You are on page 1of 9

Article

Cite This: Cryst. Growth Des. 2018, 18, 2795−2803 pubs.acs.org/crystal

Designed Synthesis and Crystallization of Isomorphic Molecular


Gyroscopes with Cell-like Bilayer Self-Assemblies
Ma. Eugenia Ochoa,† Pablo Labra-Vázquez,‡ Norberto Farfán,‡ and Rosa Santillan*,†

Departamento de Química, Centro de Investigación y de Estudios Avanzados del IPN, 07360 CDMX, México

Facultad de Química, Departamento de Química Orgánica, Universidad Nacional Autónoma de México, 04510 CDMX, México
*
S Supporting Information

ABSTRACT: Two molecular rotors featuring pyridine and


fluorobenzene rings as polar rotators and 9-octylfluorenyl
stators were synthesized. Their crystal structures were
established through SXRD techniques, crystallizing in the
monoclinic chiral P21 space group. The supramolecular
assemblies of both isomorphs showed an orientation of static
dipoles through the crystal lattice and the formation of
intriguing 2D layers that resemble cell membranes, a typical
example of an amphidynamic system. Small activation energies
and the modulation of the first-order hyperpolarizabilities of
these compounds as a function of rotational dynamics were
revealed through DFT computations at the CAM-B3LYP/M06-2X/cc-pVDZ level of theory and correlated with a potential use
of these materials as photonic switches.

1. INTRODUCTION 2. EXPERIMENTAL SECTION


Amphidynamic systems, defined as 3D molecular arrays that, 2.1. General Experimental Considerations. Materials. All
despite their long-range phase order, also display fast molecular starting materials were purchased from Sigma-Aldrich and used
without further purification. Solvents were purified by distillation over
dynamics, have the potential to act as molecular rotors, i.e., appropriate drying agents. Compounds 2−4 were obtained following
crystalline molecular machines that, when properly engineered, reported methodologies;6−8 spectroscopic data are in good agreement.
may display efficient rotational work such as that of Instrumentation. NMR spectra were recorded on Jeol ECA 500
macroscopic gyroscopes or compasses.1 These crystal systems, and Jeol 270 MHz spectrometers using deuterated solvents; chemical
expected to display interesting properties like dichroism and shifts for 1H and 13C NMR data are relative to the residual
nondeuterated solvent signal, fixed at δ = 7.26 ppm for 1H NMR
birefringence,2 usually incorporate a 1,4-diethynylphenylene and δ = 77.00 ppm for 13C NMR. Infrared spectra were registered on
ring linked at both ends to rigid frameworks that provide the an FT-IR Varian ATR spectrometer. HRMS data were acquired using
supramolecular information that dictates the crystal packing. an Agilent G1969A MS TOF spectrometer. Elemental analysis for
Our group of work has put intense collaborative efforts in MR2 was determined on a Thermofinnigan Flash 1112 equipment
studying the rotational dynamics of molecular rotors possessing (CHONS). Absorption spectra were obtained in chloroform solutions
using a PerkinElmer Lambda 2S UV/vis spectrophotometer.
diverse rigid steroidal frameworks linked through acetylenic 2.2. Synthetic Procedures. Synthesis of Molecular Rotors. 2,5-
axles to 1,4-phenylene rotators.3−5 During the course of these Bis((9,9-dioctyl-fluoren-2-yl)ethynyl)pyridine (MR1). 2-Ethynyl-9,9-
investigations, the possibility of controlling the rotational dioctylfluorene (4) (0.20 g, 0.48 mmol), 2,5-dibromopyridine (5)
dynamics of these molecular compasses like systems using a (0.06 g, 0.25 mmol), Pd(PPh3)2Cl2 (0.02 g, 0.028 mmol), and CuI
(0.009 g, 0.03 mmol) were placed in a dried round-bottom flask,
magnetic stimulus became appealing. followed by the addition of diisopropylamine (1 mL) and freshly
Herein, we describe the synthesis and supramolecular distilled THF (25 mL) under a nitrogen atmosphere. The mixture was
structure of molecular rotors featuring 9,9-dioctylfluorene refluxed 8 h, allowed to cool to room temperature, quenched with a
groups as stators and pyridine and fluorobenzene rings as saturated solution of ammonium chloride (25 mL), extracted with
polar rotators. The static dipolar moments present in these ethyl acetate (3 × 25 mL), dried over Na2SO4, evaporated to dryness,
and chromatographed on silica gel (70−230 Mesh) using hexanes. The
molecules are intended to facilitate control of rotational brown solid thus obtained was recrystallized from deuterated
dynamics upon application of an external magnetic field. chloroform and acetonitrile to yield 80 mg (37%) of MR1 as a yellow
Through SXRD analysis of these systems, we found that they solid. M.P. 78−79 °C. FTIR (ATR cm−1): ν 3268, 2923, 2875, 2316,
present peculiar solid-state self-assemblies that resemble closely
cell membranes, as well as other supramolecular features that Received: November 3, 2017
make these crystalline isomorphs interesting systems to be Revised: April 5, 2018
studied as molecular gyroscopes. Published: April 5, 2018

© 2018 American Chemical Society 2795 DOI: 10.1021/acs.cgd.7b01542


Cryst. Growth Des. 2018, 18, 2795−2803
Crystal Growth & Design Article

Scheme 1. Synthesis and Numbering for Molecular Rotors MR1 and MR2a

a
Displayed in color code are the 9-octylfluorenyl stators (blue), the polarizable rotators (red), and the acetylenic axles (black). Reagents and
conditions: (i) 1-Bromooctane, KOH, DMSO, (ii) ethynyltrimethylsilane, Pd(PPh3)2Cl2, CuI, DIPA, reflux, (iii) K2CO3, MeOH/ether, rt., (iv)
Pd(PPh3)2Cl2, CuI, DIPA, THF, reflux.

1467, 843, 827, 737, 571, 555. 1H NMR (CDCl3, 500 MHz) δ 8.82 (d, 140.29 (C-4b), 140.25 (C-4b′), 133.18 (d, 3JC‑F = 2.2 Hz, C-21),
J = 1.9 Hz, 1H, H-24), 7.84 (dd, J = 8.1, 1.9 Hz, 1H, H-22), 7.72−7.68 130.78 (C-3′), 130.71 (C-3), 127.68 (C-7′), 127.64 (C-7), 127.30 (d,
4
(m, 4H, H-4/H-4′/H-5/H-5′), 7.62−7.58 (m, 2H, H-1/H-3), 7.56− JC‑F = 3.4 Hz, C-22), 126.90 (C-6′), 126.88 (C-6), 126.03 (C-1),
7.52 (m, 3H, H-1′/H-2′/H-3′), 7.36−7.32 (m, 6H, H-6/H-6′/H-7/H- 126.01 (C-1′), 124.88 (C-20), 123.01 (C-8), 123.01 (C-8′), 120.80
7′/H-8/H-8′), 1.97 (q, J = 7.7 Hz, 8H, H-10/H-10′), 1.23−1.17 (m, (C-2), 120.66 (C-2′), 120.07 (C-5/C-5′), 119.68 (d, 3JC‑F = 5.0 Hz, C-
8H, H-16/H-16′), 1.15−1.02 (m, 32H, H-12/H-12′/H-13/H-13′/H- 24), 118.39 (C-4), 118.30 (d, 3JC‑F = 22.6 Hz, C-23), 118.21 (C-4′),
15/H-15′), 0.82 (t, J = 7.2 Hz, 12H, H-17/H-17′), 0.61 (br s, 8H, H- 97.36 (C-18), 93.34 (C-18′), 88.10 (C-19′), 82.57 (C-19), 55.16 (C-
11/H-11′). 13C NMR (CDCl3,125 MHz) δ 152.47 (C-24), 151.18 (C- 9′), 55.14 (C-9), 40.33 (C-10/C-10′), 31.78 (C-15/C-15′), 30.00 (C-
9a′), 151.07 (C-9a), 150.88 (C9b′), 150.78 (C-9b), 142.32 (C-4a), 14/C-14′), 29.22 (C-12/C-12′/C-13/C-13′), 23.70 (C-11/C-11′),
142.13 (C-4a′), 141.92 (C-20), 140.19 (C-4b/C-4b′), 138.30 (C-22), 22.59 (C-16/C-16′) 14.07 (C-17/C-17′). Anal. Calcd for C68H85F: C,
130.73 (C-3′), 127.76 (C-7), 127.74 (C-7′), 126.92 (C-6′), 126.90 (C- 88.64; H, 9.30. Found: C, 88.36; H, 9.62.
6), 126.71 (C-1), 126.22 (C-1′), 126.02 (C-21), 122.90 (C-8/C-8′), 2.3. Single X-ray Diffraction Studies. Crystals of MR1 and MR2
120.43 (C-2′), 120.14 (C-5), 120.11 (C-5′), 120.07 (C-2), 119.74 (C- suitable for single-crystal X-ray diffraction studies were obtained by
4), 119.68 (C-4′), 119.44 (C-23), 95.58 (C-18′), 92.52 (C-18), 88.70 room temperature evaporation of a saturated solution of the analytes
(C-19), 86.05 (C-19′), 55.17 (C-9′), 55.14 (C-9), 40.31 (C-10/C- in a chloroform/acetonitrile mixture.
10′), 31.76 (C-15/C-15′), 30.00 (C-14), 29.98 (C-14′), 29.21 (C-13/ The intensity data were collected on an Enraf-Nonius Kappa
C-13′), 29.20 (C-12/C-12′), 23.72 (C-11′), 23.69 (C-11), 22.57 (C- diffractometer with a CCD area detector (λMoKα = 0.71073 Å,
16/C-16′), 14.06 (C-17/C-17′). HRMS (APCI-TOF+) m/z: [M+ + monochromator: graphite) and a Bruker D8 Venture CMOS
H]+ Observed: 904.6755, required for C67H86N: 904.6754, error: 0.23 diffractometer at 173 K. The crystals were mounted on conventional
ppm MicroLoops. All heavy atoms were found by Fourier map difference
2,5-Bis(9,9-dioctyl-2-ethynyl-fluoren)fluorobenzene (MR2). Syn- and refined anisotropically. All reflection data sets were corrected for
thesized as described above for MR1, from 2-ethynyl-9,9-dioctyl- Lorentz and polarization effects. The first structure solution was
fluorene (4) (0.15 g, 0.36 mmol) and 1,4-dibromo-2-fluorobenzene obtained using the SHELXS-2017 program, and the SHELXL-2017
(6) (0.05 g, 0.18 mmol). The procedure yielded 92 mg (56%) of MR2 was applied for refinement and output data. 9 All software
as a pale yellow solid (92 mg, 56%). M.P. 116−117 °C. FTIR (ATR, manipulations were done under the WinGX environment program
cm−1) 3063, 2925, 2853, 1467, 1451, 828, 737, 721. 1H NMR (CDCl3, set. The programs Mercury 3.7 and ORTEP-3 were used to prepare
500 MHz) δ 7.72−7.69 (m, 4H, H-4/H-4′/H-5/H-5′), 7.57−7.51 (m, artwork representations.10,11 Free volumes around the rotators were
5H, H-1/H-1′/H-3/H-3′/H-21), 7.37−7.32 (m, 8H, H-6/H-6′/H-7/ approximated to the space confined by a hypothetical prism with two
H-7′/H-8/H-8′/H-22/H-24), 1.98 (t, J = 8.3 Hz, 8H, H-10/H-10′), edges defined with the distances shown in Figure 3 (∼8 and ∼10 Å)
1.26−1.19 (m, 8H, H-16/H-16′), 1.15−1.05 (m, 32H, H-12/H-12′/ and the third edge taken as the C2−C2′ distance (∼11 Å).
H-13/H-13′/H-14/H-14′/H-15/H-15′), 0.82 (t, J = 7.2 Hz, 12H, H- 2.4. Theoretical Methods. Gas phase geometries were computed
17/H-17′), 0.65−0.59 (m, 8H, H-11/H-11′). 13C NMR (CDCl3, 125 within the framework of the Density Functional Theory using the
MHz) δ 162.09 (d, 1JC‑F = 251.9 Hz, C-25), 151.07 (C-9a), 151.05 (C- Gaussian 09W software package,12 from the crystal structures of MR1
9a′), 150.84 (C-9b′), 150.80 (C-9b), 141.96 (C-4a′), 141.95 (C-4a), and MR2. Both equilibrium geometries were obtained using the

2796 DOI: 10.1021/acs.cgd.7b01542


Cryst. Growth Des. 2018, 18, 2795−2803
Crystal Growth & Design Article

Figure 1. Crystal structures of (a) MR1 and (b) MR2 with the thermal ellipsoids drawn at 50% probability for every atom other than hydrogen.

double-ζ 6-31G(d,p) polarized basis set with different hybrid Evaluation of the modulation of μ0 and β as functions of the
functionals: B3LYP, PBE0, and the long-range corrected CAM- orientation of the pyridine and fluorobenzene rotators was carried out
B3LYP and M06-2X. In every case, a former optimization at the HF/6- by varying their relative positions from the coplanar (equilibrium)
31G(d,p) was required. These relaxed conformers were confirmed as geometries. Single-point energy calculations were then performed for
energetic local minima through analytical inspection of their conformers with successive 15° increments in the dihedral angle
vibrational frequencies and were further optimized using the integral formed between the plane containing the rotator and that of the 9-
equation formalism of the Polarizable Continuum Model (IFPCM) octyl-fluorenyl stators, i.e., the [C1−C2−C20−C21] torsion angle.
with chloroform as solvent. The resulting geometries were subjected to In order to confirm or correct the NMR assignments, a Gaussian
a TD-DFT computation of the vertical excitation energies for MR1 Invariant Atomic Orbital (GIAO) computation of the isotropic
and MR2, which were compared to the experimental electronic magnetic shieldings of 13C nuclei was performed. These calculations
spectra. From all the methods, only the CAM-B3LYP and M06-2X were conducted at the PBE0/6-31G(d,p) level of theory rather than
functionals were found to predict these energies accurately, with with the M06-2X/cc-pVDZ approach, as it was unexpectedly found
differences between the experimental and computed excitations below that the former predicted the desired 13C chemical shifts with better
the usually accepted threshold of 0.3 eV (Figure S8). These functionals accuracy and performance, as reflected in lower Mean Unsigned Errors
were consequently employed with Dunning’s cc-pVDZ basis set in the
(MUE, Table S1) obtained for both MR1 and MR2. Although the
computation of first-order hyperpolarizabilities (β), as the use of such
methodology employed provides absolute chemical shifts, for clarity,
larger atomic functions has proved effective in predicting this optical
they are herein reported as scaled with respect to an external
parameter.13 For consistency reasons, the same method was applied in
tetramethylsilane (TMS) reference computed at the B3LYP/6-
the computation of rotational activation energies (Ea) and static
molecular dipoles (μ0). It should be remarked that, throughout the 311+G(2d,p) level of theory.
entire set of computations, the CAM-B3LYP/cc-pVDZ was
unambiguously found as the best DFT method in terms of accuracy 3. RESULTS AND DISCUSSION
and computational expense. Also, the use of functionals lacking long-
range corrections behaved catastrophically, particularly at predicting 3.1. Synthesis and Molecular Characterization. Sono-
electronic parameters such as electronic excitation energies and should gashira double cross-coupling reaction between 2-ethynyl-9,9-
therefore be avoided when computing NLO parameters of molecules dioctylfluorene (4) and either 2,5-dibromopyridine (5) or 1,4-
with similar quadrupolar architectures as the rotors herein studied. dibromo-2-fluorobenzene (6) allowed access to unsymmetrical
2797 DOI: 10.1021/acs.cgd.7b01542
Cryst. Growth Des. 2018, 18, 2795−2803
Crystal Growth & Design Article

Table 1. Crystal Structure and Refinement Data


compound MR1 MR2
empirical formula C67H85N C68H85F
formula weight 904.42 921.36
temperature 173 K 173 K
crystal system monoclinic monoclinic
space group P21 P21
a (Å) 12.610(16) 12.48(5)
b (Å) 14.291(18) 14.11(4)
c (Å) 16.78 (2) 16.49(5)
α (deg) 90 90
β (deg) 104.07 (4) 103.2(2)
γ (deg) 90 90
volume (Å)3 2933(6) 2825(16)
Z 2 2
density (g·cm−3) 1.024 1.083
crystal size (mm) 0.1 × 0.1 × 0.2 0.32 × 0.069 × 0.04
θ range (deg) 2.19−28.28 2.70−26.71
index ranges −16 ≤ h ≤ 16, −19 ≤ k ≤ 19, −22 ≤ l ≤ 22 −15 ≤ h ≤ 15, −17 ≤ k ≤ 17, −20 ≤ l ≤ 20
Nref 14556 10915
R (reflections) 0.0537 (8423) 0.0575 (7766)
wR2 (reflections) 0.1385 (14374) 0.1167 (10915)

molecular rotors MR1 and MR2 in moderate yields (Scheme


1).
In the 1H NMR spectrum of MR1, two spin systems could
be readily identified through homonuclear COSY (1H−1H)
2D-NMR spectroscopy. The first coupled nuclei was ascribed
to the pyridine ring, with the characteristic resonance of H-24
appearing as a highly deshielded doublet at δ = 8.82 ppm.
Aromatic hydrogens from the fluorene submolecular fragments
appeared as multiplets, with magnetic inequivalences between
the H-1/H-1′ and H-3/H-3′ nuclei from both fluorenyl stators.
Nonetheless, their signals were successfully assigned owing to
their marked long-range scalar couplings.
Most 13C nuclei were also found to be magnetically
nonequivalent in solution, which was observed as a ubiquitous
splitting of the signals in the 13C NMR spectra of both MR1
and MR2, with chemical shift differences even below 0.05 ppm.
Ambiguity found in the assignment of these resonances was
thus circumvented employing APT and 2D (1H−1H, 1H−13C)
NMR techniques and a Gaussian Invariant Atomic Orbital
(GIAO) calculation of the isotropic 13C chemical shifts for both
molecular rotors at the PBE0/6-31G(d,p) level of theory (for
details, see the Supporting Information, Table S1 and Figures
S9 and S10).
3.2. X-ray Diffraction Studies. Crystals of compounds
MR1 and MR2 suitable for SXRD experiments were grown in
chloroform/acetonitrile at room temperature (Figure 1). Figure 2. Calculated PXRD patterns for MR1 (top) and MR2
Interestingly, despite the replacement of the pyridine nitrogen (bottom).
in MR1 for a C-F moiety in MR2, both compounds crystallized
without solvent in the monoclinic P21 space group, with 2 Remarkably, in spite of the presence of the polar pyridine
molecules per unit cell and with nearly the same cell constants and fluorobenzene rotators, the crystal packing of both MR1
(Table 1). Also, both showed an anti conformation for the two and MR2 is dominated exclusively by weak nonpolar contacts
fluorenyl stators, which lie in close orthogonality to the plane among the alkyl chains of the fluorene stators (for details, see
containing the rotator and the acetylenic axles as evidenced by the Supporting Information, Figure S10). It is worth noting that
the values for the [C1−C2−C20−C21] dihedral angles of such absence of intermolecular interactions involving the
84.6° and 86.2° for MR1 and MR2, respectively. To further rotators may facilitate their rotational dynamics.
evidence the similarities of these isomorphic crystals, a Further analysis of the crystalline structure of these
comparison between their calculated PXRD patterns is depicted compounds revealed that a given rotator has 3.8−5.2 Å of
in Figure 2. distance from the ring centroid to the closest atoms within the
2798 DOI: 10.1021/acs.cgd.7b01542
Cryst. Growth Des. 2018, 18, 2795−2803
Crystal Growth & Design Article

Figure 3. Distances (Å) from a rotator centroid to the closest atoms within the crystal packing of (a) MR1 and (b) MR2.

crystal lattice of MR1 and MR2 (see Figure 3), conferring a crystallographic b axis. We expect that, given the free space
free volume of ca. 796 and 854 Å3 around the pyridine and that the rotators display and the low rotational energetic
fluorobenzene rotators, respectively, which may be enough to barriers provided by the acetylenic axles, the orientation of
allow the rotational dynamics of these molecules.3 However, these microscopic molecular dipoles might change with the
the distances between the centroids of two adjacent rotators are presence of an applied magnetic field, yielding these crystalline
of 11.0 and 10.8 Å for MR1 and MR2, respectively, which may systems as molecular analogues of macroscopic compasses.
limit correlation of internal rotational motion. Nonetheless, It is worth noting that, because of their chiral self-assembly,
cooperative effects arising by the alignment of molecular these isomorphic crystals are also of particular interest for
dipoles throughout the crystallographic c axis (vide inf ra) of nonlinear optical (NLO) applications, where noncentrosym-
these crystal arrays may help overcome this limitation. metric arrangements are required for solid-state applications,
Both MR1 and MR2 crystal structures exhibited intriguing such as the second (or higher order) harmonic generation.
2D layers that resemble the lipidic bilayers that constitute cell Moreover, due to the alignment of molecular dipoles within the
membranes (Figure 4). This peculiar self-assembly, which crystals, these materials are expected to display enhanced NLO
spreads along the crystallographic c axis, is of relevance within properties due to cooperative effects among the static dipoles
the study of solid-state molecular rotors due to the need (of aligned throughout the crystal lattice.
that particular type of molecular machines) of having an 3.3. Theoretical Studies. The molecular structure of rotors
amphidynamic character, i.e., to display fast molecular dynamics MR1 and MR2 was studied through DFT computations using
while maintaining a highly ordered phase, such as that offered the cc-pVDZ basis set with the hybrid long-range corrected
by a crystalline array. The similarity of the crystal packing of M06-2X and CAM-B3LYP functionals. Negligible geometrical
these studied molecular rotors with the lipidic bilayers in cell differences between the gas-phase geometries were found for
membranes is thus of importance due to the fact that cell both methods. When compared to the solid-state conformers,
membranes constitute an archetypal example of an amphidy- the main geometrical difference is that the former are predicted
namic system, given that they are well-known to present both to have coplanarity among the pyridine (MR1) or fluoroben-
phase order and fast molecular dynamics.14 zene (MR2) rotators and the fluorenyl stators (Figure 6). Such
As depicted in Figure 5, both MR1 and MR2 showed a conformations may reasonably be preferred due to a better
peculiar orientation of molecular dipoles throughout the stabilization of the static molecular dipoles (μ0) in these planar
2799 DOI: 10.1021/acs.cgd.7b01542
Cryst. Growth Des. 2018, 18, 2795−2803
Crystal Growth & Design Article

Figure 4. Schematic representation of 2D layers formed within the crystal packing of compounds MR1 (top) and MR2 (bottom). The crystal
structures are viewed along the crystallographic b axis, and the rotators are in space-filling representation for clarity.

Figure 5. Orientation of static molecular dipoles along the crystallographic b axis for MR1 (left) and MR2 (right).

geometries. This was confirmed through single-point energy the loss of π-conjugation and aromaticity of these molecules as
computation at the same level of theory for both conformers, they deviate from coplanarity (Figure 7). It is worth noting
showing changes in molecular dipoles (Δμ0) of ca. 0.25 and that, even considering that the nonplanar solid-state conformers
0.05 D for MR1 and MR2, respectively, favoring the planar should have been observed due to potential C−H···π contacts
conformations. involving the rotators, these weak (∼1.5 kcal/mol) 15
In order to estimate how free the pyridine and fluorobenzene interactions should not restrain rotational motion in the solid
rotators are to perform the rotational work, the energy profiles state, as efficient dynamics have been observed with even higher
of both MR1 and MR2 were computed for conformers with energetic thresholds.5
incremental 15° deviations from the equilibrium (planar) As mentioned earlier, prospective applications of molecular
geometries, allowing us to predict very small activation energies rotors include their use as microscopic analogues of compasses,
(Ea) for both rotors (2.5−4.5 kcal/mol), mainly accounted for with conformational changes effectively induced by a magnetic
2800 DOI: 10.1021/acs.cgd.7b01542
Cryst. Growth Des. 2018, 18, 2795−2803
Crystal Growth & Design Article

Pi(E) = αijEj + βijk EjEk + ... (1)


where P(E) is the molecular polarization induced by the electric
field (E), α and β are the molecular polarizability and first-order
hyperpolarizability, respectively, and i, j, k refer to the molecular
coordinate system.
The SHG response in crystals is restricted to non-
centrosymmetric space groups, as the presence of inversion
symmetry within the molecule yields β = 0 and a cancelation of
resonant dipoles is expected in centrosymmetric supra-
molecular arrangements. In noncentrosymmetric media, the
magnitude of β greatly influences the strength of the SHG
signal. This leads to the possibility of developing photonic
switches from MR1 and MR2 where, as consequence of an
induced reorientation of the rotator, a modulation of β and thus
of the SHG response could be achieved. It should be noted
Figure 6. Superposed equilibrium (M06-2X/cc-pVDZ, orange) and that, because of the quasi-centrosymmetric quadrupolar (D-π-
solid-state (blue) geometries for MR1 (top) and MR2 (bottom). A-π-D) topology of these molecules, small (but nonzero)
values of β are expected in the coplanar conformations.
stimulus. Due to the absence of inversion symmetry in the Nonetheless, β should be modified as the rotators deviate from
crystal packing of MR1 and MR2, a pulsed laser could be used coplanarity, as this process interrupts π-conjugation and
to induce not only the conformational change but also an consequently alters the molecular topology, leading to an
associated even-order nonlinear optical (NLO) response, such “OFF” state (i.e., the conformer with the lowest β) which could
as the Second Harmonic Generation (SHG). These optical be switched to an “ON” state as molecular dipoles are
phenomena may arise from a crystal as its molecules polarize reoriented due to the application of an external magnetic field,
due to the application of an external magnetic field, following16 as successfully described by our group for analogous systems.17

Figure 7. Energy profiles for the [C1−C2−C20−C21] dihedral angle in MR1 (top) and MR2 (bottom), respectively, using the cc-pVDZ basis set
with the CAM-B3LYP (blue) and M06-2X (red) functionals.

2801 DOI: 10.1021/acs.cgd.7b01542


Cryst. Growth Des. 2018, 18, 2795−2803
Crystal Growth & Design Article

Figure 8. Modulation of the static dipole moment (μ0, dotted lines) and the first-order hyperpolarizability (β, solid lines) as functions of the [C1−
C2−C20−C21] dihedral angle for MR1 (top) and MR2 (bottom) using the cc-pVDZ basis set with the CAM-B3LYP (blue) and M06-2X (red)
functionals.

The results of a DFT computation of the modulation of μ0 computed values are not particularly appealing at the molecular
and β as functions of the orientation of the fluorobenzene and level, an enhancement in the NLO response in these crystals
pyridine rotators are depicted in Figure 8. Surprisingly, while may occur due to cooperative effects caused by the crystal
the variation of μ0 followed the same trend in both rotors packing as mentioned earlier (vide supra).
(increasing as the molecule deviates from coplanarity and thus
of centrosymmetry), opposite behaviors were found for the 4. CONCLUSIONS
modulation of β. In the case of MR1, β has a minimum value at Two molecular rotors featuring 9-octylfluorenyl stators with
a dihedral angle of 0° (i.e., in the coplanar conformer) and a pyridine or fluorobenzene rings as rotators were successfully
maximum near 90°, as expected for the change in topology obtained employing a Sonogashira double cross-coupling
previously explained. Strikingly, both DFT methods employed reaction as the key synthetic step. The crystal structure of
predicted the exact opposite trend for MR2 with a maximum these compounds formed 2D layers that highly resemble cell
value of β at coplanarity, which markedly decreases as the membranes, which suggests these crystals may also display a
dihedral angle approaches 90°. dynamic behavior within a highly ordered phase. Free volumes
The rationale behind this divergent behavior is far from of ca. 796−854 Å3 were estimated around the rotators,
trivial and out of the scope of this paper. Further studies should potentially allowing their rotational dynamics in the solid
deal with computing the involved excited states, since the value state, which may occur at room temperature according to DFT
of β, commonly assumed by a two-state model,18 depends not computations. Moreover, both their chiral crystal packing and
only on the herein studied ground state dipole (μg) but also on the modulation of the first-order hyperpolarizabilites (β) with
the change in dipole moment between the involved eigenstates the orientation of the rotators make these compounds
(Δμge) and on the energy (Ege) and oscillator strength ( fge) for particularly appealing as photonic switches, which may
the transition, following reorientate upon the application of a magnetic field, potentially
Δμge fge 2 displaying rotational work such as that of a gyroscope. In silico
β∝ evaluation of the modulation of β as a function of the rotational
Ege 2 (2) dynamics showed ON/OFF ratios of ∼2.7 and ∼1.3, further
confirming a prospective application as photonic switches. The
Despite the different behaviors of both molecular rotors, experimental evaluation of the solid state rotational dynamics of
ON/OFF ratios of 2.6−2.9 for MR1 and 1.3−1.4 for MR2 are deuterated isotopologues, as well as the study of the nonlinear
expected, according to DFT computations. While the optical behavior of these materials, are ongoing.
2802 DOI: 10.1021/acs.cgd.7b01542
Cryst. Growth Des. 2018, 18, 2795−2803
Crystal Growth & Design


Article

ASSOCIATED CONTENT poly-ynes containing derivatised fluorenes in the backbone. Dalton


Trans. 2003, 74−84.
*
S Supporting Information
(8) Jiménez-Sánchez, A.; Isunza-Manrique, I.; Ramos-Ortiz, G.;
The Supporting Information is available free of charge on the Rodríguez-Romero, J.; Farfán, N.; Santillan, R. Strong Dipolar Effects
ACS Publications website at DOI: 10.1021/acs.cgd.7b01542. on an Octupolar Luminiscent Chromophore: Implications on their
Synthesis of starting materials, 1H, 13C NMR, and HRMS Linear and Nonlinear Optical Properties. J. Phys. Chem. A 2016, 120,
spectra, elemental analyses, crystal packing, optimized 4314−4324.
geometries and Gaussian Invariant Atomic Orbital 13C (9) Sheldrick, G. A short history of SHELX. Acta Crystallogr., Sect. A:
Found. Crystallogr. 2008, 64 (1), 112−122.
chemical shifts plots (DFT), as well as crystal packing of (10) Macrae, C. F.; Edgington, P. R.; McCabe, P.; Pidcock, E.;
MR1 and MR2 (PDF) Shields, G. P.; Taylor, R.; Towler, M.; van de Streek, J. Mercury:
Accession Codes visualization and analysis of crystal structures. J. Appl. Crystallogr. 2006,
39, 453−457.
CCDC 1582960 and 1582963 contain the supplementary
(11) Farrugia, L. J. ORTEP-3 for Windows-a version of ORTEP-III
crystallographic data for this paper. These data can be obtained with graphical user interface (GUI). J. Appl. Crystallogr. 1997, 30, 565.
free of charge via www.ccdc.cam.ac.uk/data_request/cif, or by (12) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.;
emailing data_request@ccdc.cam.ac.uk, or by contacting The Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci,
Cambridge Crystallographic Data Centre, 12 Union Road, B.; Petersson, G. A.; Nakatsuji, H.; Caricato, M.; Li, X.; Hratchian, H.
Cambridge CB2 1EZ, UK; fax: +44 1223 336033. P.; Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.;

■ AUTHOR INFORMATION
Corresponding Author
Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima,
T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Throssell, K.;
Montgomery, J. A., Jr.; Peralta, J. E.; Ogliaro, F.; Bearpark, M.; Heyd, J.
J.; Brothers, E.; Kudin, K. N.; Staroverov, V. N.; Kobayashi, R.;
*E-mail: rsantill@cinvestav.mx. Normand, J.; Raghavachari, K.; Rendell, A.; Burant, J. C.; Iyengar, S. S.;
ORCID Tomasi, J.; Cossi, M.; Rega, N.; Millam, J. M.; Klene, M.; Knox, J. E.;
Pablo Labra-Vázquez: 0000-0002-8843-9709 Cross, J. B.; Bakken, V.; Adam, C.; Jaramillo, J.; Gomperts, R.;
Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.;
Notes Ochterski, J. W.; Martin, R. L.; Morokuma, K.; Zakrzewski, V. G.;
The authors declare no competing financial interest.


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.
ACKNOWLEDGMENTS Gaussian 09, Revision E.01; Gaussian, Inc.: Wallingford, CT, 2009.
The authors acknowledge support from PAPIIT (IN-216616) (13) Zouaoui-Rabah, M.; Sekkal-Rahal, M.; Djilani-Kobibi, F.;
Elhorri, A. M.; Springborg, M. Performance of Hybrid DFT Compared
and CONACYT for a doctoral scholarship for P.L.-V (337958).
to MP2Methods in Calculating Nonlinear Optical Properties of
Marco A. Leyva, Geiser Cuellar, and Teresa Cortez-Picasso Divinylpyrene Derivative Molecules. J. Phys. Chem. A 2016, 120 (44),
(CINVESTAV-IPN) are gratefully acknowledged for X-ray 8843−8852.
structure determinations, HRMS analyses, and NMR experi- (14) Gennis, R. B. Biomembranes: Molecular Structure and Function;
ments, correspondingly. We would like to express our gratitude Springer-Verlag: New York, 1989.
to Prof. Pascal G. Lacroix (CNRS) for his valuable comments (15) Tsuzuki, S.; Honda, K.; Uchimaru, T.; Mikami, M.; Tanabe, K.
on this work. The Magnitude of the CH/π Interaction between Benzene and Some


Model Hydrocarbons. J. Am. Chem. Soc. 2000, 122, 3746−3753.
REFERENCES (16) Williams, D. J. Organic Polymeric and Non-Polymeric Materials
with Large Optical Nonlinearities. Angew. Chem., Int. Ed. Engl. 1984,
(1) Khuong, T.-A. V.; Zepeda, G.; Ruiz, R.; Khan, S. I.; Garcia- 23, 690−703.
Garibay, M. A. Molecular Compasses and Gyroscopes: Engineering (17) Guerrero, T.; Santillan, R.; Garcia-Ortega, H.; Morales-Saavedra,
Molecular Crystals with Fast Internal Rotation. Cryst. Growth Des. O. G.; Farfán, N.; Lacroix, P. G. Bis(4-nitroanilines) in interactions
2004, 4, 15−18. through a [small pi]-conjugated bridge: conformational effects and
(2) Michl, J.; Sykes, E. C. H. Molecular Rotors and Motors: Recent potential molecular switches. New J. Chem. 2017, 41, 11881−11890.
Advances and Future Challenges. ACS Nano 2009, 3, 1042−1048. (18) Kulasekera, E.; Petrie, S.; Stranger, R.; Humphrey, M. G. DFT
(3) Rodríguez-Molina, B.; Farfán, N.; Romero, M.; Méndez-Stivalet, Calculation of Static First Hyperpolarizabilities and Linear Optical
J. M.; Santillan, R.; Garcia-Garibay, M. A. Anisochronous Dynamics in Properties of Metal Alkynyl Complexes. Organometallics 2014, 33,
a Crystalline Array of Steroidal Molecular Rotors: Evidence of 2434−2447.
Correlated Motion within 1D Helical Domains. J. Am. Chem. Soc.
2011, 133, 7280−7283.
(4) Rodríguez-Molina, B.; Ochoa, M. E.; Romero, M.; Khan, S. I.;
Farfán, N.; Santillan, R.; Garcia-Garibay, M. A. Conformational
Polymorphism and Isomorphism of Molecular Rotors with Fluoroar-
omatic Rotators and Mestranol Stators. Cryst. Growth Des. 2013, 13,
5107−5115.
(5) Czajkowska-Szczykowska, D.; Rodríguez-Molina, B.; Magaña-
Vergara, N. E.; Santillan, R.; Morzycki, J. W.; Garcia-Garibay, M. A.
Macrocyclic molecular rotors with bridged steroidal frameworks. J.
Org. Chem. 2012, 77, 9970−9978.
(6) Liu, L.; Wong, W.-Y.; Lam, Y.-W.; Tam, W.-Y. Exploring a series
of monoethynylfluorenes as alkynylating reagents for mercuric ion:
Synthesis, spectroscopy, photophysics and potential use in mercury
speciation. Inorg. Chim. Acta 2007, 360, 109−121.
(7) Khan, M. S.; Al-Mandhary, M. R. A.; Al-Suti, M. K.; Ahrens, B.;
Mahon, M. F.; Male, L.; Raithby, P. R.; Boothby, C. E.; Kohler, A.
Synthesis, characterisation and optical spectroscopy of diynes and

2803 DOI: 10.1021/acs.cgd.7b01542


Cryst. Growth Des. 2018, 18, 2795−2803

You might also like