ArticlePDF Available

Quantifying the Effects of Halogen Bonding by Haloaromatic Donors on the Acceptor Pyrimidine

Wiley
ChemPhysChem
Authors:

Abstract and Figures

The effects of intermolecular interactions by a series of haloaromatic halogen bond donors on the normal modes and chemical shifts of the acceptor pyrimidine are investigated by Raman and NMR spectroscopies and electronic structure computations. Halogen bond interactions with pyrimidine's nitrogen atoms shift normal modes to higher energy and shift 1H and 13C NMR peaks upfield in adjacent nuclei. This perturbation of vibrational normal modes is reminiscent of the effects of hydrogen bonded networks of water, methanol, or silver on pyrimidine. The unexpected observation of vibrational red shifts and downfield 13C NMR shifts in some complexes suggests that other intermolecular forces such as pi-interactions are competing with halogen bonding. Natural bond orbital analyses indicate a wide range of charge transfer from pyrimidine to different haloaromatic donors is possible and computed halogen bond binding energies can be larger than a typical hydrogen bond. These results emphasize the importance in strategic selection of substituents and electron withdrawing groups in developing supramolecular structures based on halogen bonding.
Content may be subject to copyright.
www.chemphyschem.org
Accepted Article
A Journal of
Title: Quantifying the effects of halogen bonding by haloaromatic
donors on the acceptor pyrimidine
Authors: Thomas L Ellington, Peyton L Reves, Briana L Simms, Jamey
L Wilson, Davita L Watkins, Gregory S Tschumper, and
Nathan I Hammer
This manuscript has been accepted after peer review and appears as an
Accepted Article online prior to editing, proofing, and formal publication
of the final Version of Record (VoR). This work is currently citable by
using the Digital Object Identifier (DOI) given below. The VoR will be
published online in Early View as soon as possible and may be different
to this Accepted Article as a result of editing. Readers should obtain
the VoR from the journal website shown below when it is published
to ensure accuracy of information. The authors are responsible for the
content of this Accepted Article.
To be cited as: ChemPhysChem 10.1002/cphc.201700114
Link to VoR: http://dx.doi.org/10.1002/cphc.201700114
ARTICLE
Quantifying the effects of halogen bonding by haloaromatic
donors on the acceptor pyrimidine
Thomas L. Ellington, Peyton L. Reves, Briana L. Simms, Jamey L. Wilson, Davita L. Watkins,* Gregory
S. Tschumper,* and Nathan I. Hammer*[a]
Abstract: The effects of intermolecular interactions by a series of
haloaromatic halogen bond donors on the normal modes and
chemical shifts of the acceptor pyrimidine are investigated by
Raman and NMR spectroscopies and electronic structure
computations. Halogen bond interactions with pyrimidine’s nitrogen
atoms shift normal modes to higher energy and upfield shift 1H and
13C NMR peaks in adjacent nuclei. This perturbation of vibrational
normal modes is reminiscent of the effects of hydrogen bonded
networks of water, methanol, or silver on pyrimidine. The
unexpected observation of vibrational red shifts and downfield 13C
NMR shifts in some complexes suggests that other intermolecular
forces such as interactions are competing with halogen bonding.
Natural bond orbital analyses indicate a wide range of charge
transfer is possible from pyrimidine to different haloaromatic donors
and computed halogen bond binding energies can be larger than a
typical hydrogen bond. These results emphasize the importance in
strategic selection of substituents and electron withdrawing groups
in developing supramolecular structures based on halogen bonding.
1. Introduction
The inclusion of halogen-bonded interactions in self-assembled
material building blocks has recently garnered a great deal of
attention for enhanced and selective morphological control.[1-11]
Halogen bonding is defined by IUPAC as a noncovalent
interaction involving the net attraction between an electrophilic
region of a halogen atom of a molecular entity and the
nucleophilic region of another.[12] Here, a halogen bond acceptor
is a molecule with an electron rich region, such as the nitrogen
atoms on pyrimidine, which interacts with a positive region of
electrostatic potential (ESP) on a halogen atom on an adjacent
molecule. When covalently bound to another atom, the electron
density of halogen atoms undergo an anisotropic redistribution in
which an area of positive ESP is localized on the halogen atom
and is highly directional, aligned with the R-X covalent bond
(referred to a -hole, blue region in Figure 1), leading to near
collinear intermolecular interactions.[13-18] This linearity has been
touted as a basis for the competition of halogen bonding with the
much more recognizable hydrogen bonding in complex chemical
environments.[7, 19-23] Due to the formation of a -hole, a belt-like
region of negative ESP also emerges that is perpendicular to the
R-X covalent bond, as shown in Figure 1. This nucleophilic
region can interact with Lewis acids, allowing for amphoteric
behaviour by halogen bond donors acting as both Lewis acids
and bases.[24]
The interaction strength of these highly-directional
noncovalent interactions is tuneable through either the
modification of the substituent groups, with stronger interactions
when these groups are sufficiently electron withdrawing, or the
halogen atom which results in stronger interactions with less
electronegative and more polarizable halogen atoms (I > Br > Cl >
F).[17-18] For example, Resnati and co-workers recently
demonstrated a nice correlation between the redshift in the C-I
stretching vibrational frequency in halogen bond donor molecules
of co-crystals with the careful selection of electron withdrawing
groups on pyridine-based acceptors.[25] In that case, changing the
electron withdrawing groups of the halogen bond acceptors led to
the tunability in interaction strength. Due to the strength and the
tunability of the interaction, halogen bonding is indeed becoming
more useful as a tool to direct and control molecular assembly.[1, 7,
23, 26-29] We have recently taken advantage of these interactions for
the creation of co-crystals for optoelectronic device applications.[8,
11] For example, we demonstrated that the crystal structures that
result from using pyridyl thiophene-based donors exhibit strong
halogen bond interactions, ranging from −7.5 to −8.7 kcal mol-1,
supplemented by other secondary interactions including
stacking interactions. In fact, the stacking interactions were
slightly larger in magnitude than the halogen bond interactions for
co-crystals containing (iodoethynyl)difluorobenzene.[11]
Pyrimidine (Pm, left-most structure in Figure 2) is a two-
nitrogen analogue of pyridine whose incorporation into building
blocks offers two sites for halogen bond interactions. We and
others have previously elucidated the effects of hydrogen
bonding by water, methanol, and other co-solvent molecules on
pyridine’s and Pm’s vibrational normal modes.[30-34] In particular,
[a] T. L. Ellington, P. L. Reves, B. L. Simms, J. L. Wilson, Prof. D. L.
Watkins, Prof. G. S. Tschumper, Prof. N. I. Hammer
Department of Chemistry and Biochemistry
University of Mississippi
P.O. Box 1848, University, Mississippi 38677 (United States)
E-mail: dwatkins@olemiss.edu
tschumpr@olemiss.edu
nhammer@olemiss.edu
Supporting information for this article is given via a link at the end of
the document.
Figure 1. Iodopentafluorobenzene includes a region of positive ESP indicated
by blue at the terminus of the molecule and a nucleophilic region of negative
ESP perpendicular to the R-X covalent bond (orange belt). A second
perspective with a solid surface (right) is included to clearly show both positive
and negative ESP regions. See Experimental Section for computational details.
10.1002/cphc.201700114
ChemPhysChem
This article is protected by copyright. All rights reserved.
ARTICLE
we showed that electron density is transferred from Pm’s lone
pairs out into the extended hydrogen bonded network and that
this charge redistribution leads to blue-shifting in select normal
modes with the two properties being directly correlated. We also
recently demonstrated that this effect is present in surfaced
enhanced Raman spectroscopic (SERS) interactions and that
the magnitude of charge transfer from Pm to adjacent silver
atoms is only about twice that for water alone.[35]
As revealed in an extensive review, many previous
spectroscopic studies have concentrated on analysis of the
halogen bond donor and in particular, shifting in the donor’s C-X
stretching frequencies when participating in a halogen bond.[18]
Here, in contrast, we experimentally and computationally
investigate the effects of halogen bond interactions by a series
of haloaromatic donors on the acceptor Pm and explore the
possible presence of other forces present in solution such as
stacking interactions. We also compare the strength of the
interactions and magnitude of the spectral shifts to that
previously observed for hydrogen bonded systems. The
haloaromatic series and Pm is shown in Figure 2 and includes
bromobenzene (BrB), iodobenzene (IB), bromopentafluorobenzene
(BrPFB), and iodopentafluorobenzene (IPFB). Bromo- and iodo-
based halogen bond donors are selected as the electron
acceptors based on the polarizability of the halogen atoms.
Donors, BrPFB and IPFB, were employed as the inductive effect
provided by the fluoro substituents increases the magnitude of
the -hole on halogen atomin turn, increasing the strength of
the halogen bonding interaction. In addition to shifts in Pm’s
vibrational spectrum, we also show that Pm’s 1H and 13C NMR
are very sensitive to halogen bonded interactions.
Experimental Section
Raman spectroscopy
Either a LabView-controlled Jobin-Yvon Ramanor HG2-S double
grating Raman spectrometer with photomultiplier tube detection
or a Horiba Scientific LabRAM HR Evolution Raman
Spectroscopy system with CCD camera detection was used for
the acquisition of solution phase Raman spectra. Resolution of
these instruments is much less than 1 cm-1 and their use allows
for the detection of small vibrational energy shifts due to
noncovalent interactions.[32, 34-38] Solutions only contained Pm
and the corresponding haloaromatic donors.
NMR spectroscopy
1H and 13C NMR spectra were recorded on a Bruker Avance
DRX-500 (500 MHz) spectrometer and are reported in ppm
using the solvent as an internal standard (perdeuterated
toluene) at 2.08 ppm and 20.43 ppm (CH3).
Computational Methods
Full geometry optimizations and corresponding harmonic
vibrational frequency computations with Raman activities were
performed on each monomer and each halogen bonded
complex with the hybrid M06-2X[39] density functional in
conjunction with a correlation consistent basis sets augmented
with diffuse functions (aug-cc-pVTZ).[40-41] Additionally, the basis
set for bromine[42] and iodine[43] atoms contain a small-core
energy-consistent relativistic pseudo-potential (aug-cc-pVTZ-PP).
This prescription was selected based on the extensive
calibration recently performed by Kozuch and Martin.[44] For
brevity, the previously mentioned basis sets will be referred to as
aVTZ throughout this work. Further, a full natural bond orbital
(NBO) analysis was performed on all clusters in order to assess
the magnitude of charge transfer (|q|) between pyrimidine and
its halogen bond donor. A pruned numerical integration grid
composed of 99 radial shells and 590 angular points per shell
was employed along with a threshold of < 10-9 for the RMS
change in the density matrix during the self-consistent field
procedure. The threshold for removing linear dependent basis
functions (magnitude of the eigenvalues of the overlap matrix)
was tightened from 10-6 to 10-7. All electronic energies have
been converged to at least 1 × 10-9 Eh while the Cartesian forces
of the gradient did not exceed 1 × 10-5 Eh/a0. Furthermore, pure
angular momentum (i.e., 5d, 7f, etc.) basis functions were used
instead of their Cartesian counterparts (i.e., 6d, 10f, etc.). All
computations were performed using the analytic gradients and
Hessians available in the Gaussian09 software package.[45]
Electrostatic potential maps were constructed using a total
electron density isosurface value of 0.004 electrons Bohr-3.
Optimized Cartesian coordinates of each monomer and each
halogen bonded complex can be found in the Supporting
Information.
2. Results
2.1 Raman spectroscopic results
Figure 3 shows six spectroscopic regions of Raman vibrational
spectra of solutions of Pm and IPFB. Raman spectra of Pm
interacting with BrB, IB, and BrPFB are included in the
Supporting Information (Figure S1 - S3). Four of Pm’s normal
modes,6b, 1, 9a, and 8b, unambiguously exhibit blue shifts
(shifts to higher energy) of +4, +5, +2, and +5 cm-1, respectively.
Raman spectra of solutions with the other three haloaromatic
donors do not exhibit blue shifts. In fact, solutions of Pm with IB
and BrB actually yield a slight red shift for 1. The maximum
observed blue shifts in IPFB solutions are smaller than we
reported previously for Pm hydrogen bonding to water[32] (13, 14,
5, 12 cm-1), ethylene glycol[34] (9, 14, 3, 9 cm-1), or methanol[34] (6,
12, 3, 8 cm-1). These modes involve motions of Pm’s carbon
and nitrogen atoms and their blue shifts have been previously
associated with changes in the C-N bond lengths as a result of
charge transfer. The result that blue shifts are only observed
with IPFB immediately suggests that other competing
interactions are likely present in solution that are larger in
Figure 2. From left to right: pyrimidine (Pm), bromo- (BrB) or iodobenzene
(IB), and bromo- (BrPFB) or iodopentafluorobenzene (IPFB).
10.1002/cphc.201700114
ChemPhysChem
This article is protected by copyright. All rights reserved.
ARTICLE
magnitude than the halogen bond and are playing more
dominate roles.
2.2 NMR spectroscopic results
NMR spectroscopy is commonly employed to identify the
signatures of noncovalent interactions and afford insight into the
strength of those contacts.[46-48] We complexed each halogen
bond donor with Pm at a 1:2 ratio in halogen bond promoting
solvent (i.e., 250 mM; toluene-d8). As the electron density at the
nitrogen atom changes due to halogen bond formation, the
nuclei bewteen the two nitrogen atoms (i.e., the C-H group)
becomes slightly more shielded. This shielding is observed as a
difference in chemical shift of a few hundreths of a ppm when
comparing the spectrum of neat Pm to that of a 1:2 complex of
the halogen bond acceptor and donor. Figure 4 shows the 1H
and 13C NMR spectra for the resonance signals that correspond
to the hydrogen and carbon (1H = 9.163 and 13C = 159.580 ppm)
residing between Pm’s nitrogen atoms. In the case of the 1H
NMR, essentially no perturbation is observed when Pm is
interacting with BrB or IB. While for halogen bonded complexes
of Pm with either BrPFB or IPFB, an upfield shift is observed
with IPFB showing the largest chemical shift difference (9.163 to
9.136 ppm). Conventionally, attention is given to the 13C NMR
spectra as carbon atoms are less susceptible to media (i.e.,
solvent, pH, etc.) effects.[46, 49] The carbon spectra for the
halogen bonding complexes show a similar trend to that of the
proton; however, an unexpected downfield shift is observed with
BrPFB (159.580 to 159.592 ppm). Such deviations are not
unprecedented. For example, Goroff and co-workers reported an
unusual effect in the 13C NMR spectra of iodoalkynes.[49] In this
case, the authors noted - experimentally and computationally -
that combinations of solvent and bonding environment promote
dramatic changes in the 13C NMR spectrum making it
significantly different than anticipated. Our results were
confirmed by changing the solvent from toluene-d8 to benzene-
d6 and also observing a similar trend with pyridine as the
halogen bond donor (Figure S4 - S11 of Supporting Information).
Nonetheless, the largest change in the 13C chemical shifts are
for IPFB from 159.580 to 159.509 ppm.
2.3 Computational results
Figure 5 shows optimized molecular geometries for Pm
interacting with one or two molecules of BrB, IB, BrPFB or IPFB
at the M06-2X/aVTZ level of theory. The addition of a second
halobenzene or halopentafluorobenzene is made possible due
the two nitrogen atoms of Pm. Through a vibrational analysis,
each optimized structure reported was verified as a minimum (ni
= 0) on the M06-2X/aVTZ potential energy surface. Of the
structures characterized in this work, the global minimum of
Pm/BrB and of Pm/IB are C1 structures that deviate only slightly
from a perfectly perpendicular Cs orientation of the two rings and
have halogen bond intermolecular separations of 3.11 and 3.14
Å, respectively, which correspond to reductions of the sum of the
van der Waals radii (rsvdW) of 8.5% and 11.0% relative to the
sum of nitrogen (1.55 Å) and the halogen (Br: 1.85 Å, I: 1.98 Å)
van der Waals radii.[50] A second low-energy co-planar Cs isomer
of Pm/BrB (+0.12 kcal mol-1) and of Pm/IB (+0.09 kcal mol-1) was
also found and with similar halogen bond intermolecular
separations of 3.10 Å (8.8%) and 3.11 Å (11.9%). In contrast,
the global minimum of each pentafluorohalobenzene (Pm/BrPFB
and Pm/IPFB) prefers Cs symmetry and has slightly shorter
halogen bond intermolecular separations of 2.93 Å (13.8%) and
2.95 Å (16.4%), respectively. For the case of Pm/BrPFB, a C1
symmetric analogue (+0.15 kcal mol-1) was found with the same
halogen bond intermolecular separation of 2.93 Å (13.8%). This
C1 symmetric analogue of Pm/IPFB collapses to the Cs
symmetric structure described above. The addition of a second
halo- or halopentafluorobenzene, leads to various low-energy
isomers whose relative energies (E) doesn’t exceed 0.25 kcal
Figure 4. 1H (left) and 13C (right) NMR spectra of the proton and carbon atom
of pyrimidine indicated by the circle.
10.1002/cphc.201700114
ChemPhysChem
This article is protected by copyright. All rights reserved.
ARTICLE
mol-1. Furthermore, changes in intermolecular separations and
binding energies per haloaromatic donor are similar to those
seen in complexes with a single donor, suggesting that
cooperative effects are minimal. The key feature differentiating
the 1:1 and 1:2 Pm:haloaromatic complexes is whether the rings
are co-planar or perpendicular.
For Pm/(IB)2 a non-planar Cs symmetric structure was
found to be the global minimum at the M06-2X/aVTZ level of
theory with a halogen bond intermolecular separation of 3.15 Å
(10.8%). A second energetically competitive structure of C2
symmetry was also found, though it was determined to be a
transition state (ni = 1) on the M06-2X/aVTZ potential energy
surface. Further, a third low-energy planar C2v local minimum
(+0.17 kcal mol-1) was found and has a halogen bond
intermolecular separation of 3.14 Å (11.0%). Upon fluorination,
the global minimum was identified as a planar C2v symmetric
structure with a halogen bond intermolecular separation of 2.96
Å (16.2%). Additionally, a slightly higher energy C2 isomer
(+0.22 kcal mol-1) was found to be a local minimum at this level
of theory with a halogen bond intermolecular separation of 2.98
Å (15.6%). A summary of these results along with computed
binding energies and results of the NBO calculations can be
found in Table 1. Table 2 compares the maximum experimental
frequency shifts with those computed for the molecular
complexes.
3. Discussion
The result here that four vibrational modes of Pm are observed
to blue shift when complexed with IPFB is reminiscent of the
effects of hydrogen bonding by water or methanol with Pm.
Earlier, we demonstrated that these shifts are directly
proportional to the magnitude of partial charge transfer from Pm
to the hydrogen bond donors.[34] Here, we also see perturbations
in the NMR spectra of Pm when complexed with the halogen
bond donors.
Traditionally, techniques such as 15N[51] and 19F NMR[52] as
well as solid state nuclear magnetic resonance (SSNMR)[53] are
employed to investigate displacements in chemical shifts due to
halogen bonding. However, these techniques are less practical
when compared to the standard 1H and 13C NMR where
specialized chemicals (e.g., isotopic labelling) and large sample
quantities are not necessary to achieve quality data. Studies
employing solution-phase NMR demonstrate that not only can
the presence of a halogen bond be determined but the relative
strength of the intermolecular interaction of various halogen
bond donors and their corresponding acceptors can also be
elucidated.[26, 54-59] For example, the measure of efficiency for a
halogen bond donor (Cl < Br < I) was originally established by
using 19F NMR spectroscopy.[60-61] Fluorinated halogen bond
Figure 5. Optimized molecular geometries of pyrimidine interacting with one or two molecules of BrB, IB, BrPFB, or IPFB at the M06-2X/aVTZ level of theory.
Table 1.
Point group symmetries, binding energies (Ebind), relative energies
(E), halogen bond intermolecular separations
)
, van der Waals radii
reduction upon complexation (rsvdW, %), and magnitude of charge transfer
(|q|, me) from pyrimidine upon complexation. Energies are reported in kcal
mol-1.
Complex
Symmetry
Ebind
E
RX···N
rsvdW
|q|
Pm/BrB
C1
2.27
0.00
3.11
8.5
6.2
Cs
2.16
+0.12
3.10
8.8
Pm/IB
C1
3.38
0.00
3.14
11.0
13.2
Cs
3.29
+0.09
3.11
11.9
Pm/BrPFB
Cs
4.04
0.00
2.93
13.8
15.7
C1
3.89
+0.15
2.93
13.8
Pm/IPFB
Cs
5.83
0.00
2.95
16.4
30.0
Pm/(BrB)2-A
C2
4.56
0.00
3.11
8.5
12.6
Pm/(BrB)2-B
C2
4.32
+0.24
3.10
8.8
Pm/(IB)2
Cs
6.69
0.00
3.15
10.8
24.5
C2v
6.52
+0.17
3.14
11.0
Pm/(BrPFB)2
C2v
7.88
0.00
2.95
13.2
27.0
Pm/(IPFB)2
C2v
11.03
0.00
2.96
16.2
50.5
C2
10.81
+0.22
2.98
15.6
Ebind = E(Complex) E(Pm) E(halo- or haloperfluorobenzene)
ΔE = Ebind(local minimum) Ebind(global minimum)
10.1002/cphc.201700114
ChemPhysChem
This article is protected by copyright. All rights reserved.
ARTICLE
acceptors result in definitively stronger interactions via an
inductive effect and result in large high-field shifts in chemical
signals upon halogen bonding formation.
Although solution-based NMR spectroscopy has shown to
be a convenient method towards detecting the occurrence of
halogen bonding, these studies are modest and place particular
emphasis on the halogen bond donor.[26] Nitrogen-containing
molecules like Pm are ubiquitous in self-assembling complexes,
yet their characterization within halogen bond assemblies has
been limited.[48, 62] This lack of data on nitrogen-containing
halogen bonded acceptors is mostly due to the complexity of the
molecules and strengths of interaction, affording less distinctive
differences when compared to that of non-interacting building
blocks.[49] Here, we employ our knowledge of the Pm structure to
note significant perturbations in the NMR spectra of Pm when
complexed with IPFB. The substantial difference in chemical
shift is most likely due to the polarizability of iodine and the
inductive effect of the fluoro groups increasing the localized area
of electron density on the atom. As the lone pairs on Pm’s
nitrogen atoms interact with the region of positive electrostatic
potential on the iodine atom in IPFB, the electron density of the
nitrogen atoms decrease as some of this density is transferred.
The NMR peaks corresponding to the proton and carbon nuclei
between the two nitrogen atoms thus becomes more shielded
(i.e., spatial shielding)[63] with complexation and are the most
affected. Although the deviations in NMR chemical shifts are
small (<1 ppm), they can be accounted for by weak
intermolecular interactions between the solvent and solute that
conceivably screen the effects of halogen bonding. Nonetheless,
the trend is consistent with the blue shifts observed in the
experimental Raman spectra.
Here, |q| for Pm/(IPFB)2 is computed to be significantly
larger than that for any of the other complexes. This
corresponds to the largest shifts in the NMR spectra and the
only observed blue shifts in the experimental Raman spectra.
Furthermore, as seen in Tables 1 and 2, the |q| and rsvdW
results of halogen bonded complexes display a clear trend and
track the increasing interaction strength. These results might
appear somewhat inconsistent with the computed vibrational
results in Table 2, which suggest that blue shifts should also be
observed for the other haloaromatic donors as well, especially
BrPFB. Other intermolecular forces, such as -type interactions,
are known to compete with halogen bonding in the condensed
phases[8, 11] and are likely competing here with halogen bonding
in these solutions. The numerical agreement between the
experimentally observed blue shifts for IPFB solutions and the
computed shifts for the 1:1 Pm:IPFB complex suggests that the
halogen bond interaction begins to dominate this interaction as
the halogen bond strength increases. The unexpected red shift
in Pm’s 1 mode with BrB and IB and the downfield shift in the
13C NMR spectrum of BrPFB, however, also suggest that other
interactions such as -stacking are competing with halogen
bonding.
In the cases of similar 1:2 structures of Pm interacting with
water or methanol, |q| at the B3LYP/6-311++(2df,2pd) level of
theory was previously reported to be 38 me in both cases.[32, 34]
At the M06-2X/aVTZ level of theory, |q| for these 1:2
complexes is 31 me for water and is 34 me for methanol. This
compares to 51 me for the Pm/(IPFB)2 structure computed here
at the M06-2X/aVTZ level of theory. The magnitudes of |q| for
Pm/(IB)2, Pm/IPFB, Pm/(BrPFB)2 are computed to be much
smaller, lying between 24 and 30 me with binding energies
between approximately 3 and 4 kcal mol-1. The binding energy
for Pm/(IPFB)2, on the other hand, is computed to be over 5 kcal
mol-1, similar in magnitude to the strength of hydrogen bonding.
Together, these results suggest that, in solution, an appreciable
interaction strength may be required for the halogen bond
interaction to dominate over competing molecular forces. For
example, Zarić and co-workers demonstrate the significance of
perturbations involving -stacking interactions and other
competing noncovalent interactions in pyridine and hydrated
pyridine.[64] We also recently showed that although hydrogen
bond interactions are possible in similar systems, only -
stacking interactions are energetically competitive with halogen
bonding bonding. For example, in that study we found that C-
H∙∙∙F interactions are only about 1 kcal mol-1, roughly 1/8th the
magnitude of halogen interaction energies.[11] Herrebout and co-
workers previously reported vibrational red shifts of benzene's
ring breathing mode when CF3X (where X=Cl,Br,I) is the
halogen bond donor and benzene is the acceptor. Such C-X∙∙∙
interactions would explain the red shifts observed here in Pm’s
ring breathing mode when complexed with BrPFB and Pm.
Wang and co-workers also previously showed that as the
halogen bond strength between pyridine and different
haloaromatic halogen bond donors increases, the - interaction
also strengthens.[65] Here, of all the Br containing complexes, the
lowest energy 1:2 Pm/(BrPFB)2 complex is the only one that
exhibits coplanarity, which could allow it to more easily form -
interactions and could explain the downfield shift in the 13C NMR.
Although the lowest energy 1:2 Pm/(IPFB)2 complex is also
coplanar, its halogen bond interaction strength is much greater.
Table 2. Select experimental and computed Raman vibrational frequencies (in cm-1) and their subsequent perturbations upon complexation. The original locations
are the experimental Raman vibrational frequencies of pyrimidine before complexation.
Mode
Original
Location
Bromobenzene
Iodobenzene
Bromopentafluorobenzene
Iodopentafluorobenzene
BrB
(BrB)2
Expt
IB
(IB)2
Expt
BrPFB
(BrPFB)2
Expt
IPFB
(IPFB)2
Expt
ν6b
626
+2
+4
1
+2
+5
1
+4
+8
0
+5
+11
+4
ν6a
681
+1
+1
1
+1
+2
0
+1
+2
0
+1
+2
+1
ν1
990
+2
+3
2
+4
+7
2
+4
+9
0
+6
+10
+5
ν9a
1139
+2
+4
0
+1
+2
1
+1
+3
+1
+1
+2
+2
ν8a
1564
+1
+3
1
0
+1
0
0
+6
0
1
+1
+2
ν8b
1565
+2
+4
0
+3
+4
0
+4
+8
0
+7
+8
+6
10.1002/cphc.201700114
ChemPhysChem
This article is protected by copyright. All rights reserved.
ARTICLE
Conclusions
The effects of intermolecular interactions by a series of
haloaromatic halogen bond donors on the vibrational normal
modes and NMR chemical shifts of the acceptor Pm have been
investigated by both Raman and NMR spectroscopies, as well
as electronic structure computations. Strong halogen bond
interactions were observed to induce upfield shifts in NMR
spectra and blue shifts in certain vibrational normal modes of
Pm. The consistency between the NMR and Raman results
suggest that these methods may be very sensitive probes for
detecting such weak interactions in the condensed phases.
These results also suggest that there is a delicate balance of
intermolecular interactions competing in solution that could
affect solid state crystal structures and that a relatively strong
interaction strength on the order of a hydrogen bond is needed
for halogen bonding to dominate and govern observed
spectroscopic shifts.
Acknowledgements
This material (experimental part) is based upon work supported
by the National Science Foundation under Grant Numbers CHE-
0955550 and CHE-1532079 (N.I.H). This material
(computational part) is based on work supported by the
Mississippi Center of Supercomputing Research and the
National Science Foundation under Grant Numbers CHE-
1338056 and IIA-1430364 (G.S.T). D.L.W. appreciates financial
support of this work from Oak Ridge Associated Universities
through the Ralph E. Powe Award. We would like to also thank
John T. Kelly for contributions to this work.
Keywords: halogen bonding • ab initio calculations vibrational
spectroscopy • Raman spectroscopy • NMR spectroscopy
[1] P. Metrangolo, F. Meyer, T. Pilati, G. Resnati, G.
Terraneo, Angew. Chem. Int. Ed. 2008, 47, 6114-6127.
[2] A. Abate, S. Biella, G. Cavallo, F. Meyer, H. Neukirch,
P. Metrangolo, T. Pilati, G. Resnati, G. Terraneo, J.
Fluorine Chem. 2009, 130, 1171-1177.
[3] E. Cariati, G. Cavallo, A. Forni, G. Leem, P.
Metrangolo, F. Meyer, T. Pilati, G. Resnati, S. Righetto,
G. Terraneo, E. Tordin, Cryst. Growth Des. 2011, 11,
5642-5648.
[4] J. Marti-Rujas, L. Colombo, J. Lu, A. Dey, G. Terraneo,
P. Metrangolo, T. Pilati, G. Resnati, Chem. Commun.
2012, 48, 8207-8209.
[5] A. Priimagi, G. Cavallo, P. Metrangolo, G. Resnati, Acc.
Chem. Res. 2013, 46, 2686-2695.
[6] V. Vasylyeva, S. K. Nayak, G. Terraneo, G. Cavallo, P.
Metrangolo, G. Resnati, CrystEngComm 2014, 16,
8102-8105.
[7] A. Mukherjee, S. Tothadi, G. R. Desiraju, Acc. Chem.
Res. 2014, 47, 2514-2524.
[8] J. Wilson, J. S. Dal Williams, C. Petkovsek, P. Reves,
J. W. Jurss, N. I. Hammer, G. S. Tschumper, D. L.
Watkins, RSC Adv. 2015, 5, 82544-82548.
[9] A. Vanderkooy, M. S. Taylor, J. Am. Chem. Soc. 2015,
137, 5080-5086.
[10] O. S. Bushuyev, D. Tan, C. J. Barrett, T. Friscic,
CrystEngComm 2015, 17, 73-80.
[11] S. T. Nguyen, A. L. Rheingold, G. S. Tschumper, D. L.
Watkins, Cryst. Growth Des. 2016, 16, 6648-6653.
[12] G. R. Desiraju, P. Shing Ho, L. Kloo, A. C. Legon, R.
Marquardt, P. Metrangolo, P. Politzer, G. Resnati, K.
Rissanen, Pure Appl. Chem. 2013, 85, 1711-1713.
[13] J. P. M. Lommerse, A. J. Stone, R. Taylor, F. H. Allen,
J. Am. Chem. Soc. 1996, 118, 3108-3116.
[14] T. Clark, M. Hennemann, J. S. Murray, P. Politzer, J.
Mol. Model. 2007, 13, 291-296.
[15] P. Politzer, P. Lane, M. C. Concha, Y. Ma, J. S. Murray,
J. Mol. Model. 2007, 13, 305-311.
[16] P. Politzer, J. S. Murray, T. Clark, Phys. Chem. Chem.
Phys. 2010, 12, 7748-7757.
[17] C. Wang, D. Danovich, Y. Mo, S. Shaik, J. Chem.
Theory Comput. 2014, 10, 3726-3737.
[18] G. Cavallo, P. Metrangolo, R. Milani, T. Pilati, A.
Priimagi, G. Resnati, G. Terraneo, Chem. Rev. 2016,
116, 2478-2601.
[19] P. Metrangolo, H. Neukirch, T. Pilati, G. Resnati, Acc.
Chem. Res. 2005, 38, 386-395.
[20] P. Politzer, J. S. Murray, P. Lane, Int. J. Quantum
Chem 2007, 107, 3046-3052.
[21] C. B. Aakeröy, M. Fasulo, N. Schultheiss, J. Desper, C.
Moore, J. Am. Chem. Soc. 2007, 129, 13772-13773.
[22] A. Priimagi, G. Cavallo, A. Forni, M. Gorynsztejn
Leben, M. Kaivola, P. Metrangolo, R. Milani, A.
Shishido, T. Pilati, G. Resnati, G. Terraneo, Adv. Funct.
Mater. 2012, 22, 2572-2579.
[23] F. M. Amombo Noa, S. A. Bourne, H. Su, E. Weber, L.
R. Nassimbeni, Cryst. Growth Des. 2016, 16, 4765-
4771.
[24] P. Metrangolo, G. Resnati, IUCrJ 2014, 1, 5-7.
[25] V. Vasylyeva, L. Catalano, C. Nervi, R. Gobetto, P.
Metrangolo, G. Resnati, CrystEngComm 2016, 18,
2247-2250.
[26] P. Metrangolo, G. Resnati, Chem. Eur. J. 2001, 7,
2511-2519.
[27] X. Ding, M. Tuikka, M. Haukk, in Recent Advances in
Crystallography, InTech, 2012, pp. 143-168.
[28] C. B. Aakeröy, M. Baldrighi, J. Desper, P. Metrangolo,
G. Resnati, Chem. Eur. J. 2013, 19, 16240-16247.
[29] M. R. Scholfield, C. M. Zanden, M. Carter, P. S. Ho,
Protein Sci. 2013, 22, 139-152.
[30] S. Schlücker, R. K. Singh, B. P. Asthana, J. Popp, W.
Kiefer, J. Phys. Chem. A 2001, 105, 9983-9989.
[31] S. Schlücker, J. Koster, R. K. Singh, B. P. Asthana, J.
Phys. Chem. A 2007, 111, 5185-5191.
[32] A. A. Howard, G. S. Tschumper, N. I. Hammer, J. Phys.
Chem. A 2010, 114, 6803-6810.
[33] J. C. Howard, N. I. Hammer, G. S. Tschumper,
ChemPhysChem 2011, 12, 3262-3273.
[34] A. M. Wright, A. A. Howard, J. C. Howard, G. S.
Tschumper, N. I. Hammer, J. Phys. Chem. A 2013,
117, 5435-5446.
[35] J. T. Kelly, A. K. McClellan, L. V. Joe, A. M. Wright, L.
T. Lloyd, G. S. Tschumper, N. I. Hammer,
ChemPhysChem 2016, 17, 2782-2786.
[36] K. L. Munroe, D. H. Magers, N. I. Hammer, J. Phys.
Chem. B 2011, 115, 7699-7707.
[37] K. A. Cuellar, K. L. Munroe, D. H. Magers, N. I.
Hammer, J. Phys. Chem. B 2014, 118, 449-459.
[38] A. M. Wright, L. V. Joe, A. A. Howard, G. S.
Tschumper, N. I. Hammer, Chem. Phys. Lett. 2011,
501, 319-323.
[39] Y. Zhao, D. G. Truhlar, Theor. Chem. Acc. 2008, 120,
215-241.
[40] T. H. Dunning, J. Chem. Phys. 1989, 90, 1007-1023.
[41] R. A. Kendall, T. H. Dunning, R. J. Harrison, J. Chem.
Phys. 1992, 96, 6796-6806.
[42] K. A. Peterson, D. Figgen, E. Goll, H. Stoll, M. Dolg, J.
Chem. Phys. 2003, 119, 11113-11123.
[43] K. A. Peterson, B. C. Shepler, D. Figgen, H. Stoll, J.
Phys. Chem. A 2006, 110, 13877-13883.
10.1002/cphc.201700114
ChemPhysChem
This article is protected by copyright. All rights reserved.
ARTICLE
[44] S. Kozuch, J. M. Martin, J. Chem. Theory Comput.
2013, 9, 1918-1931.
[45] M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E.
Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani,
V. Barone, B. Mennucci, G. A. Petersson, H. Nakatsuji,
M. Caricato, X. Li, H. P. Hratchian, A. F. Izmaylov, J.
Bloino, G. Zheng, J. L. Sonnenberg, M. Hada, M.
Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida,
T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven,
J. A. Montgomery Jr., J. E. Peralta, F. Ogliaro, M. J.
Bearpark, J. Heyd, E. N. Brothers, K. N. Kudin, V. N.
Staroverov, R. Kobayashi, J. Normand, K.
Raghavachari, A. P. Rendell, J. C. Burant, S. S.
Iyengar, J. Tomasi, M. Cossi, N. Rega, N. J. Millam, M.
Klene, J. E. Knox, J. B. Cross, V. Bakken, C. Adamo, J.
Jaramillo, R. Gomperts, R. E. Stratmann, O. Yazyev, A.
J. Austin, R. Cammi, C. Pomelli, J. W. Ochterski, R. L.
Martin, K. Morokuma, V. G. Zakrzewski, G. A. Voth, P.
Salvador, J. J. Dannenberg, S. Dapprich, A. D. Daniels,
Ö. Farkas, J. B. Foresman, J. V. Ortiz, J. Cioslowski, D.
J. Fox, Gaussian, Inc., Wallingford, CT, USA, 2009.
[46] R. Glaser, N. Chen, H. Wu, N. Knotts, M. Kaupp, J. Am.
Chem. Soc. 2004, 126, 4412-4419.
[47] A.-C. C. Carlsson, M. Uhrbom, A. Karim, U. Brath, J.
Gräfenstein, M. Erdélyi, CrystEngComm 2013, 15,
3087.
[48] N. Han, Y. Zeng, C. Sun, X. Li, Z. Sun, L. Meng, J.
Phys. Chem. A 2014, 118, 7058-7065.
[49] P. D. Rege, O. L. Malkina, N. S. Goroff, J. Am. Chem.
Soc. 2002, 124, 370-371.
[50] A. Bondi, J. Phys. Chem. 1964, 68, 441-451.
[51] P. Cerreia Vioglio, L. Catalano, V. Vasylyeva, C. Nervi,
M. R. Chierotti, G. Resnati, R. Gobetto, P. Metrangolo,
Chem. Eur. J. 2016, 22, 16694-16694.
[52] M. G. Sarwar, B. Dragisic, L. J. Salsberg, C. Gouliaras,
M. S. Taylor, J. Am. Chem. Soc. 2010, 132, 1646-1653.
[53] P. Cerreia Vioglio, L. Catalano, V. Vasylyeva, C. Nervi,
M. R. Chierotti, G. Resnati, R. Gobetto, P. Metrangolo,
Chem. Eur. J. 2016, 22, 16819-16828.
[54] P. Cardillo, E. Corradi, A. Lunghi, S. Valdo Meille, M.
Teresa Messina, P. Metrangolo, G. Resnati,
Tetrahedron 2000, 56, 5535-5550.
[55] A. De Santis, A. Forni, R. Liantonio, P. Metrangolo, T.
Pilati, G. Resnati, Chem. Eur. J. 2003, 9, 3974-3983.
[56] K. Bouchmella, B. Boury, S. G. Dutremez, A. van der
Lee, Chem. Eur. J. 2007, 13, 6130-6138.
[57] S. Libri, N. A. Jasim, R. N. Perutz, L. Brammer, J. Am.
Chem. Soc. 2008, 130, 7842-7844.
[58] A.-C. C. Carlsson, J. Gräfenstein, A. Budnjo, J. L.
Laurila, J. Bergquist, A. Karim, R. Kleinmaier, U. Brath,
M. Erdélyi, J. Am. Chem. Soc. 2012, 134, 5706-5715.
[59] R. A. Thorson, G. R. Woller, Z. L. Driscoll, B. E. Geiger,
C. A. Moss, A. L. Schlapper, E. D. Speetzen, E. Bosch,
M. Erdélyi, N. P. Bowling, Eur. J. Org. Chem. 2015,
2015, 1685-1695.
[60] M. T. Messina, P. Metrangolo, W. Panzeri, E. Ragg, G.
Resnati, Tetrahedron Lett. 1998, 39, 9069-9072.
[61] A. Lunghi, P. Cardillo, T. Messina, P. Metrangolo, W.
Panzeri, G. Resnati, J. Fluorine Chem. 1998, 91, 191-
194.
[62] A.-C. C. Carlsson, J. Grafenstein, J. L. Laurila, J.
Bergquist, M. Erdelyi, Chem. Commun. 2012, 48,
1458-1460.
[63] J. W. Emsley, J. Feeney, L. H. Sutcliffe, in High
Resolution Nuclear Magnetic Resonance
Spectroscopy, Pergamon, 1966, pp. 665-870.
[64] D. B. Ninković, G. V. Janjić, S. D. Zarić, Crystal Growth
& Design 2012, 12, 1060-1063.
[65] W. Wang, Y. Zhang, Y.-B. Wang, J. Phys. Chem. A
2012, 116, 12486-12491.
10.1002/cphc.201700114
ChemPhysChem
This article is protected by copyright. All rights reserved.
ARTICLE
Entry for the Table of Contents
ARTICLE
Affecting the acceptor: The effects
of halogen bond interactions on the
acceptor pyrimidine are elucidated
using Raman and NMR
spectroscopies. Changes in
vibrational normal modes and
chemical shifts stem from charge
transfer to haloaromatic donors and
interactions. Computations reveal
some halogen bond binding energies
larger than a typical hydrogen bond.
Thomas L. Ellington, Peyton L. Reves,
Briana L. Simms, Jamey L. Wilson,
Davita L. Watkins,* Gregory S.
Tschumper,* and Nathan I. Hammer*
Quantifying the effects of halogen
bonding by haloaromatic donors on
the acceptor pyrimidine
((Insert TOC Graphic here))
10.1002/cphc.201700114
ChemPhysChem
This article is protected by copyright. All rights reserved.
... However, examination of this question has been fragmentary, with little in the way of general trends emerging from past work. A large fraction of the past work concentrated on the effects of the XB upon the electron donor unit [39][40][41][42], meaning it ignored issues arising within the acid. A few cases have been identified where XB formation leads to a red shift in the internal stretching frequency within the halogen donor, usually in small molecules such as a dihalogen [43,44], FX [45,46], CH 3 X [47], or CF 3 X [48], and there are cases where a blue shift has been observed [49], but little systematic work has addressed this issue in larger systems. ...
... A number of works have considered very small Lewis acids such as dihalogens [45,46,[51][52][53]. NMR coupling constants have been computed for the specific pair of atoms involved directly in the bond [39,[54][55][56] but little attention has been paid to the more peripheral nuclei or to the coupling constants within the Lewis acid unit. In connection with larger systems, a certain amount of attention has been drawn toward halobenzenes [57][58][59][60][61][62] where the X atom is connected to a simple phenyl ring. ...
Article
Full-text available
The relationship between the strength of a halogen bond (XB) and various IR and NMR spectroscopic quantities is assessed through DFT calculations. Three different Lewis acids place a Br or I atom on a phenyl ring; each is paired with a collection of N and O bases of varying electron donor power. The weakest of the XBs display a C–X bond contraction coupled with a blue shift in the associated frequency, whereas the reverse trends occur for the stronger bonds. The best correlations with the XB interaction energy are observed with the NMR shielding of the C atom directly bonded to X and the coupling constants involving the C–X bond and the C–H/F bond that lies ortho to the X substituent, but these correlations are not accurate enough for the quantitative assessment of energy. These correlations tend to improve as the Lewis acid becomes more potent, which makes for a wider range of XB strengths.
Article
Full-text available
As a prototypical molecule in the important class of halopyrimidines, 2-chloropyrimidine has been the subject of numerous spectroscopic studies. However, its absorption spectrum under vacuum ultraviolet (VUV) radiation has not yet been reported. Here, we close this gap by presenting high-resolution VUV photoabsorption cross sections in the 3.7–10.8 eV range. Based on time-dependent density functional theory (TDDFT) calculations performed within the nuclear ensemble approach, we are able to characterize the main features of the measured spectrum. By comparing the present results for 2-chloropyrimidine with those of 2-bromopyrimidine and pyrimidine, we find that the effect of the halogen atom increases remarkably with the photon energy. The two lowest-lying absorption bands are overall similar for the three molecules, apart from some differences on the vibrational progressions in band I (3.7–4.6 eV) and minor energy shifts in band II (4.6–5.7 eV). Larger shifts appear in band III (5.7–6.7 eV), especially when comparing pyrimidine with the two halogenated species. The three molecules absorb more strongly in the region of band IV (6.7–8.2 eV), where the bands look qualitatively different because the mixing of excited configurations is strongly dependent on the species. At higher energies (8.2–10.8 eV) the three spectra no longer resemble each other. An important finding of this study is the very satisfactory comparison between experiment and theory, as the combination of TDDFT calculations with the nuclear ensemble approach yields cross sections much closer to experiment than the simpler vertical approximation, in shape and magnitude, and across the whole spectral range surveyed here.
Article
This Article revisits the “Definition of the Halogen Bond (IUPAC Recommendations 2013)” [Desiraju, G. R. Pure Appl. Chem. 2013, 85 (8), 1711–1713], recommendations that fail to include the fundamental, underlying concept of (electrophilic) σ- and p-/π-hole theory and orbital-based charge transfer interactions that accompany halogen bond formation. An electrophilic σ-hole, or p-/π-hole, is an electron-density-deficient region of positive polarity (and positive potential) on the electrostatic surface on the side of halogen along, or orthogonal to, a covalently bonded halogen in a molecular entity that leads to the development of a noncovalent interaction─a halogen bond─when in close proximity to an electron-density-rich nucleophilic region on the same or another identical or different molecular entity, with which it interacts. This Article re-examines the characteristic features of the halogen bond and lists a wide variety of donors and acceptors that participate in halogen bonding. We add caveats that are essential for identifying halogen bonding in chemical systems, necessary for the appropriate use of the terminologies involved. Illustrative examples of chemical systems that feature inter- and intramolecular halogen bonds and other noncovalent interactions in the crystalline phase are given, together with a case study of some dimer systems using first-principles calculations. We also point out that the π-hole/belt (or p-hole/belt) that may develop on the surface of a halogen derivative in halogenated molecules may be prone to forming a π-hole/belt (or p-hole/belt) halogen bond when in close proximity to nucleophiles on another similar or different molecular entity.
Article
Full-text available
The relative contributions of halogen and hydrogen bonding to the interaction between graphitic carbon nitride monomers and halogen bond (XB) donors containing C−X and C≡C bonds were evaluated using computational vibrational spectroscopy. Conventional probes into select vibrational stretching frequencies can often lead to disconnected results. To elucidate this behavior, local mode analyses were performed on the XB donors and complexes identified previously at the M06‐2X/aVDZ‐PP level of theory. Due to coupling between low and high energy C−X vibrations, the C≡C stretch is deemed a better candidate when analyzing XB complex properties or detecting XB formation. The local force constants support this conclusion, as the C≡C values correlate much better with the σ‐hole magnitude than their C−X counterparts. The intermolecular local stretching force constants were also assessed, and it was found that attractive forces other than halogen bonding play a supporting role in complex formation.
Article
The potential formation of halogen bonded complexes between a donor, heptafluoro-2-iodopropane (HFP), and the three acceptor heterocyclic azines (azabenzenes: pyridine, pyrimidine, and pyridazine) is investigated herein through normal mode analysis via Raman spectroscopy, density functional theory, and natural electron configuration analysis. Theoretical Raman spectra of the halogen bonded complexes are in good agreement with experimental data providing insight into the Raman spectra of these complexes. The exhibited shifts in vibrational frequency of as high as 8 cm-1 for each complex demonstrate, in conjunction with NEC analysis, significant evidence of charge transfer from the halogen bond acceptor to donor. Here, an interesting charge flow mechanism is proposed involving the donated nitrogen lone pair electrons pushing the dissociated fluorine atoms back to their respective atoms. This mechanism provides further insight into the formation and fundamental nature of halogen bonding and its effects on neighboring atoms. The present findings provide novel and deeper characterization of halogen bonding with applications in supramolecular and organometallic chemistry.
Article
The effects of trimethylamine N‐oxide (TMAO) on guanidinium chloride and hydrogen‐bonded networks of water are explored in this joint Raman spectroscopic and quantum chemical study. Both TMAO and guanidinium are osmolytes that affect the stability of proteins, as TMAO is known to stabilize and counteract the destabilizing effects of guanidinium. While guanidinium is very similar in chemical structure to urea, the exact mechanisms of the molecular interactions between guanidinium, TMAO, and proteins continue to be investigated. Herein, we use Raman spectroscopy to elucidate the physical interactions between TMAO and guanidinium in aqueous solutions to better understand how these important osmolytes interact with each other and affect adjacent hydrogen‐bonding networks of water. Comparing experiment to theory yields good agreement and allows for the identification and tracking of different vibrational modes. It was determined that adding TMAO into an aqueous solution of guanidinium induces a blue shift (shift to higher energy) in guanidinium's H‐N‐H bending modes, which is indicative of direct interactions between the two osmolytes and similar to the earlier results observed for TMAO interacting with urea.
Article
The fundamental underpinnings of noncovalent bonds are presented, focusing on the σ-hole interactions that are closely related to the H-bond. Different means of assessing their strength and the factors that control it are discussed. The establishment of a noncovalent bond is monitored as the two subunits are brought together, allowing the electrostatic, charge redistribution, and other effects to slowly take hold. Methods are discussed that permit prediction as to which site an approaching nucleophile will be drawn, and the maximum number of bonds around a central atom in its normal or hypervalent states is assessed. The manner in which a pair of anions can be held together despite an overall Coulombic repulsion is explained. The possibility that first-row atoms can participate in such bonds is discussed, along with the introduction of a tetrel analog of the dihydrogen bond.
Article
A supramolecular ladder sustained by halogen bonds with rungs based upon a photoproduct, namely rctt -tetrakis(5′-pyrimidyl)cylcobutane, generated in the solid state is reported.
Article
The term halogen bonding describes the tendency of halogen atoms to interact with lone pair possessing atoms. The binding features and structural properties of halogen bonding are discussed and applied to drive the intermolecular self-assembly of hydrocarbons and perfluorocarbons in chemo-, site-, and enantioselective supramolecular synthesis. The halogen bonding is thus an effective and reliable tool in crystal engineering at the disposal of the supramolecular chemist.
Article
Natural abundance ¹⁵N solid-state NMR spectroscopy provides an effective method for the direct evaluation of halogen bond (XB) geometry. The change in the ¹⁵N SSNMR chemical shifts in halogen-bonded co-crystals of different dipyridyl derivatives with halobenzenes and diiodoalkanes as XB-donors is generally greater than ¹³C chemical shifts and it is shown to experimentally correlate with the normalized distance parameter of the XB. The same overall trend is confirmed by DFT calculations of the chemical shifts. More information can be found in the Full Paper by R. Gobetto, P. Metrangolo et al. (DOI: 10.1002/chem.201603392).
Article
Among recent advances towards efficient semiconducting materials, rational design guidelines have emerged focusing on the synergy between various intermolecular interactions to improve the solid-state order of π-conjugated molecules in organic electronic devices. Herein, we focus our attention on halogen bonding (XB) interactions and the crucial role of electron withdrawing substituents (e.g., nitro and fluoro) towards influencing solid-state properties via secondary interactions. Employing iodoethynyl benzene derivatives (F2BAI and (NO2)2BAI) and thiophene/furan-based building blocks equipped with pyridyl groups as self-assembling domains (PyrTF and PyrT2), co-crystals driven by XB and π-stacking interactions were formed and studied. Spectroscopic and thermal analysis of 1:1 mixtures provide initial evidence of co-crystallization. X-ray crystallography affords the inherent solid state packing motifs within each assembly. Computational studies support experimental observations, revealing the dominant interactions and contribution of each substituent group towards increasing the stability of the resulting assemblies.
Article
Solid-state nuclear magnetic resonance (SSNMR) spectroscopy is a versatile characterization technique that can provide a plethora of information complementary to single crystal X-ray diffraction (SCXRD) analysis. Herein, we present an experimental and computational investigation of the relationship between the geometry of a halogen bond (XB) and the SSNMR chemical shifts of the non-quadrupolar nuclei either directly involved in the interaction (15N) or covalently bonded to the halogen atom (13C). We have prepared two series of X-bonded co-crystals based upon two different dipyridyl modules, and several halobenzenes and diiodoalkanes, as XB-donors. SCXRD structures of three novel co-crystals between 1,2-bis(4-pyridyl)ethane, and 1,4-diiodobenzene, 1,6-diiodododecafluorohexane, and 1,8-diiodohexadecafluorooctane were obtained. For the first time, the change in the 15N SSNMR chemical shifts upon XB formation is shown to experimentally correlate with the normalized distance parameter of the XB. The same overall trend is confirmed by density functional theory (DFT) calculations of the chemical shifts. 13C NQS experiments show a positive, linear correlation between the chemical shifts and the C−I elongation, which is an indirect probe of the strength of the XB. These correlations can be of general utility to estimate the strength of the XB occurring in diverse adducts by using affordable SSNMR analysis.
Article
The similarity and differences of the three host compounds H1 = 9,9′-(biphenyl-2,2′-diyl)difluoren-9-ol, H2 = 2,2′,7,7′-tetrabromo-9,9′-(biphenyl-2,2′-diyl)difluoren-9-ol, and H3 = 2,2′,7,7′-tetra-tert-butyl-9,9′-(1,4-phenylene)difluoren-9-ol which form inclusion compounds with 3-bromopyridine and its chloro-analogue guest are compared with respect to hydrogen bonding and halogen bonding. In all cases the hydrogen bonding motif (Host)O-H···O(Host)-H···N(Guest) predominates while the halogen···halogen interactions are of secondary importance in the packing of the structures. The structural data are supported by thermal analysis, Hirshfeld surface analysis, and IR spectroscopy.
Article
The competition for binding and charge-transfer (CT) from the nitrogen containing heterocycle pyrimidine to either silver or to water in surface enhanced Raman spectroscopy (SERS) is discussed. The correlation between the shifting observed for vibrational normal modes and CT is analyzed both experimentally using Raman spectroscopy and theoretically using electronic structure theory.Discrete features in the Raman spectrum correspond to the binding of either water or silver to each of pyrimidine's nitrogen atoms with comparable frequency shifts. Natural bond orbital (NBO) calculations in each chemical environment reveals that the magnitude of charge transfer from pyrimidine to adjacent silver atoms is only about twice that for water alone.These results suggest that the choice of solvent plays a role in determining the vibrational frequencies of nitrogen containing molecules in SERS experiments.
Article
The halogen bond occurs when there is evidence of a net attractive interaction between an electrophilic region associated with a halogen atom in a molecular entity and a nucleophilic region in another, or the same, molecular entity. In this fairly extensive review, after a brief history of the interaction, we will provide the reader with a snapshot of where the research on the halogen bond is now, and, perhaps, where it is going. The specific advantages brought up by a design based on the use of the halogen bond will be demonstrated in quite different fields spanning from material sciences to biomolecular recognition and drug design.
Article
Six halogen-bonded cocrystals involving aromatic donors have been studied by far-IR spectroscopy. Characteristic redshift and intensity increase of the C-I stretching band have been observed, which provided a distinct signature of the halogen bond involving iodopentafluorobenzene.
Article
Although recognized as a significant force in crystal engineering, halogen bonding (XB) has been scarcely investigated in "bottom-up" approaches towards organic electronics. We report, herein, the use of a thiophene-based building block, pyridyl-thiophene (Pyr-T), to achieve an assembly driven by XB and π-π stacking interactions with iodopentafluorobenzene (IPFB). Spectroscopic and thermal analysis of the co-crystal provide indirect evidence of the assembly. The combined effects of XB and π-π stacking are confirmed experimentally via X-ray crystallography. Density functional theory (DFT) computations support experimental observations. The results of the study speak to the use of halogen bond driven self-assembly in organic electronic device application. This journal is