δAB
and δBC in A−B−C systems following a bond-valence model,
In analogy with previous analysis of systems featuring a dependent on the single adimensional parameter k and
hydrogen bonding, several model functions f(δ) could be independent of the nature of the atomic species A, B, and C.
adopted for correlating δAB and δBC. In particular, the O−H and Therefore, eq 7 would allow for the direct comparison of δAB
H···O distances in O−H···O systems have been fitted23 by and δBC values in three-body systems featuring different
adopting an expression directly derived from the bond valence composing atoms.
model,24−29 which has been widely applied as a tool for the In the following, A−B−C systems will be analyzed on the
interpretation and prediction of bond lengths in inorganic basis of the nature of the composing atoms.
crystals.25 Notably, this model, which was rationalized in terms E−X−Y Fragments. These fragments are mainly found in
of the molecular orbital theory,26 has been successfully applied CT adducts of sulfur and selenium compounds, and, among the
not only to ionic interactions, but also to covalent compounds, consistent number of structures reported, many examples
including those characterized by metallic bonding.27 According pertain to CT adducts formed by sulfur donors (E = S) and
to this model, the bond-valence24c sAB and the length dAB of an diiodine (X = Y = I). In the past we proposed a classification of
A−B bond are related by the approximated equation: these CT complexes into three categories: (i) weak or medium-
2770 dx.doi.org/10.1021/cg201328y | Cryst. Growth Des. 2012, 12, 2769−2779
Crystal Growth & Design Article
Downloaded by UNIV EASTERN FINLAND on September 8, 2015 | http://pubs.acs.org
Publication Date (Web): May 11, 2012 | doi: 10.1021/cg201328y

Figure 1. Scatter plots of the experimental bond distances (Å) for selected A−B−C (A, B, C = S, Se, Cl, Br, I) three-body systems obtained from a
CSD21 search by imposing the linearity of the fragments (angle > 165°) and interatomic distances lower than the sum of van der Waals radii of the
involved species: (a) I−I−I 1618 (1084); (b) Br−Br−Br 181 (127); (c) S−I−I 147 (104); (d) Se−I−I 43 (32); (e) Cl−Se−Cl 155 (68); (f) Br−
Se−Br 194 fragments (86 structures).

weak adducts, in which the S···I2 interaction can be seen as a between the iodine atoms ranges from values slightly lower
perturbation induced by the donor on the diiodine molecule. In than 1 to about 0.6 (dI−I < 2.86 Å). (ii) Very strong adducts in
these systems the I−I bond order nI−I, defined according to which the donor−acceptor interaction is so strong that nI−I
original Pauling’s equation31 as a function of the distance dI−I becomes lower than 0.4 (dI−I > 3.01 Å). (iii) Strong adducts
2771 dx.doi.org/10.1021/cg201328y | Cryst. Growth Des. 2012, 12, 2769−2779
Crystal Growth & Design Article
Downloaded by UNIV EASTERN FINLAND on September 8, 2015 | http://pubs.acs.org
Publication Date (Web): May 11, 2012 | doi: 10.1021/cg201328y

Figure 2. Structural data of the S−I−I fragments found in the CSD Figure 4. Structural data of the Se−X−Y fragments overlapped with
reported as scatter plot of δ1 vs δ2 where δ1 = δS−I and δ2 = δI−I (see eq those of the structural data reported in the scatter plot of Figure 3
1). The solid curve represents the least-squares fit of the data adopting depicted as dots [S−X−Y (·); Se−I−I (red open circles); Se−I−Br
eq 7 as a model. The red point has not been included in the fitting. (red open squares); Se−I−Cl (red plus signs); Se−Br−Br (red open
Fitted parameter k = 0.161; rmsd = 0.036; normalized rmsd = 0.133. triangles)] reported as scatter plot of δ1 vs δ2 where δ1 = δSe−X and δ2 =
δX−Y (see eq 1). The solid curve represents the least-squares fit of all
data adopting eq 7 as a model. Data with δX−Y > 0.35 along with the
red point in Figure 2 have been omitted from the fitting. Fitted
parameter k = 0.158; rmsd = 0.034; normalized rmsd = 0.089.

spectroscopy.15,32 In fact, the first class of compounds shows


FT-Raman spectra characterized by only one peak due to the
νI−I stretching vibration falling in the low frequency region at
frequencies lower than 213 cm−1 (gas phase) depending
directly on the entity of the donor−acceptor interaction
between the S donor molecule and diiodine. In the spectra of
the very strong adducts, no peaks assignable to the iodine−
iodine stretching vibration are found, thus supporting a
description of these very polarized systems as [>C−S−I]+···I−
CT adducts. For strong adducts, FT-Raman spectra are
characterized by three peaks in the low frequency region as
expected for a S−I−I three-body system.
The present class of compounds, where a continuous
variation of dS−I vs dI−I bond lengths (Figure 1c) is observed
in the condensed phases, shows close structural analogies with
systems featuring hydrogen bonding situations. In his review of
the hydrogen bonding in the solid state, Steiner observed that,
having defined X−H as the donor group and A as the hydrogen
Figure 3. Structural data of the S−X−Y fragments [S−I−Br (red open bonding acceptor, “the hydrogen bond phenomenon is a very
circles); S−I−Cl (red open squares); S−Br−Br (red open triangles)] broad one: there are dozens of different X−H···A interactions
reported as scatter plot of δ1 vs δ2 where δ1 = δS−X and δ2 = δX−Y (see that occur commonly in the condensed phases” and that
eq 1). The least-squares fit (eq 7) of the data provides k = 0.156 (rmsd “dissociation energies span more than two orders of magnitude,
= 0.028; normalized rmsd = 0.103). The solid curve represents the about 0.2−40 kcal·mol−1”, concluding that “the nature of the
least-squares fit of data including the systems featuring the S−I−I
interaction is not constant but includes electrostatic, covalent,
fragments (Figure 2): k = 0.160; rmsd = 0.035; normalized rmsd =
0.100. The red point in the scattergraph depicted in Figure 2 has not and dispersion contributions in varying weights”.33 Therefore,
been considered in the overall fit. while an electrostatic model could satisfactory explain the
chemical bonding in unbalanced S−I−I fragments in which one
of the two bond distances approach the sum of the relevant van
embracing all the intermediate cases characterized by quite der Waals radii, the other bond being substantially a single
balanced bond orders (0.4 < nI−I < 0.6; 2.86 Å < dI−I < 3.01 covalent bond, it becomes more and more inadequate to
Å).1,3a,31 This classification is supported by FT-Raman describe the chemical bonding on approaching more balanced
2772 dx.doi.org/10.1021/cg201328y | Cryst. Growth Des. 2012, 12, 2769−2779
Crystal Growth & Design Article
Downloaded by UNIV EASTERN FINLAND on September 8, 2015 | http://pubs.acs.org
Publication Date (Web): May 11, 2012 | doi: 10.1021/cg201328y

Figure 6. Structural data of the E−X−E fragments overlapped with


Figure 5. Structural data of the X−E−Y fragments overlapped with
those of the structural data reported in the scatter plot of Figure 5
those of the structural data reported in the scatter plot of Figure 4
depicted as dots [E−X−Y (·); X−E−Y (·); E−X−E (○)] reported as
depicted as dots [E−X−Y (·); Br−S−Br (red open triangles); I−S−I
scatter plot of δ1 vs δ2 where δ1 = δE−X and δ2 = δX−E (see eq 1). Data
(red open circles); Cl−S−Cl (red open squares); I−Se−I (green open
for GIGBED, TUNCAG, and VIYRIE were neglected since they
circles); I−Se−Br (red solid circles); Br−Se−Br (green open
feature atomic positions with mixed occupancies. Because of the
triangles); Cl−Se−Br (green solid circles); Cl−Se−Cl (green open
peculiar structure of the bromoselenate(I) anion [Se16Br18]2− in
squares)] reported as scatter plot of δ1 vs δ2 where δ1 = δX−E and δ2 =
MUHGUR the two points circled in blue were also neglected. The
δE−Y (see eq 1). The solid curve represents the least-squares fit of all
solid curve represents the least-squares fit of all data (excluding blue-
data adopting eq 7 as a model. The data with a red circle around them
circled data and those not considered in the fitting of Figures 2, 4, and
and those not considered in the fitting of Figures 2 and 4 have not
5) adopting eq 7 as a model. Fitted parameter k = 0.157; rmsd =
been taken into account in the overall fit. Fitted parameter k = 0.157;
0.049; normalized rmsd = 0.066.
rmsd = 0.050; normalized rmsd = 0.067.

we obtain the scatter plot shown in Figure 4, in which the solid


systems. This is the reason why, in our opinion, both the well- line represents the least-squares fit of the data according to eq
known CT molecular orbital (CTMO) scheme and the 7. All δE‑X/δX‑Y data with few exceptions, share the same
simplified molecular orbital model for electron-rich 3-center correlation (k = 0.158; rmsd = 0.034), with those related to
4-electron (3c-4e) systems by Rundle-Pimentel34,35 can selenium containing systems gathering on the right-hand side of
appropriately describe these systems.1,15b Obviously, in the the scatter plot in agreement with the stronger donor ability of
case of the 3c-4e bond model, the three pz orbitals of the three selenium donors as compared to the sulfur congeners.
aligned atoms will contribute differently to the three molecular X−E−Y Fragments. These fragments are mainly found in
orbitals having different positions in the energy scale. T-shaped 10-E-3 adducts between chalcogenone donors and
In Figure 2 the structural data for S−I−I systems are dihalogens featuring the central chalcogen atom in a hyper-
reported as a scatter plot of δS−I vs δI−I (see eq 1). A least- valent state. This class of compounds represents the only other
squares fit of the data according to eq 7 [root-mean-square possibility to linearly arrange one chalcogen and two halogen
deviation (rmsd) = 0.036] leads to a k value of 0.161. Because atoms. In Figure 5 all data found in the CSD for all the
parameter δ is used instead of the absolute values of the bond combinations X−E−Y (E = S, Se; X, Y = I, Br, Cl) are reported
distances, the same scatter plot (Figure 3) can also include data as a scattergraph of the parameters δX−E and δE−Y. The data
from the CSD relative to sulfur CT adducts with I−Br, I−Cl, relative to the systems E−X−Y previously discussed have also
and Br−Br (red symbols). Interestingly, all structural data for been introduced in the diagram as dots. All data from E−X−Y
sulfur CT adducts fall in the same correlation following the and X−E−Y systems, with few exceptions, share the same
application of eq 7 (rmsd on all data = 0.035; k = 0.160). As correlation curve following application of eq 7 (k = 0.157; rmsd
expected on the basis of the higher acidity of Br2, IBr, and ICl = 0.050).
with respect to I2, the data relative to the corresponding S- E−X−E Fragments. Among the products that can be
adducts are located on the right-hand side of the scatter plot obtained from the reactions of chalcogen donors and
with respect to the straight line of equation δSX = δXY (see dihalogens, E−X−E fragments are mainly found in [>C−E−
Figure 3). X−E−C<]+ so-called halonium complexes. These compounds
If the use of δ structural parameters is extended to the formally feature a central X+ (X = halogen) linearly coordinated
analogous systems featuring a Se−X−Y moiety (X = Y = I, Br; by two donor molecules. So far, most of these compounds
X = I, Y = Br, Cl), and all data related to the E−X−Y (E = S, reported in the literature feature a central I+ interacting with
Se) fragments retrieved from the CSD are processed together, either S- or Se-donors.3,15b,36−38 As for the linear three-body
2773 dx.doi.org/10.1021/cg201328y | Cryst. Growth Des. 2012, 12, 2769−2779
Crystal Growth & Design Article

Table 2. Number of Fragments (Structures) Belonging to Linear X−Y−Z (X, Y, Z = Cl, Br, I) Three-Body Systems Found in the
CSD21 for the Considered Sequences with Interatomic Distances Ranging from the Sum of the Covalent Radii up to That of
Van der Waals Radii
Y−Z = I−I I−Br I−Cl Br−Br Br−Cl Cl−Cl
X=I 1618 (1084) 13 (9) 9 (9) 2 (2) 5 (4)
Br 13 (9) 68 (51) 4 (4) 181 (127) 4 (4)
Cl 8 (8) 4 (4) 74 (59) 4 (4) 4 (2) 10 (10)
Downloaded by UNIV EASTERN FINLAND on September 8, 2015 | http://pubs.acs.org
Publication Date (Web): May 11, 2012 | doi: 10.1021/cg201328y

Figure 7. Structural data of the Br−Br−Br fragments overlapped with Figure 8. Structural data of the I−I−I fragments overlapped with those
those of the structural data reported in the scatter plot of Figure 6 of the structural data reported in the scatter plot of Figure 7 depicted
depicted as dots [E−X−Y (·); X−E−Y (·); E−X−E (·); Br−Br−Br as dots [E−X−Y (·); X−E−Y (·); E−X−E (·); Br−Br−Br (·); I−I−I
(○)]. The data are reported as scatter plot of δ1 vs δ2 where δ1 = (○)]. The data are reported as a scatterplot of δ1 vs δ2 where δ1 =
δBr1−Br2 and δ2 = δBr2−Br3 (see eq 1). Data for NOFYUC have not been δI1−I2 and δ2 = δI2−I3 (see eq 1). Data for AVIPIE, IWOFAB,
considered since they feature Br−Br bond distances shorter than the TEQCOH, and ZOLQUM have been neglected since they feature I−I
sum of two bromine covalent radii; the point circled in green bond distances shorter than the sum of two iodine covalent radii. The
corresponds to the tribromide in TIJLII whose structure was solved solid curve represents the least-squares fit of all data (excluding those
with an R factor of 11.08%. The solid curve represents the least- not considered in the fittings in Figures 2 and 4−7, and those featuring
squares fit of all data (excluding the circled point as well as those not δI−I > 0.35; see text) adopting eq 7 as a model. Fitted parameter for I−
considered in the fitting of Figures 2 and 4−6) adopting eq 7 as a I−I fragments: k = 0.150, rmsd = 0.010, normalized rmsd = 0.026; for
model. Fitted parameter for Br−Br−Br fragment: k = 0.164, rmsd = all data: k = 0.152; rmsd = 0.042; normalized rmsd = 0.057.
0.029, normalized rmsd = 0.048; for all data: k = 0.158; rmsd = 0.046;
normalized rmsd = 0.061. for E−X−Y and X−E−Y fragments (see above), where the
terminal halogen atoms of the three-body systems can be easily
systems discussed above, a correlation holds between the two involved in intermolecular contacts.
E−X bond distances: on reinforcing one bond, a lengthening of X−Y−Z Trihalogen Fragments. A survey of the crystal
the other is observed, so that the total length of the E−I−E structures deposited into the CSD containing any linear
framework is almost independent of the nature of the organic sequence of three halogen atoms, having each of the two
fragments bearing the chalcogens. Accordingly, the structural interatomic distances spanning from the sum of the covalent
data relative to these fragments reported as δE−X vs δX−E (eq 1) radii to the sum of the van der Waals radii provided the results
can be fitted along with those related to the three-body systems collected in Table 2. While no examples of I−Br−Cl and Br−
described above (eq 7; k = 0.157; rmsd = 0.049; Figure 6). Cl−Cl sequences were found, all the other combinations occur,
With the exception of two points, all the other points related to the I−I−I systems being the most numerous followed by Br−
E−X−E systems gather in the restricted area typical of systems Br−Br, Cl−I−Cl, and Br−I−Br ones (Table 2). For our
having fairly well balanced bonds. Two main reasons account purposes, the discussion can be limited to triiodine and
for this: (i) the potential energy surfaces (PESs) calculated for tribromine sequences, which are the most numerous. In these
the [HS−I−SH]− and [HSe−I−SeH]− model compounds are cases also, the structural data, expressed as δBr−Br and δI−I (eq
steeper than those calculated for CT and “T-shaped” adducts 1) share the same correlation found for the data corresponding
(see below); (ii) the organic framework incorporating the to the three-body systems discussed above and reported as dots
chalcogen donor can shield the E−X−E fragment from further in Figures 7 and 8, respectively. Remarkably, a certain number
intermolecular interactions, differently from what can happen of data, mainly belonging to triiodine fragments, spread out in
2774 dx.doi.org/10.1021/cg201328y | Cryst. Growth Des. 2012, 12, 2769−2779
Crystal Growth & Design Article
Downloaded by UNIV EASTERN FINLAND on September 8, 2015 | http://pubs.acs.org
Publication Date (Web): May 11, 2012 | doi: 10.1021/cg201328y

Figure 9. Potential energy surfaces (PESs) calculated for X2, X3− and for the model adducts [HE−X−X]− [E = S, Se; X = I (a), Br (b)]. d(X−X) −
d(X2)eq represents the difference between the X−X bond lengths d(X−X) with respect to the optimized ones d(X2)eq in the free dihalogens.

the regions of the scatter plot corresponding to very share the correlation defined by the points corresponding to all
unbalanced bonds. It is evident that for these three-body the other three-body systems considered so far (k = 0.152;
systems the correlation between the two bonds, expected rmsd = 0.042 for the cumulative fitting curve in Figure 8).
according to the 3c-4e model, is lost, one of the bonds being These experimental observations represent the main reason
longer than expected and the total bond order being lower than why the 3c-4e or CT models in our opinion are more adequate
1. In our opinion, this can be reasonably explained by to describe the “halogen bond” in linear systems featuring a
considering that the influence of the surrounding chemical central halogen atom than a purely electrostatic model based on
environment on the chemical bonding within the three-body the interaction between the positive electrostatic potential on
systems becomes stronger and stronger on moving toward the outer side of the halogen (σ-hole) and the negative site of a
bond distances that approach the sum of van der Waals radii. Lewis base (LB), which remains important in any case in
This notwithstanding, the great majority of triiodine systems driving the approach of LB to the halogen site.39 The use of the
(δI−I ≤ 0.35, corresponding to dI−I ≤ 3.6 Å; 87% of the data) 3c-4e model to describe linear three-body systems belonging to
2775 dx.doi.org/10.1021/cg201328y | Cryst. Growth Des. 2012, 12, 2769−2779
Crystal Growth & Design Article

H (−0.00220 au), at the Se−I bond critical point (BCP), and


the calculated Wiberg index (0.23) for this very polarized bond
indicate a partial covalent nature for it. Analogous results were
obtained for adducts featuring Se−X−Y moieties (X = Y = I; X
= I, Y = Br, Cl) with QTAIM parameters at the BCPs of the
Se−X and X−Y bonds, indicating a 3c-4e nature also for these
systems.
In order to better understand the reasons why the considered
three-body systems feature such bonding variability, potential
energy surfaces (PESs) were calculated at the DFT level for
some representative model systems (see the Experimental
Section). For the sake of comparison, the corresponding
surface calculations were extended to dihalogen molecules (X2).
In Figure 9, the PES curves calculated for X2, X3−, and the
model CT adducts [HS−X−X]− and [HSe−X−X]− are
depicted (X = I, a; X = Br, b). The PESs calculated for the
Figure 10. Potential energy surfaces (PESs) calculated for model T- three-body systems present a wide flat region around the
shaped adducts [X−HE−X]− systems (E = S, Se; X = Cl, Br, and I). energy minima as compared to the PES curves calculated for
Downloaded by UNIV EASTERN FINLAND on September 8, 2015 | http://pubs.acs.org

d(X−E) − d(X−E)eq represents the differences between the HE−X the 2c-2e systems X2. Because of the flat potential holes,
bond lengths d(X−E) with respect to the optimized ones d(X−E)eq.
distortions imposed by the organic framework or weak packing
interactions of few kcal mol−1 are sufficient to impose the small
Publication Date (Web): May 11, 2012 | doi: 10.1021/cg201328y

energy variations required to move the three-body systems


from the equilibrium distances. PES curves calculated for the
model systems [X−HE−X]− (E = S, Se; X = Cl, Br, I) feature
similar trends as those discussed above for trihalides (Figure
10).
Se−Se−Se Fragments. Recently, analogies and differences
between polychalcogenides and polyhalides, mainly tritellurides
and triiodides, have been comparatively discussed with the help
of quantum chemical DFT calculations, and it was shown that
homo- and heteronuclear trichalcogen systems behave as three-
body systems strictly related to those discussed above.41 The
main difference between isoelectronic (22 e) X3− trihalides and
E 3 4− trichalcogenides is the overall charge, with the
consequence that trihalides are stable anions, while trichalco-
genides cannot exist without a delocalization of the negative
charge. This is the reason why linear trichalcogen fragments can
only be found when the chalcogen atoms are bonded to organic
fragments or are part of metal complexes.
As a study case, we have searched the CSD looking for Se−
Figure 11. Structural data of the Se−Se−Se fragments overlapped with Se−Se linear systems. All data found are reported in Figure 11
those of the structural data reported in the scatter plot of Figure 8 as δSe−Se parameters, together with the data relative to all the
depicted as dots [E−X−Y (·); X−E−Y (·); E−X−E (·); X−X−X (·); other systems previously considered (Figure 8), showing that
Se−Se−Se (○)]. The data are reported as a scatter plot of δ1 vs δ2 all share a common correlation following application of eq 7 (k
where δ1 = δSe1−Se2 and δ2 = δSe2−Se3 (see eq 1). Fitted parameter for all = 0.153, rmsd = 0.043). Also for Se−Se−Se fragments, several
data (excluding those not considered in the fittings depicted in Figures points in the scatter plot belong to symmetric or slightly
2 and 4−8): k = 0.153; rmsd =0.043; normalized rmsd =0.057.
asymmetric systems (higher covalent contribution); however,
the points corresponding to very asymmetric systems (higher
chalcogen-dihalogen adducts has been recently supported by
applying the quantum theory of atoms in molecules (QTAIM) electrostatic contribution) are equally numerous. As observed
to a series of different adducts of selone donors with dihalogens in systems featuring hydrogen bonding,33 also in Se−Se−Se
and pseudo-halogens among which are some very polarized T- three-body systems there is no indication of a critical distance at
shaped compounds containing the I−Se−Cl, I−Se−Br, and I− which the bond switches from the substantially covalent to the
Se−CN fragments.12 In all three cases, the three-center predominantly electrostatic nature. However, also in this case,
delocalization index δ(A,B,C) introduced by Ponec and co- although the electrostatic interaction can explain the very
workers40 to better distinguish between hypervalent 3c-4e unbalanced situations satisfactorily, it becomes more and more
compounds (A−B−C) and 2c-2e systems (A−B) within the inadequate to explain the nature of the chemical bonding in
QTAIM framework indicate for all three compounds a 3c-4e more symmetric cases.
bond nature [δ(I,Se,X) = −0.203, −0.243, and −0.074 for X = Interestingly, the trend of the calculated PES curve for the
Cl, Br, and CN, respectively]. In particular, in the case of the [H2Se−HSe−SeH2]+ model compound (Figure 12) shows that
adduct with ICN featuring the fragment I−Se−CN, both the also in this case few kcal mol−1 are enough to displace the
calculated negative value of the local electronic energy density, three-body system from the point of minimum energy.
2776 dx.doi.org/10.1021/cg201328y | Cryst. Growth Des. 2012, 12, 2769−2779
Crystal Growth & Design Article
Downloaded by UNIV EASTERN FINLAND on September 8, 2015 | http://pubs.acs.org

Figure 12. Potential energy surface (PES) calculated for the hypothetical model compound [H2Se−HSe−SeH2]+. d(Se−Se) − d(Se−Se)eq
represents the differences between the Se−Se bond lengths d(Se−Se) with respect to the optimized one d(Se−Se)eq.


Publication Date (Web): May 11, 2012 | doi: 10.1021/cg201328y

CONCLUSIONS mandatory to consider the presence of a halogen bonding also


in the cases featuring balanced A−X and X−B distances. Under
Sequences of three aligned halogen, chalcogen, or mixed
such circumstances, both the CT and the 3c-4e bond models
halogen/chalcogen atoms represent recurrent motifs either as a
seem to be more appropriate to describe the nature of the
functional group in some archetypes of products ensuing from
halogen bonding, since the different match in the energies of
the reaction of dihalogens and chalcogen donors, or as
the combining orbitals is responsible for the different covalent
supramolecular assemblies in many other solid products,
and electrostatic contributions.
including polyhalides. Sequences for almost all possible
combinations of three aligned chalcogen/halogen atoms are
found in the CSD, some of them being very numerous. The
analysis of the retrieved data indicates that all systems must be
■ EXPERIMENTAL SECTION
Structural data were retrieved from CSD by means of the
considered strictly correlated, sharing the same nature in terms software Conquest v5.32.21 The nonlinear least-squares fit42 of
of chemical bonding. Except for the case of the E−X−E experimental data was carried out adopting eq 7 as a model
fragment, all the other linear three-body systems here (initial guess of k = 0.37). Theoretical calculations were carried
considered show a continuous variation of the distances of out on dihalogens X2, trihalides X3− (X = Cl, Br, and I),
the two bonds from balanced situations up to very unbalanced hypervalent [X−HE−X]− systems (E = S, Se; X = Cl, Br, and
ones without indications of critical distances at which the bonds I), charge transfer (CT) adducts HE·X2− (E = S, Se; X = Br and
switch from the substantially covalent to the predominantly I) and, for the sake of comparison, on the 22-electron three-
electrostatic nature. Indeed, by expressing all bond distances as body system [H2Se−HSe−SeH2]+ with the commercial suite of
relative variation δ with respect to the sum of the covalent radii programs Gaussian 09 package.43 On the grounds of the results
of the involved atomic species, all data belonging to systems as reported previously and comparing different functionals in
varied as CT adducts of chalcogen donors with halogens and closely related systems,2b,44 the mPW1PW functional45 was
interhalogens (Figures 2−4), hypervalent chalcogen T-shaped adopted along with the Schäfer, Horn, and Ahlrichs basis set46
adducts (Figure 5), halonium complexes of chalcogen donors for hydrogen and the completely uncontracted basis sets
(Figure 6), sequences of three halogen atoms (Figures 7−8), LANL0847 for halogen and chalcogen species with effective
can be fitted with a single semiempirical equation resulting core potentials (ECPs), supplemented by d-type polarization
from the bond-valence model. Remarkably, the same functions. In a few cases, such as for [X−HE−Cl]− (E = S, Se;
correlation can be extended to polychalcogenides, as X = Cl, Br), a quadratically convergent SCF procedure was
demonstrated for the compounds featuring Se−Se−Se frag- needed. For all species, the molecular geometry was optimized
ments, confirming the analogies in the electronic structures of (with the tight cutoffs on forces and step size) and verified by a
all the considered systems. frequency calculation, carried out by determining the second
It is important to point out that many among the three-body derivatives of the energy with respect to the Cartesian
systems considered here display a central halogen atom and coordinates of the nuclei. For three-body systems of the type
therefore represent peculiar cases of arrangements in which a X−Y−Z, PESs were subsequently carried out by calculating the
halogen bonding is operating. If we had to limit the presence of total electronic energy at geometries obtained by systematically
a halogen bonding to the cases in which one of the two bonds scanning the variation of the X−Y and Y−Z bond distances,
A−X, X−B is close to the sum of the relevant van der Waals with the geometrical constraint of keeping constant the X−Z
radii, an electrostatic model could be adequate to the distance. All other bonds, such as chalcogen−hydrogen bonds,
description of this interaction. However, the lack of were not subject to any constraints. All calculations were
discontinuity in the correlation between the two distances in carried out with an E4 Workstation equipped with four quad-
the classes of three-body arrangements considered renders it core Opteron processors and 16 GB of RAM. Molden 5.048 and
2777 dx.doi.org/10.1021/cg201328y | Cryst. Growth Des. 2012, 12, 2769−2779
Crystal Growth & Design Article

GaussView 5.049 were exploited for analyzing the results of the hypervalent does not necessarily require the involvement of a d orbital
calculations. from the central E atom in order to explain the bond nature of the


bonding in the X−E(R)−X system.14
AUTHOR INFORMATION (17) Arca, M.; Demartin, F.; Devillanova, A. F.; Garau, A.; Isaia, F.;
Lippolis, V.; Piludu, S.; Verani, G. Polyhedron 1998, 17, 3111−3119.
Corresponding Author (18) (a) Nakanishi, W. In Handbook of Chalcogen Chemistry;
*E-mail: devilla@unica.it. Devillanova, F. A., Ed.; RSC Publishing: Cambridge, 2007; Chapter
Notes 10.3, pp 644−668; (b) Bigoli, F.; Deplano, P.; Devillanova, F. A.;
The authors declare no competing financial interest. Girlando, A.; Lippolis, V.; Mercuri, M. L.; Pellinghelli, M. A.; Trogu, E.

■ REFERENCES
(1) Aragoni, M. C.; Arca, M.; Devillanova, F. A.; Garau, A.; Isaia, F.;
F. Inorg. Chem. 1996, 35, 5403−5406. (c) Devillanova, F. A.; Garau,
A.; Isaia, F.; Lippolis, V.; Verani, G.; Cornia, A.; Fabretti, A. C.;
Girlando, A. J. Mat. Chem. 2000, 1281−1286.
Lippolis, V.; Mancini, A. Bioinorg. Chem. Appl., 2006, article ID 58937, (19) (a) Arca, M.; Demartin, F.; Devillanova, A. F.; Garau, A.; Isaia,
1−12. F.; Lippolis, V.; Verani, G. J. Chem. Soc., Dalton Trans. 1999, 3069−
(2) (a) Bricklebank, N.; Skabara, P. J.; Hibbs, D. E.; Hursthouse, M. 3073. (b) Aragoni, M. C.; Arca, M.; Devillanova, A. F.; Isaia, F.;
B.; Malik, K. M. A. J. Chem. Soc., Dalton Trans. 1999, 3007−3014. Lippolis, V.; Mancini, A.; Pala, L.; Slawin, A. M.; Woollins, J. D. Chem.
(b) Aragoni, M. C.; Arca, M.; Demartin, F.; Devillanova, F. A.; Garau, Commun. 2003, 2226−2227. (c) Blake, A. J.; Devillanova, A. F.; Gould,
A.; Isaia, F.; Lelj, F.; Lippolis, V.; Verani, G. Chem.Eur. J. 2001, 7, R. O.; Li, W.-S.; Lippolis, V.; Parsons, S.; Radek, C.; Schröder, M.
3122−3133. Chem Soc. Rev. 1998, 27, 195−206. (d) Svensson, P. H.; Kloo., L.
(3) (a) Aragoni, M. C.; Arca, M.; Demartin, F.; Devillanova, F. A.; Chem. Rev. 2003, 103, 1649−1684.
Downloaded by UNIV EASTERN FINLAND on September 8, 2015 | http://pubs.acs.org

Garau, A.; Isaia, F.; Lippolis, V.; Verani, G. Trends Inorg. Chem. 1999, (20) For example, in the crystal packing of [>C−E−E−C<]+2
6, 1−18. (b) Boyle, P. D.; Godfrey, S. M. Coord. Chem. Rev. 2001, 223, dications, E−E−Nu linear arrangements are frequently observed
265−299. (Nu = nucleophile).18
Publication Date (Web): May 11, 2012 | doi: 10.1021/cg201328y

(4) (a) Bigoli, F.; Pellinghelli, A. M.; Deplano, P.; Devillanova, F. A.; (21) Cambridge Structural Database; Conquest v. 5.32; August 2011
Lippolis, V.; Mercuri, M. L.; Trogu, E. F. Gazz. Chim. It. 1994, 124, updated.
445−454. (b) Kuhn, N.; Kratz, T.; Henkel, G. Chem. Ber. 1994, 127, (22) Halogen Bonding: Fundamentals and Applications; Metrangolo,
849−851. (c) Kuhn, N.; Fawzi, R.; Kratz, T.; Steimann, M.; Henkel, G. P.; Resnati, G., Eds.; Structure and Bonding Series; Mingos, D. M. P.,
Phosphorus Sulfur Silicon Relat. Elem. 1996, 112, 225−233. Ed.; Springer-Verlag: Berlin, 2008.
(5) Aragoni, M. C.; Arca, M.; Blake, A. J.; Devillanova, F. A.; du (23) Steiner, T.; Saenger, W. Acta Crystallogr. B 1994, 50, 348−357.
Mont, W.-W.; Garau, A.; Isaia, F.; Lippolis, V.; Verani, G.; Wilson, C. (24) (a) Brown, I. D. J. Chem. Educ. 1976, 53, 100−101. (b) Brown,
Angew. Chem., Int. Ed. 2001, 40, 4229−4232. I. D. J. Chem. Educ. 1976, 53, 231−232. (c) Donnay, G.; Allman, R.
(6) Ouvrard, C.; Le Questel, J.-Y.; Berthelot, M.; Laurence, C. Acta Am. Mineral. 1970, 55, 1003−1015.
Crystallogr., Sect. B 2003, 59, 512−526. (25) Urusov, V. S.; Orlov, I. P. Crystallogr. Rep. 1999, 44, 686−760.
(7) (a) Corban, G. J.; Hadjikakou, S. K.; Hadjiliadis, N.; Kubicki, M.; (26) Burdett, J. K.; Hawthorne, F. C. Am. Mineral. 1993, 78, 884−
Tiekink, E. R. T.; Butler, I. S.; Drougas, E.; Kosmas, A. M. Inorg. Chem. 892.
2005, 44, 8617−8627. (b) Daga, V.; Hadjikakou, S. K.; Hadjiliadis, N.; (27) O’Keefe, M.; Breese, N. E. J. Am. Chem. Soc. 1991, 110, 1506−
Kubicki, M.; Santos, J. H. Z.; Butler, I. S. Eur. J. Inorg. Chem. 2002, 1510.
1718−1728. (c) Antoniadis, C. D.; Corban, G. J.; Hadjikakou, S. K.; (28) Keller, E.; Krämer, V. Acta Crystallogr. B 2006, 62, 411−416.
Hadjiliadis, N.; Kubicki, M.; Warner, S.; Butler, I. S. Eur. J. Inorg. Chem. (29) See for example: (a) Brown, I. D.; Altermatt, D. Acta
2003, 1635−1640. (d) Antoniadis, C. D.; Hadjikakou, S. K.; Crystallogr. B 1985, 41, 244−247. (b) Brese, N. E.; O’Keeffe, M. Acta
Hadjiliadis, N.; Kubicki, M.; Butler, I. S. Eur. J. Inorg. Chem. 2004, Crystallogr. B 1991, 47, 192−197. (c) Adams, S. Acta Cryst. B 2001, 57,
4324−4329. 278−287. (d) Hu, S.-Z.; Tu, L.-D.; Huang, Y.-Q.; Li, Z.-X. Inorg. Chim.
(8) (a) Bigoli, F.; Demartin, F.; Deplano, P.; Devillanova, F. A.; Isaia, Acta 1995, 232, 161−165. (e) Brown, I. D.; Gillespie, R. J.; Morgan, K.
F.; Lippolis, V.; Mercuri, M. L.; Pellinghelli, M. A.; Trogu, E. F. Inorg. R.; Tun, Z.; Ummat, P. K. Inorg. Chem. 1984, 23, 4506−4508.
Chem. 1996, 35, 3194−3201. (b) Aragoni, M. C.; Arca, M.; Demartin,
(f) Palenik, G. J. Inorg. Chem. 1997, 36, 4888−4890. (g) Kanowitz, S.
F.; Devillanova, F. A.; Garau, A.; Isaia, F.; Lippolis, V.; Verani, G. J.
M.; Palenik, G. J. Inorg. Chem. 1998, 37, 2086−2088. (h) Wood, R.
Am. Chem. Soc. 2002, 124, 4538−4539.
(9) Boyle, P. D.; Cross, W. I.; Godfrey, S. M.; McAuliffe, C. A.; M.; Palenik, G. J. Inorg. Chem. 1998, 37, 4149−4151.
(30) In the case of chalcogen atoms E and halogen atoms X (E = S,
Prichard, R. G.; Teat, S. J. Chem. Soc., Dalton Trans. 1999, 2219−2224.
(10) Rudd, M. D.; Linderman, S. V.; Husebye, S. Acta Chem. Scand. Se; X = Cl, Br, I), the BV parameters (as extracted from
1997, 51, 689−708. BVparms.cif)29b of E−X bonds and the sum of the corresponding
(11) (a) Aragoni, M. C.; Arca, M.; Devillanova, F. A.; Grimaldi, P.; covalent radii are indeed very close. Sum of covalent radii (BV
Isaia, F.; Lelj, F.; Lippolis, V. Eur. J. Inorg. Chem. 2006, 2166−2174. parameters in parentheses): S−Cl, 2.01 (2.03); S−Br, 2.16 (2.17); S−
(b) Aragoni, M. C.; Arca, M.; Demartin, F.; Devillanova, F. A.; Garau, I, 2.35 (2.36); Se−Cl, 2.15 (2.16); Se−Br, 2.30 (2.33); Se−I, 2.49
A.; Isaia, F.; Lippolis, V.; Verani, G. Dalton Trans. 2005, 2252−2258. (2.54); Cl−Br, 2.13 (2.19); Cl−I, 2.32 (2.31); Br−I, 2.47 (not
(12) Juárez-Pérez, E. J.; Aragoni, M. C; Arca, M.; Blake, A., J.; reported); Se−S, 2.18 (2.25).
Devillanova, A. F.; Garau, A.; Isaia, F.; Núñez, R.; Pintus, A.; Wilson, (31) Pauling, L. J. Am. Chem. Soc. 1947, 69, 542−553.
C. Chem.Eur. J. 2011, 17, 11497−11514. (32) Lippolis, V.; Isaia, F. in Handbook of Chalcogen Chemistry;
(13) Perkins, C. W.; Martin, J. C.; Arduengo, A. J.; Lau, W.; Alegrie, Devillanova, F. A., Ed.; RSC Publishing: Cambridge, 2007; Chapter
A.; Kocki, J. K. J. Am. Chem. Soc. 1980, 102, 7753−7759. 8.2, pp 477−499.
(14) Akiba, K.-Y. Chemistry of Hypervalent Compounds; Wiley-WCH: (33) Steiner, T. Angew. Chem., Int. Ed. 2002, 41, 48−76.
New York, 1999. (34) Hackand, R. J.; Rundle, R. E. J. Am. Chem. Soc. 1951, 73, 4321−
(15) (a) Landrum, G. A.;Goldberg, N.; Hoffmann, R.; J. Chem. Soc., 4324.
Dalton Trans., 1997, 3605−3613. (b) Aragoni, M. C.; Arca, M.; (35) Pimentel, G. C. J. Chem. Phys. 1951, 19, 446−448.
Devillanova, F. A.; Garau, A.; Isaia, F.; Lippolis, V.; Mancini, A. (36) du Mont, W.-W.; Bätcher, M.; Daniliuc, C.; Devillanova, F. A.;
Bioinorg. Chem. Appl. , 2007, article ID 17416, 1−46. Druckenbrodt, C.; Jeske, J.; Jones, P. G.; Lippolis, V.; Ruthe, F.;
(16) The notation 10-X-2 and 10-E-3 for CT and “T-shaped” Seppälä, E. Eur. J. Inorg. Chem. 2008, 29, 4562−4577.
hypervalent adducts, respectively, proposed by Arduengo et al.13 (37) Timilselvi, A.; Mugesh, G. Bioorg. Chem. Lett. 2010, 20, 3692−
identifies these compounds concisely, considering that the term 3697.

2778 dx.doi.org/10.1021/cg201328y | Cryst. Growth Des. 2012, 12, 2769−2779


Crystal Growth & Design Article

(38) Corban, G. J.; Hadjikakou, S. K.; Tsipis, A. C.; Kubicki, M.;


Bakas, N.; Hadjiliadis, N. New. J. Chem. 2011, 35, 213−224.
(39) Politzer, P.; Murray, J. S.; Clark, T. Phys. Chem. Chem. Phys.
2010, 12, 7748−7757.
(40) Ponec, R.; Mayer, I. J. Phys. Chem. 1977, 101, 1738−1741.
(41) Aragoni, M. C.; Arca, M.; Devillanova, A. F.; Isaia, F.; Lippolis,
V. Phosphorus, Sulfur Silicon Relat. Elem. 2008, 183, 1036−1045.
(42) Harris, D. C. J. Chem. Educ. 1998, 75, 119−121.
(43) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.;
Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci,
B.; Petersson, G. A.; Nakatsuji, H.; Caricato, M.; Li, X.; Hratchian, H.
P.; Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.;
Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima,
T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Montgomery, J. A. Jr.;
Peralta, J. E.; Ogliaro, F.; Bearpark, M.; Heyd, J. J.; Brothers, E.; Kudin,
K. N.; Staroverov, V. N.; Kobayashi, R.; Normand, J.; Raghavachari, K.;
Rendell, A.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Rega,
N.; Millam, J. M.; Klene, M.; Knox, J. E.; Cross, J. B.; Bakken, V.;
Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.;
Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Martin, R. L.;
Downloaded by UNIV EASTERN FINLAND on September 8, 2015 | http://pubs.acs.org

Morokuma, K.; Zakrzewski, V. G.; Voth, G. A.; Salvador, P.;


Dannenberg, J. J.; Dapprich, S.; Daniels, A. D.; Farkas, Ö .;
Foresman, J. B.; Ortiz, J. V.; Cioslowski, J.; Fox, D. J.Gaussian 09,
Revision A.1; Gaussian, Inc.: Wallingford, CT, 2009.
Publication Date (Web): May 11, 2012 | doi: 10.1021/cg201328y

(44) (a) Aragoni, M. C.; Arca, M.; Demartin, F.; Devillanova, F. A.;
Lelj, F.; Isaia, F.; Lippolis, V.; Mancini, A.; Pala, L.; Verani, G. Eur. J.
Inorg. Chem. 2004, 3099−3109. (b) Aragoni, M. C.; Arca, M.;
Devillanova, F. A.; Hursthouse, M. B.; Huth, S. L.; Isaia, F.; Lippolis,
V.; Mancini, A.; Ogilvie, H. R.; Verani, G. J. Organomet. Chem. 2005,
690, 1923−1934.
(45) Adamo, C.; Barone, V. J. Chem. Phys. 1998, 108, 664−675.
(46) Schäfer, A.; Horn, H.; Ahlrichs, R. J. Chem. Phys. 1992, 97,
2571−2577.
(47) Roy, L. E.; Hay, P. J.; Martin, R. L. J. Chem. Theory Comput.
2008, 4, 1029−1031.
(48) Schaftenaar, G.; Noordik, J. H. J. Comput.-Aided Mol. Des. 2000,
14, 123−134.
(49) Dennington, R.; Keith, T.; Millam, J. GaussView, Version 5;
Semichem Inc.: Shawnee Mission, KS, 2009.

2779 dx.doi.org/10.1021/cg201328y | Cryst. Growth Des. 2012, 12, 2769−2779

You might also like