You are on page 1of 26

CrystEngComm

View Article Online


PAPER View Journal | View Issue

Post-synthetic modification of supramolecular


Cite this: CrystEngComm, 2021, 23,
assemblies of β-diketonato Cu(II) complexes:
4344 comparing and contrasting the molecular
topology by crystal structure and quantum
computational studies†
Published on 03 June 2021. Downloaded on 8/11/2021 8:41:07 PM.

Mahesha, a K. J. Pampa,b C. S. Karthik, c


M. K. Hema,a
P. Malluc and N. K. Lokanath *a

The post-synthetic approach is a powerful strategy to modify organic linkers and/or metal node
functionality without disrupting the parent metal and ligand coordination. Engineering new crystal forms of
the same molecule and investigating the relationship among them is a challenging area of crystallography
to explore the structure–property relationship. In this regard, a new crystal form of homoleptic 4,4,4-
trifluoro-1-(2-naphthyl)-1,3-butanedione (TFNB) Cu(II) complex [Cu(TFNB)2(C2H6SO)] (1) is characterized by
the single crystal X-ray diffraction method which revealed that complex 1 is an isostructural solvate of
previously reported [Cu(TFNB)2(C2H6SO)2] (2). In complex 1, a single DMSO solvent molecule is apically
coordinated to the Cu(II) ion to form a distorted square pyramidal geometry with two chelating rings of
TFNB, while in complex 2, the solvent molecules are present in pairs related by the inversion center leading
to a six coordinated octahedral geometry. Besides, the newly synthesized heteroleptic Cu(II) complex of
TFNB with 8-hydroxyquinoline exists in two pseudopolymorphic forms, [Cu(TFNB)(8HQ)]2 (3) and
[Cu(TFNB)(8HQ)]·CH3COOH (4) via non-bridging solvent incorporation. Complex 3 is a phenoxide bridged
binuclear complex that adopts a distorted square pyramidal geometry. In complex 4, the bridging action of
phenoxide is broken by the effect of a guest solvent (acetic acid) molecule, leading to the formation of a
Received 3rd March 2021, discrete mononuclear complex with a square planar geometry. The solvent induced structural
Accepted 12th May 2021
transformations on the metal coordination sphere and molecular interaction topology of Cu(II) complexes,
via single crystal coordinated solvent exchange (complexes 1 and 2) and solvent molecule participation in
DOI: 10.1039/d1ce00304f
the crystal lattice (complexes 3 and 4), were comparatively analyzed and investigated by combined
rsc.li/crystengcomm crystallographic and quantum computational studies.

1. Introduction chemistry, materials science, and inorganic and bioinorganic


chemistry.2–6 Crystal engineering is a pivotal approach to
The area of supramolecular chemistry is at the forefront of design novel crystalline materials with intriguing
chemical crystallography research to unveil and understand architectures and desirable properties. The critical target of
the chemistry beyond the covalent bond.1 It is an crystal engineering is to develop new supramolecular
interdisciplinary field of science that encompasses aesthetic synthetic strategies to design a distinct supramolecular
crystal structures to explore their application in solid state architecture through different crystal forms for the same
molecule.7 In this technique, the assembly of multifunctional
a
ligands and active metal ions into the target crystalline
Department of Studies in Physics, University of Mysore, Manasagangotri, Mysuru-
570 006, Karnataka, India. E-mail: lokanath@physics.uni-mysore.ac.in
architectures is achieved via supramolecular interactions. As a
b
Department of Biotechnology, University of Mysore, Manasagangotri, Mysuru-570 result, unique metallosupramolecular systems with diverse
006, Karnataka, India one, two and three-dimensional architectures have been
c
Department of Chemistry, SJCE, JSS Science and Technology University, Mysuru- constructed.8–10 In recent years, the study of supramolecular
570 006, Karnataka, India
coordination complexes (SCCs) with metal–organic
† Electronic supplementary information (ESI) available. CCDC 2060389–2060391
contain the supplementary crystallographic data of complexes 1, 3, and 4. For
frameworks (MOFs) has been an emerging research area of
ESI and crystallographic data in CIF or other electronic format see DOI: 10.1039/ coordination chemistry.11 MOFs comprise infinite networks of
d1ce00304f inorganic clusters (metal centers) linked by organic ligands

4344 | CrystEngComm, 2021, 23, 4344–4369 This journal is © The Royal Society of Chemistry 2021
View Article Online

CrystEngComm Paper

via covalent or coordination bonds. In SCCs, the metal considered only for structural comparison with complex
centers are assembled with the active binding sites of ligating 3.33,34 The different conformers are achieved by the PSM
molecules to construct the unique supramolecular method. The supramolecular aggregation of the conformers
frameworks.12,13 Engineering different crystal forms of the in the crystalline environment is stabilized via distinct non-
same compound by choosing adequate supramolecular covalent interaction. DFT calculation is performed for the
building blocks (isomers, solvates, co-crystals, and their X-ray crystal structures of homoleptic and heteroleptic
polymorphs) is an opportunity to modify the properties of pseudopolymorphs containing different coordination spheres
crystals with applicative implication. On this account, the around the Cu(II) center. The strength of metal–ligand
design and development of polymorphs of coordination bonding and the nature of weakly coordinated solvent
complexes play a vital role in supramolecular chemistry.14–16 molecules are described via the quantum theory of atoms in
Polymorphism has been extensively studied in organic molecules (QTAIM) and the non-covalent interaction (NCI)
chemistry, and growing attention is being devoted to index model. This helps to understand the coordination
polymorphism in coordination complexes (MOFs, SCCs, and geometry around the Cu(II) ion in the pseudopolymorphs and
coordination polymers).17,18 Post-synthetic modification discerns the complexes' supramolecular self-assemblies in
(PSM) of coordination complexes is a powerful approach to the solid state. Further, the non-covalent interactions of the
Published on 03 June 2021. Downloaded on 8/11/2021 8:41:07 PM.

design multifunctional materials with precisely defined pseudopolymorphs were also analyzed using Hirshfeld
crystal structures (polymorphs) and tunable physical surface analysis, 2-D fingerprint plots, enrichment ratios, and
properties.19–21 The structural transformation induced by 3-D molecular electrostatic potential maps. Further, the
solvent molecule participation as a guest or terminal ligand energetic difference in interaction topologies of
in the crystal lattice (pseudopolymorphism), especially single pseudopolymorphic complexes is calculated, compared and
crystal coordinated solvent exchange (SCCSE) transformations visualized via energy framework analysis.
of coordination complex crystals, has been reported as a
prominent strategy in the PSM method. The analysis of SCCSE 2. Experimental section
transformation (structural construction and evolvement) is
vital to understand the origin of the changes in properties 2.1 Synthesis and crystal growth
that accurately reveal the structure–property relationship of The required chemicals and general solvents for the synthesis
the inorganic clusters.22–24 The reports on solvent induced and crystallization process were obtained from Sigma-Aldrich
structural transformations of heteroleptic metal complexes and Merck Ltd, India. The homoleptic Cu(II) complex of TFNB
involving an exchange of guest solvent molecules and SCCSE was prepared using the synthetic procedure reported earlier.30
transformations are rare. Hence, PSM should be a suitable The product yield obtained is 92%. The obtained compound
approach to calibrate the effect of given solvents on crystal was soluble in THF, DMA, DMF, and DMSO and partially
structures. An explicit understanding of the molecular soluble in MeOH and EtOH. The compound is dissolved in 3
topology and supramolecular framework modification driven mL of EtOH and heated to 80 °C, and 1 mL of DMSO is added
by a transformation in the coordination sphere and structural (3 : 1) to obtain a saturated solution. The test tube is wrapped
conformation provides the basis for designing new with paraffin foil and kept undisturbed for slow evaporation of
architectures for different practical applications.25–27 solvents to get good quality single crystals. X-ray diffraction
Supramolecular topology in polymorphs can be well analysis revealed that it is a new pseudopolymorphic form (1)
recognized and geometrically analyzed by the X-ray of previously reported complex 2.30
crystallography method. The technological advances in Synthesis of the new heteroleptic Cu(II) complex was
computational methodology are emerging as a robust strategy carried out in ethanol medium by taking an equimolar (1 : 1 :
for the advancement in crystal engineering.28,29 1) ratio of the ligands and metal salt. Initially, an ethanolic
In this article, we investigated the solvent induced solution of 8HQ (1.13 mmol) is prepared in a round bottom
structural transformations and directed supramolecular flask and an aqueous solution of CuSO4·5H2O (1.13 mmol) is
aggregation of a homoleptic complex [Cu(TFNB)2(C2H6SO)] added slowly into it. A catalytic amount of triethylamine was
(1) and compared it with its solvato-supramolecular isomers added and the mixture was stirred for 20 min. Then, the
[Cu(TFNB)2(C2H6SO)2] (2) and [Cu(TFNB)2(C2H6SO)2] (5) prepared ethanolic solution of TFNB (1.13 mmol) was added
reported previously.30 Also, pseudopolymorphic heteroleptic dropwise to the resulting mixture and heated for 30 min. The
Cu(II) complexes, [Cu(TFNB)(8HQ)]2 (3) and [Cu(TFNB) obtained precipitate was filtered, washed with water and
(8HQ)]·CH3COOH (4), obtained by the PSM method are dried. The obtained compound yield is 72%. The formation of
studied using a crystallographic and computational the compound is confirmed by EDAX and FTIR analysis. The
methodology. β-Diketones and 8-hydroxyquinoline are simple scanning electron microscopy image shows a needle-shaped
and potent organic linkers owing to their strong chelating morphology which revealed the crystalline nature of the
ability with stable coordination arrangements, flexible synthesized compound (Fig. 1a). The peaks of essential
ligating nature, and keto–enol resonance forms with active elements Cu, C, N, O and F which constitute the heteroleptic
functionality.31,32 Similar phenolate bridged β-diketone Cu(II) Cu(II) complex were observed in the EDAX spectrum (Fig. 1b).
complexes (complexes 6 and 7) reported earlier are also In the FTIR spectrum (Fig. 1c), the bands observed at 1135

This journal is © The Royal Society of Chemistry 2021 CrystEngComm, 2021, 23, 4344–4369 | 4345
View Article Online

Paper CrystEngComm
Published on 03 June 2021. Downloaded on 8/11/2021 8:41:07 PM.

Fig. 1 (a) SEM photograph, (b) EDAX spectrum and (c) FTIR spectrum of the synthesized heteroleptic Cu(II) complex.

and 1185 cm−1 are assigned to symmetric and asymmetric CF3 2.2 X-ray diffraction studies
stretching vibrations. The stretching vibrations at 1607 and X-ray quality single crystals of the complex were selected using a
1568 cm−1 are due to the υ(CO) group. The bands at 1466 polarizing microscope. Intensity data were collected on a Rigaku
and 1135 cm−1 are assigned to υ(C–O) and υ(C–C) stretching Xtalab mini single crystal X-ray diffractometer using graphite-
vibrations. The absence of a phenolic υ(O–H) stretching monochromated Mo Kα radiation (λ = 0.71073 Å, 293 K). The
vibration indicates the deprotonation of the hydroxyl complete data were processed using the d*trek program, and
hydrogen atom of the ligands to form the Cu(II) complex.33,34 absorption correction was conducted with the multi-scan
The saturated solution of the heteroleptic Cu(II) complex employed in CrystalClear-SM Expert.35 The crystal structures of
is prepared using DMA, DMF, DMSO and acetic acid the complexes were solved by direct methods (SHELXS), and the
solvents. Single crystals were grown in all the solvents by position of all non-hydrogen atoms was identified and refined
the slow evaporation method. Interestingly, single crystal on F 2 by a full-matrix least-squares procedure using anisotropic
X-ray analysis unveiled the similarity in crystal structure displacement parameters (SHELXL).36–38 All the hydrogen atoms
(phenolate bridged dinuclear form) with identical were located in difference Fourier maps and treated as riding
structural conformation in all the solvents except acetic on their parent atoms with isotropic thermal displacement
acid (mononuclear form). In an attempt to investigate the parameters [Uiso(H) = 1.2Ueq(CH, CH2) and Uiso(H) =
effect of solvent exchange, the crystalline sample of the 1.5Ueq(CH3)]. The geometrical calculations and crystal packing
dinuclear complex was immersed in acetic acid and left diagrams were prepared by the crystallographic program
undisturbed at room temperature. X-ray diffraction analysis PLATON39 and MERCURY 4.2.0 (ref. 40) software.
revealed that the structure of the dinuclear complex (3) is
transformed and it was exactly coincident with that of the
mononuclear complex (4), indicating that the solvent 2.3 Hirshfeld surface analysis
exchange induces the change from a dinuclear to The molecular Hirshfeld surface and 2-D fingerprint plot (FP)
mononuclear coordination complex. analyses for the pseudopolymorphic crystal structures were

4346 | CrystEngComm, 2021, 23, 4344–4369 This journal is © The Royal Society of Chemistry 2021
View Article Online

CrystEngComm Paper

performed using CrystalExplorer-17.5 software.41 The non- weak van der Waals interactions. 2-D scatter plots are
covalent interactions in the crystalline state were visualized generated to understand the strength of non-covalent
and quantified using a crystallographic file (cif). High interactions with the spikes that appear to be associated with
resolution molecular Hirshfeld surfaces mapped with the critical points' interaction. All the above analyses were carried
dnorm (−0.25 a.u. to 0.01 a.u.) and shape index functions out using Multiwfn version 3.7 (ref. 46) and visualized by VMD
(−0.25 a.u. to −0.01 a.u) were generated for the homoleptic software47 based on Gaussian cube files generated by Multiwfn.
and heteroleptic polymorphs. By mapping the dnorm function,
the distances of any point on the surface to the nearest 2.6 Density functional theory calculations
interior and exterior atom (di, de) are explored. The 2-D FPs
DFT calculations were carried out to analyse the electronic
were generated by combining pairs (di, de) obtained from the
structures of pseudopolymorphs using the Gaussian16
3-D Hirshfeld surface.42 The percentage contribution of
quantum chemistry software package and visualized by
various intermolecular contact pairs is obtained from the
Gaussview 6.0.16.48,49 The atomic coordinates of the crystal
FPs. Further, enrichment ratios (ERs) were determined from
structure were used as the initial guess for electronic
the crystal contact surface's decomposition values between
structure optimization. The calculations were performed
interacting chemical species pairs. The likelihood of two
using the B3LYP functional with the mixed basis set
Published on 03 June 2021. Downloaded on 8/11/2021 8:41:07 PM.

chemical species forming intermolecular contacts to stabilize


6-311G(d,p) for non-metal atoms and LanL2DZ for the Cu(II)
the crystal packing is derived and analysed.43
metal centre.50 The structural coordinates were optimized to
the ground state, and the frequency calculations reveal the
2.4 Interaction energies and energy frameworks
correspondence of electronic structure to local minima on
The molecular pairwise interaction energies of the the potential energy surface. The ground state energy, global
pseudopolymorphic crystal structures were calculated using and local indices for different conformers of Cu(II) complexes
CrystalExplorer-17.5. The hydrogen positions of the complex are calculated by Koopman's approximation. Further, natural
molecules were normalized to standard neutron diffraction charge, valence electron configuration, and bonding orbital
values for the calculation. The cluster of molecular fragments calculation were done in full NBO mode to explore the role of
with a radius of 3.8 Å with respect to the central complex stabilized hyperconjugated orbitals of the possible natural
molecule was generated using crystallographic symmetry Lewis structure.51
operations. The different interaction energy components
such as electrostatic, polarization, dispersion, exchange- 3. Results and discussion
repulsion, and total energies were calculated for the cluster.
3.1 Synthesis and crystal growth
In the result, the fragments are represented with colour
coding, which helps to identify the symmetry position and The heteroleptic Cu(II) complex was synthesized by the reflux
energy of that particular molecule with respect to the central method. The reaction is carried out with an equimolar
one. The obtained energy values are used to construct the mixture of TFNB and 8HQ ligands with aqueous copper
crystal structure supramolecular topology and visualized sulfate solution. Scheme 1 shows the resultant schematic
through energy frameworks.44 synthesis of the heteroleptic Cu(II) complex. Suitable single
crystals were grown in a saturated solution of the compound
2.5 Quantum theory of atoms in molecule analysis prepared using DMSO and acetic acid. Interestingly, they
resulted in two novel pseudopolymorphic forms that
Topological analysis has emerged as a powerful tool to explore
crystallized in the P21/n space group. The ligand TFNB has
covalent, ionic, and hydrogen bonding interactions based on
recently been used for the preparation of homoleptic Cu(II)
the molecule's electron density using Bader's (QTAIM) theory.
complexes [Cu(TFNB)2(C2H6SO)2] (2) and [Cu(TFNB)2(C3H7-
The wave function files (.wfn) for the QTAIM and NCI index
NO)] (5).30 The crystal structure of complex 1 presented here
topological analyses were generated from the DFT calculations
is an isostructure of complex 2 reported previously. Further,
using Gaussian software. The interactions are specified by the
it also vindicates our heuristics that the mixed-ligand system
bond path, a single maximum line of electron density that
is an essential precondition for exploring bio-active ligands'
connects the nuclei of two atoms in an equilibrium
behavior in the coordination complex. Taking a synthetic cue
configuration with the stabilized potential energy. The nature
from our earlier work on β-diketonato ligands, the task of
of interactions in the molecule is assessed using the values of
generating elusive phenoxide bridged dimers can be
electron density ρ(r), Laplacian ∇ 2ρ(r), potential V(r), kinetic
envisaged and achieved by transforming the homoleptic
energy density G(r) and total energy densities H(r) at the bond
complex to a heteroleptic complex.
critical points (BCPs). The interaction energies for different
molecular interactions were also estimated through Eint = −0.5
V(r) × 2625.5 (kcal mol−1).45 The negative region of sign(λ2)ρ 3.2 Single crystal structure analysis
with a larger value is indicative of attractive interactions (blue); Crystal structure of [Cu(TFNB)2(C2H6SO)] (1), [Cu(TFNB)
meanwhile, if sign(λ2)ρ is large and positive it indicates steric (8HQ)]2 (3), and [Cu(TFNB)(8HQ)]·CH3COOH (4) complexes.
interaction (red). The green colored spikes near zero indicate The molecular structures of the homoleptic (1) and

This journal is © The Royal Society of Chemistry 2021 CrystEngComm, 2021, 23, 4344–4369 | 4347
View Article Online

Paper CrystEngComm
Published on 03 June 2021. Downloaded on 8/11/2021 8:41:07 PM.

Scheme 1 The pathway for the formation of heteroleptic Cu(II) complex pseudopolymorphs by the PSM method.

heteroleptic complexes (3 and 4) were determined by the introduced the utility of a simple metric to assess the
single-crystal X-ray diffraction method, and the ORTEP view molecular shape by the τ5 parameter for five-coordinated
of the complexes with a selective atomic numbering scheme transition metal complexes. τ5 = ( β − α)/60, where β and α are
is presented in Fig. 2. The crystal structure data and the two largest angles around the central atom; τ5 is 0 for a
refinement parameters of the homoleptic and heteroleptic regular square pyramidal geometry and 1 for a regular
complexes are summarized in Table S1.† The selected bond trigonal bipyramidal geometry.52 The geometry index for
lengths, bond angles, and torsion angles of the crystal four-coordinated complexes is represented by τ4. τ4 = [360° −
structures are given in the ESI† (Tables S3–S5) and compared (α + β)]/141°, τ4 = 0 for a perfect square planar geometry and
with the optimized electronic structure. 1 for a perfect tetrahedral geometry.53 The coordination
Crystal structure analysis revealed that homoleptic geometry around Cu(II) is described as a slightly distorted
complex 1 is a neutral uninuclear Cu(II) complex that square pyramidal with an Addison parameter (τ5) of 0.0586.
crystallizes in the triclinic space group P1̄ with two symmetry Meanwhile, in complex 2, the expansion of the asymmetric
independent TFNB molecules and one crystallographically unit unveils that the Cu(II) ion is in a distorted octahedral
unique DMSO molecule coordinated to the Cu(II) ion in the coordination environment. Two TFNB ligand molecules are
asymmetric unit. Complex 1 is the isostructural form of coordinated to the central metal ion in equatorial positions
complex 2.49 Both complexes are solvates including DMSO as and two DMSO solvent molecules in axial positions (Fig.
the guest coordinating solvent. In complex 1, a single DMSO S1b†). Interestingly, the oxygen donors in complex 1 are
molecule is apically coordinated to the Cu(II) ion, whereas in inclined in a cis configuration, whereas complex 2 has a trans
complex 2, the guest molecules are present in pairs (related configuration. The coordination of the TFNB ligand to the
to each other by the centre of inversion). However, a Cu(II) ion generates two six-membered chelate rings. The
significant structure conformational difference and variation non-coplanar nature of chelate rings in complex 1 is
in supramolecular architecture is observed in both indicated by the dihedral angle of 12.83(13)°, whereas in
homoleptic complexes. The complex 1 cation features a five- complex 2, perfect planarity is observed since they are related
coordinate Cu(II) centre with doubly chelated –O donors of by the centre of inversion. Two distinct naphthalene rings in
the TFNB ligand in equatorial positions, and the apically complex 1 are deviated from their respective chelating ring
coordinated oxygen atom of DMSO supplementing the fifth by 12.41(13)° and 17.81(12)°, while in complex 2, the
coordination site (Fig. S1a†). Addison and co-workers chelating diketone ring is twisted by 27.51(11)°. In complex

4348 | CrystEngComm, 2021, 23, 4344–4369 This journal is © The Royal Society of Chemistry 2021
View Article Online

CrystEngComm Paper

complexes is composed of deprotonated phenoxide oxygen


(O1), pyridine nitrogen (N1) of 8HQ, and keto and enolic
oxygen atoms (O2 and O3) of TFNB. The four donor atoms in
the basal plane form the five and six membered chelating rings
with the central copper metal core. Significantly, the bridging
mode of μ2 phenolate oxygen (Oi) occupies the axial position in
complex 3 to complete the stable five coordination geometry.
The Cu(II) cationic centre geometry in complex 3 is a distorted
square pyramid with a τ5 of 0.28 [Fig. S1(c)†]. The nearly perfect
square planar coordination environment of the Cu(II) centre in
complex 4 is unveiled by the τ4 index of 0.061 [Fig. S1(d)†]. In
complex 3, the coordinating atoms O1, O2, O3, N1, Cu1, and
O3i are deviated from the base plane of a square pyramid (O1,
N1, O2, and O3) by the distance −0.141(6), 0.147(4), −0.158(4),
0.151(5), −0.1743(7) and −2.516(4) Å, indicating a slightly more
Published on 03 June 2021. Downloaded on 8/11/2021 8:41:07 PM.

distorted geometry due to the bridging action of phenolate


oxygen. In complex 4, the deviation of chelating atoms from
the basal plane is in the range of 0.016–0.021 Å showing the
planarity of the coordination polyhedra. The dihedral angle
between five and six membered chelating rings in complex 3
and 4 is 23.8(3)° and 1.62(10)°, indicating higher distortion in
the dinuclear unit. The naphthalene ring and 8HQ rings are
deviated from their respective chelating ring by 26.5(3)° and
2.7(2)° in complex 3, whereas 5.74(10)° and 1.48(10)° deviations
are observed in complex 4. Two types of coordination modes
are displayed by the 8HQ ligand: phenolate oxygen connects
the metal ion via a μ2 bridging mode in complex 3, while a
monodentate coordination mode is observed in complex 4. In
complex 3, the μ2 bridging action of the phenolate oxygen
atom (axial and equatorial) coordinates Cu1 and its symmetric-
related Cu1i through the facing O1i and O1 atoms. The double-
bridged phenoxide oxygen forms two inverted symmetry
dependent square pyramidal geometries with the copper centre
Cu1i–O3 = 2.389(6) Å. The bond angle of 93.90(17)° is observed
at the bridgehead phenolate oxygen [Cu1–O3–Cu1i]. The
Fig. 2 ORTEP of homoleptic complex 1 and heteroleptic complexes 3 equatorial planes formed by the chelating action around the
and 4 with the selective atomic numbering scheme.
symmetry related Cu1 and Cu1i atoms are perfectly parallel
and separated by 2.606 Å. The Cu2O2 formed a slightly
distorted quadrilateral geometry from the ideal rhombus in
1, the coordinating atoms O1, O2, O3, O4, and Cu1 are which the centre of inversion of the dimeric unit resides at the
deviated from the base plane of a square pyramid (O1, O2, intersection of two diagonals. The torsion angle of the
O3, and O4) by the distance 0.031(3), −0.032(3), −0.031(4), rhombus core (Cu1–O1–Cu1i–O1i = 0°) substantiates the
0.032(3) and 0.1503(4) Å, indicating a slightly distorted syn-periplanar configuration of the bridged ring. The dinuclear
geometry due to single solvent coordination. unit of 3 is termed as anti-coplanar, as the two terminal Cu–N
X-ray crystallographic investigation revealed that bonds are oriented in an anti-arrangement for the Cu2O2 plane.
heteroleptic Cu(II) complexes 3 and 4 exist in Structural comparison of homoleptic and heteroleptic
pseudopolymorphic forms. Fig. 2 shows the ORTEP of complexes. The structural parameters correlating the
complexes 3 and 4 with the selective atomic numbering geometry of the β-diketone homoleptic Cu(II) complexes (1, 2,
scheme. Complex 3 is a doubly phenoxide bridged and 5)30 and heteroleptic complexes (3, 6 and 7)33,34 are
centrosymmetric dinuclear complex containing a Cu2O2 procured from CSD and listed in Table 1 for comparative
rhombus core, whereas complex 4 is a mononuclear species analysis. The crystal structures of the homoleptic and
with the solvent acetic acid in the lattice. Both complexes heteroleptic complexes for structural comparison are
crystallize in the monoclinic crystal system with an identical displayed in Fig. S2 and S3† respectively.
space group. The Cu(II) centre adopts a five coordinated Interestingly, the homoleptic Cu(II) complex of the
polyhedral geometry in complex 3, while in complex 4, a four chelating β-diketonate (TFNB) ligand exists in three
coordinated geometry is observed. The basal plane in both pseudopolymorphic forms (1, 2 and 5) (Fig. S2†). The Cu(II)

This journal is © The Royal Society of Chemistry 2021 CrystEngComm, 2021, 23, 4344–4369 | 4349
View Article Online

Paper CrystEngComm

Table 1 Structural parameters correlating the geometry of related mononuclear β-diketone complexes (1, 2 and 5) and phenoxide bridged binuclear
heteroleptic Cu(II) complexes (3, 6 and 7)

Bond lengths (Å)


Complex CN In-plane Axial τ4/τ5 ρ (Å) Ref.
[Cu(TFNB)2(C2H6SO)] 5 Cu1–O1 = 1.941(3) Cu1–O5 = 2.235(3) 0.058 0.150 Complex 1
Cu1–O2 = 1.945(3)
Cu1–O3 = 1.942(3)
Cu1–O4 = 1.939(3)
[Cu(TFNB)(C2H6SO)]2 6 Cu1–O1 = 1.948() Cu1–O3 = 2.452() — 0.00 Complex 2
Cu1–O2 = 1.943()
[Cu(TFNB)(C3H7NO)]2 5 Cu1–O1 = 1.944() Cu1–Oi = 2.286(6) 0.127 0.177 Complex 5
Cu1–O2 = 1.929()
Cu1–O2 = 1.952()
Cu1–O3 = 1.938()
[Cu(TFNB)(8HQ)]2 5 Cu1–O1 = 1.905(6) Cu1–Oi = 2.381(6) 0.280 0.139 Complex 3
Cu1–N1 = 1.979(6)
Cu1–O2 = 1.941(5)
Published on 03 June 2021. Downloaded on 8/11/2021 8:41:07 PM.

Cu1–O3 = 1.913(6)
[Cu(TFPB)(8HQ)]2 5 Cu1–O1 = 1.945(4) Cu1–Oi = 2.387(3) 0.281 0.130 Complex 6
Cu1–N1 = 1.978(4)
Cu1–O2 = 1.955(3)
Cu1–O3 = 1.930(4)
[Cu(TFTB)(8HQ)]2 5 Cu1–O1 = 1.948(5) Cu1–Oi = 2.398(6) 0.280 0.129 Complex 7
Cu1–N1 = 1.995(6)
Cu1–O2 = 1.955(5)
Cu1–O3 = 1.943(6)

cationic centre of complexes 1 and 5 adopts a five-coordinate by linking the Cu(II) metal centres through bridged
distorted square-pyramidal geometry with trigonality indices phenoxide oxygen which leads to short Cu⋯Cu
(τ5) of 0.281 and 0.127, whereas complex 2 has a six intramolecular separations (3.133, 3.198 and 3.212 Å). In 3, 6
coordinated distorted octahedral geometry. The central Cu(II) and 7, the bridged four-membered rhombus core Cu2O2 units
ion in complexes 1 and 5 is shifted ( ρ) by 0.120 and 0.142 Å are constrained to be planar by the presence of the symmetry
from the basal plane towards the axially coordinated DMSO element with a Cu–O–Cui bond angle of 93.26(17)°,
and DMF solvent molecule, respectively, whereas in complex 94.64(13)° and 94.75(17)°, respectively. In the three
2, the metal ion is on the exact basal plane with zero shift. μ-phenoxide bridged dinuclear Cu(II) complexes listed in
The superposition of homoleptic pseudopolymorphs (Fig. 3a) Table 1, the τ5 value varies from 0.02 to 0.11, indicating that
shows the coordination sphere's structural transformation in these structures are closer to the ideal square pyramidal
a cis and trans configuration. This reveals that the complex configuration. The Cu(II) ions are displaced by a distance (ρ)
with six coordination exists in a trans configuration, whereas of 0.139, 0.130 and 0.129 Å from the basal plane (O1, N1, O2,
the complexes with five coordination exist in a cis O3 and Cu) toward the apical Oi atom. The observed
configuration irrespective of the guest coordinating solvent. dimerization of the Cu(II) complex in 3, 6 and 7 might be due
The asymmetric bridging mode of phenoxide oxygen of to the absence of a strong donor in the equatorial location,
the 8HQ ligand has also been established in heteroleptic which leads to the default coordination of phenoxide oxygen
complexes 6 and 7 (Fig. S3†). In complexes 3, 6 and 7, the present in the neighbouring complex. Puckering analysis
formation of dimeric units [Cu(TFNB)(8HQ)]2, [Cu(TFPB) unveiled that the metallocycle Cg(5) chelating ring formed by
(8HQ)]2 and [Cu(TFTB)(8HQ)]2 respectively is accomplished the 8HQ ligand is puckered in dinuclear complexes 3, 6 and

Fig. 3 Molecular overlay of (a) homoleptic (β-diketone) Cu(II) complexes and (b) heteroleptic phenoxide bridged dinuclear complexes.

4350 | CrystEngComm, 2021, 23, 4344–4369 This journal is © The Royal Society of Chemistry 2021
View Article Online

CrystEngComm Paper

7 with a puckering amplitude Q(2) = 0.088(5), 0.102(3) and projection of all the complexes viewed along the
0.106(4) Å, respectively. The ring conformation on the crystallographic a-axis is shown in Fig. S5.†
puckering surface ϕ(2) = 3(5)°, 1(2)° and 6(3)° leads to the In homoleptic complexes 1 and 2, C–H⋯O, C–H⋯F and
formation of an envelope conformation on the Cu1 centre di-hydrogen (H⋯H) molecular contacts form two S(5) and
(Fig. S4†). Further, this envelope configuration of the one S(6) loop; besides, one supramolecular ring with the
chelating ring is substantiated from the pseudorotation graph-set notation of S(5) formed by H⋯H interactions is
parameters P = 167.7(39)°, 165.4(24)°, 172.4(24)° and τ(M) = observed only in complex 2 (Fig. 4a and b). The presence of
7.7(4)°, 8.8(2)° and 9.2(3)° for the reference Cu1–O1 bond.54 coordinated DMSO is a major participant in the various non-
The overlay of heteroleptic complexes 3, 6 and 7 is displayed covalent interactions to stabilize the crystal packing. In
in Fig. 3b which emphasizes the similar structural complex 1, the central symmetry independent molecule
conformation for the phenoxide bridged dinuclear complexes forms three inversion-related dimers with ring motifs of
6 and 7. Interestingly, complex 3 has a similar coordination R22(22), R22(16) and R22(8) by C7–H7⋯S1, C20–H20⋯F6 and
bridging mode but stabilized with a completely different C29–H29C⋯O5 hydrogen bond inter-contacts (Fig. 5). The
structural conformation. R22(22) and R22(8) dimers propagate alternatively along the
Supramolecular architecture description of Cu(II) complex b-axis to generate an infinite one-dimensional zig-zag chain.
Published on 03 June 2021. Downloaded on 8/11/2021 8:41:07 PM.

crystal structures. The supramolecular assembly of Cu(II) The R22(16) loop interconnects the 1-D chain which leads to
complexes in the crystalline phase is constructed and the formation of a 2-D planar sheet in the bc plane. Further,
stabilized through various hydrogen bonding and π–π C12–H12⋯F5, C29–H29A⋯F3, C29–H29A⋯O1 and C29–
interactions. The conformational difference in the H29C⋯O3 C29–H29C⋯O5 and C20–H20B⋯S1 short contacts
homoleptic (complex 1 and 2) and heteroleptic also significantly contribute to the packing of molecules to
pseudopolymorphs (3 and 4) establishes the unique and produce the 3-D supramolecular architecture in complex 1
aesthetic supramolecular assemblies through different kinds (Fig. 5d), whereas in complex 2, two inversion dimers are
of non-covalent interactions. The geometries of various observed with R22(20) and R22(8) homosynthons formed by
hydrogen bonds in homoleptic and heteroleptic Cu(II) C12–H12⋯O3 and C16B–H16B⋯O3 contacts (Fig. 6). The
complexes are listed in Table S2.† The crystal packing R22(8) inversion dimers propagate along the crystallographic

Fig. 4 Supramolecular rings formed by intramolecular hydrogen bonds in the homoleptic [(a) complexes 1 and (b) 2] and heteroleptic [(c)
complexes 3 and (d) 4] Cu(II) complexes.

This journal is © The Royal Society of Chemistry 2021 CrystEngComm, 2021, 23, 4344–4369 | 4351
View Article Online

Paper CrystEngComm
Published on 03 June 2021. Downloaded on 8/11/2021 8:41:07 PM.

Fig. 5 Three discrete dimers, (a) R22(22), (b) R22(16) and (c) R22(8) formed through intermolecular hydrogen bonds. (d) 3D packing of complex 1
viewed along the a-axis.

Fig. 6 Two distinct dimers, (a) R22(20) and (b) R22(8) formed via intermolecular hydrogen bonds. (c) 3D packing of complex 2 viewed along the b-axis.

4352 | CrystEngComm, 2021, 23, 4344–4369 This journal is © The Royal Society of Chemistry 2021
View Article Online

CrystEngComm Paper

a-axis, which are interconnected by C12–H12⋯O3 contacts C14–H14⋯π (aryl) interaction. Another carbon atom (C16) of
with a R22(20) ring motif to form the 2-D planar sheet. These the pyridine ring of the molecule at (x, y, z) is in contact with
hydrogen bonds are assisted by various C–F⋯π and π⋯π the molecule at (1/2 − x, 1/2 − y, z and 1/2 − x, 1/2 − y, z)
interactions to stabilize the overall crystal packing (Fig. 6c). through the C14 atom of the phenyl ring to form inter dimer
In heteroleptic complexes 3 and 4, the intramolecular π⋯π stacking interaction (Fig. 7b). In complex 4, the solvent
hydrogen bond contacts of the type O–H⋯O, C–H⋯O and C– acetic acid molecule in the lattice plays a vital role in the
H⋯F form a metal chelate supramolecular S(5) ring, two aryl crystal packing with O–H⋯O and C–H⋯O intermolecular
S(5) ring and one S(6) ring motif (Fig. 4c and d). In complex hydrogen bond interactions with adjacent molecules
3, the short contact interaction C20–H20⋯F1 with four (Fig. 8a). The acetic acid's carboxylate oxygen acts as a donor
adjacent molecules establishes the interlocked wave-like to form the O4–H4⋯O1 contact with phenoxide oxygen, and
architecture in the bc plane. With a unique position and naphthyl carbon acts as a donor to form C14–H14⋯O5
directionality, the –CF3 moiety exhibits an electrostatic interaction with the O5 acceptor which leads to the 2-D
F3⋯F3 interaction (θ1 = θ2) which leads to the formation of packing formed by interconnection of adjacent complex
an infinite 1-D chain along the crystallographic a-axis. molecules. Besides, the π⋯π (C6–C22) contacts, C6–C15 and
Besides, this halogen bond interaction is assisted by C–H⋯π C7–C16 stacking interactions increase the intercomplex
Published on 03 June 2021. Downloaded on 8/11/2021 8:41:07 PM.

(aryl) and π⋯π interactions to increase the crystal packing stability. The solvent–complex interactions lead to the
stability (Fig. 7a). The phenyl ring carbon atom (C14) on formation of a 3-D array of molecular channels along the
either side of the dimeric unit acts as a donor to the ring crystallographic b-axis where the bridging solvent molecules
centroid of the six membered pyridine ring (Cg4) to form occupy the channels alternatively (Fig. 8c).55,56

Fig. 7 (a) Partial structural motif of endless chains formed via F⋯F and π–π (C14–C16) interactions leading to the formation of the supramolecular
assembly. (b) Packing of complex 3 viewed along the b-axis.

This journal is © The Royal Society of Chemistry 2021 CrystEngComm, 2021, 23, 4344–4369 | 4353
View Article Online

Paper CrystEngComm
Published on 03 June 2021. Downloaded on 8/11/2021 8:41:07 PM.

Fig. 8 (a) Solvent bridged 1D chain formed through O–H⋯O and C–H⋯O contacts. (b) Two dimensional assembly fused by C–H⋯F and strong
π–π interaction. (c) 3D packing of complex 4 viewed along the crystallographic a-axis.

3.3 Hirshfeld surface analysis solvatomorphs and pseudopolymorphic crystals with variable
solvents or counterions.57–59
Diverse non-covalent interactions have stabilized the 3-D molecular HS and their associated 2-D FPs for
supramolecular architectures of the pseudopolymorphic homoleptic and heteroleptic pseudopolymorphic Cu(II)
Cu(II) complexes. The non-covalent interactions in the lattice complexes are generated. The contact points are highlighted
space are visualized through a 3-dimensional electron density in the HS of isostructures (complex 1 and 2), which display
map. Hirshfeld surface (HS) and the associated 2-D FPs the significant difference in the overall intermolecular
together help to quantify the molecular contacts interactions and similarity in the inter-contacts generated by
conveniently. It is an effective tool to analyze the similarities the coordinated solvent molecule (Fig. 9). However, the
and differences in the self-assembly of molecules to construct discrepancy in non-covalent interactions' strength can be
the unique structural topology in isostructures, polymorphs, gauged from the altered area of the red coloured region at

Fig. 9 View of the dnorm mapped Hirshfeld surface and shape index surface for homoleptic Cu(II) complexes 1 (a) and 2 (b).

4354 | CrystEngComm, 2021, 23, 4344–4369 This journal is © The Royal Society of Chemistry 2021
View Article Online

CrystEngComm Paper

the respective contact points. The two similar circular red dnorm and shape index property with a transparent view to
spots in both complexes 1 and 2 near the coordinating allow the visualization. The spots on the HS recapitulate the
solvent are related to the R22(8) inversion dimer formed by C– information of non-covalent interaction effectively, which
H⋯O interaction (Fig. 9a). Interestingly in complex 1, reveals that complexes 3 and 4 are stabilized via different
halogen (F) and chalcogen atoms are (O and S) involved in packing modes with distinctive supramolecular self-assembly.
major non-covalent interactions to establish the 3-D The red spots on the Hirshfeld surface of complex 3 indicate
supramolecular assembly. The two red spots near the –CF3 the contacts corresponding to centrosymmetric dimers due
moiety indicate the significant C6–H6⋯F2 and C20–H20⋯F6 to F⋯F interactions; however, adjacent dimers are connected
bonding interactions between inversion dimers with a R22(16) by very weak C–H⋯F interactions. However, the red dip
ring motif. Further, one faint red spot near DMSO indicates pattern on the shape index mapping near the –CF3 moiety
the C–H⋯S contact between inversion dimers with a R22(22) (Fig. 10a) represents the region of F⋯F contacts. In complex
homo-synthon that propagate along the crystallographic 3, centrosymmetric dimers face one another through the
b-axis alternatively with R22(8) dimers. Finally, a weak tetrel fluorine atoms which are linked via dihalogen F⋯F
bond is also observed in complex 1, wherein two chelating intermolecular contacts. The F⋯F distance of 2.751 Å is
oxygen atoms approach the carbon atoms of the naphthyl notably shorter than the sum of van der Waals radii of
Published on 03 June 2021. Downloaded on 8/11/2021 8:41:07 PM.

ring via face-to-face O⋯C interactions and contribute to the fluorine atoms (2.94 Å). This shows the halogen interaction's
crystal packing's stability (Fig. 9b). In both isostructures, key relevance (F⋯F and C–H⋯F) in driving the
various C–F⋯π contacts are established by the supramolecular self-assembly of complex 3. The red hollow
electronegative fluorine atoms and the electropositive Cg region near the 8HQ pyridine (C16) and naphthyl ring (C14)
centre of the naphthyl ring. The shape index property of indicates the π–π (C14⋯C26) interaction. The bow-tie
pseudopolymorphic complexes 1 and 2 is shown by the blue- patterns formed by red and blue triangles on the surface
coloured bulging area around the –CF3 moiety and the show that the motifs are related to one another via π–π
orange coloured hollow area about the centroid of the stacking interactions (Fig. 10a).
naphthyl ring (complex 1: Cg3, Cg4, and Cg5; complex 2: Cg3 The presence of lattice acetic acid in the motif of
and Cg4) is attributed to C–F⋯π interactions. The π–π heteroleptic complex 4 alters the structural conformation and
stacking interactions in the complexes are represented by the supramolecular assembly compared to solvatomorphic
red-blue coloured bow-tie patterns (adjacent triangles) on the isomer 3. The circular red spot on the surface is attributed to
shape-index surface. The observed bow-tie patterns (red and the C–H⋯O hydrogen bond with a distance of 2.46 Å
blue) in the same region are characteristic of molecular (Fig. 10b). Four faint red spots of lower intensity, two on the
stacking interactions; the blue triangle denotes the convex pyridine and naphthyl ring, indicate the C6⋯C15 and
zone owing to the presence of ring carbon atoms of the C7⋯C16 contacts. The bifurcated interaction nature of the
complex inside the surface, while the red triangles signify the C6 atom exhibits two distinct π–π (C15 and C22) interactions
concave zone originating from carbon atoms of the π-stacked to construct the 2-D stacked network. The weak tetrel bond
molecule above it. (O2⋯C4) formed by chelating oxygen with the carbon atom
Fig. 10 illustrates the 3-D Hirshfeld surface of heteroleptic of the chelating ring via face-to-face interactions also plays a
pseudopolymorphs 3 and 4 that have been mapped over the vital role in stabilizing the molecular packing. The lone

Fig. 10 View of the dnorm mapped Hirshfeld surface and shape index surface for heteroleptic Cu(II) complexes 3 (a) and 4 (b).

This journal is © The Royal Society of Chemistry 2021 CrystEngComm, 2021, 23, 4344–4369 | 4355
View Article Online

Paper CrystEngComm

pair–π (Cg6) interaction formed by the chalcogen atom (O5) interaction. C–H⋯F interactions make H⋯F intermolecular
is represented by an orange coloured hollow region about the contacts (de + di ≈ 2.6 Å) a significant contributor to the
phenyl ring. The three bow-tie patterns formed by adjacent Hirshfeld surface, with 20.9% and 19.9% for complexes 3
red and blue triangles near the phenyl (Cg6), pyridine (Cg3) and 4, respectively. The H⋯O contacts (12.5%) play an
and naphthyl ring (Cg4) on the shape index surface are essential role in the packing modes of complex 4 which are
attributed to π–π stacking interaction in complex 4. shown as a pair of relatively sharp spikes in the vicinity of de
Fingerprint and enrichment ratio analysis. The similarities + di ≈ 2.3 Å corresponding to C–H⋯O and O–H⋯O
and discrepancies in various intermolecular contacts in interactions induced by the lattice solvent. The
different molecular environments of the pseudopolymorphic halogen⋯halogen contacts in complex 3 are shown in the FP
Cu(II) complexes are quantified by FPs (Fig. 11) as generated (de + di ≈ 2.8 Å) corresponding to the F⋯F interaction (3.2%)
from the HS. A pie chart represents the resulting percentage responsible for the formation of a 1-D linear chain. The C⋯C
contribution of various interactions. contact's contribution is shown as areas on the diagonal at de
In homoleptic Cu(II) complexes 1 and 2, the ≈ di ≈ 1.7–2.2 Å with 2.7% and 8.4% attributed to the π⋯π
intermolecular H⋯F, H⋯H and H⋯C interactions are the stacking interactions. The difference in the nature of π⋯π
dominant contributors to the overall molecular crystal stacking in 3 and 4 is clearly reflected in the presence of a
Published on 03 June 2021. Downloaded on 8/11/2021 8:41:07 PM.

packing. In complex 1, H⋯F > H⋯C > H⋯H comprising higher proportion of C⋯C contacts in complex 4 compared
30.8%, 24.2% and 21% of the total number of contacts to 3. Further, the complex's molecular surface is also
(Fig. 11a), whereas in complex 2, H⋯H > H⋯F > H⋯C with populated by C⋯O (1.9% and 3.5%) and C⋯F (2.4% and
a contribution of 38.7%, 23.3% and 11.2%, respectively 1.3%) contacts.
(Fig. 11b). The characteristic spike at de + di ≈ 2.2–2.3 Å To understand the propensity for privileged and disfavoured
resembles the shortest H⋯H contact and a subtle feature of contacts between a pair of chemical species in a crystal
the FP shows the distinct splitting of the short H⋯H structure, the enrichment ratios (ER) were determined for all
fingerprint in complex 1. FP splitting is observed when the the Cu(II) complexes. The calculated ER values for polymorphic
shortest contact has occurred between three atoms, rather complexes 1 and 2 (Table 2) reveal that H⋯H contacts are more
than a direct two atom contact. The crystal structure of favoured in complex 2 (EHH = 0.99) compared to complex 1
complex 2 is dominated by H⋯H contacts, whereas the H⋯F (EHH = 0.73) since they dominate (38.7%) the interaction
contact is the most dominant in complex 1. The H⋯F surface in complex 2. The EHO, EHF and EHS values are larger
contacts are shown as a pair of spikes at de + di ≈ 2.4–2.5 Å; than unity in both complexes, indicating that H⋯O, H⋯F and
these contacts in complex 1 correspond to C–H⋯F hydrogen H⋯S interactions are the most privileged contacts for
bonds responsible for the 2-D packing with the formation of constructing stable supramolecular architecture. The vice versa
R22(16) centrosymmetric dimers. The H⋯O contacts are trend is observed for H⋯C and C⋯C inter-contacts, where in
shown as two spikes at de + di ≈ 2.5 Å for complex 1, while in complex 1, H⋯C contacts show an increased propensity with
complex 2 a pair of relatively sharp spikes are observed with EHC = 1.12 and C⋯C inter-contacts are less favoured. In
the shortest de + di ≈ 2.3 Å corresponding to C–H⋯O complex 2, the C⋯C contacts are remarkably enriched (ECC =
interactions. The C⋯C contact's contribution is shown as 2.63) owing to the relatively high value of their proportion on
areas on the diagonal at de ≈ di ≈ 1.6–2.3 Å with 2.2% and the molecular surface. Remarkably, the favoured chalcogen
6% for complexes 1 and 2. This is attributed to the above- C⋯O contacts are observed only in complex 1 with ECO = 1.11
mentioned π⋯π stacking interactions in the crystal as discussed in HSA.
structures. The chalcogen C⋯O contacts are observed only in The features of all the inter-contacts and their ER values
complex 1, and are absent in complex 2. Further, the for heteroleptic solvatomorphic complexes 3 and 4 are
molecular surface is also populated by H⋯C, H⋯S, C⋯F and tabulated (Table 2). The H⋯H contacts are favoured in both
F⋯F contacts in both complexes. complexes in which EHH is close to unity since it comprises
The relative contribution of different interactions in the majority of the interaction surface (32.3 and 31.5%). The
heteroleptic complexes 3 and 4 is presented in 2D histograms structures of 3 and 4 are characterized by favoured H⋯O
(Fig. 11c and d). The differences in the crystal environment contacts with an ER greater than unity. Remarkably, in the
of the pseudopolymorphic forms are highlighted by structure of complex 3 the favoured H⋯F contacts are the
variations in their FPs. The dihydrogen H⋯H molecular privileged ones for the molecular packing (EHF = 1.18), while
contacts comprising 32.3% and 31.5% of the overall contacts the same contacts in complex 4 are even more favoured (EHF
for complexes 3 and 4 are the major contributors to the = 1.52). The halogen bonding interactions (C–F⋯F–C) are
crystal packing. The shortest H⋯H contacts are evident in enriched in complex 3 with EFF = 1.45. These
the FPs as a characteristic spike at de + di = 2.3–2.4 Å and the halogen⋯halogen contacts are not favoured in complex 4
splitting of the short H⋯H is identified in the fingerprint of since they are absent in the molecular interaction surface.
complex 4. The structures of the complexes are also The H⋯C contacts have shown an increased propensity to
dominated by H⋯C contacts with 25.4% to 14.5% that are form (EHC = 1.13) in the structure of 3 and are less favoured
shown in the form of wings with the shortest de + di ≈ 2.8– (EHC = 0.66) in complex 4. The C⋯C inter-contacts in the
2.9 Å (Fig. 11c and d), which are recognized as the C–H⋯π structure of 4 are greatly enriched (ECC = 2.15), which is due

4356 | CrystEngComm, 2021, 23, 4344–4369 This journal is © The Royal Society of Chemistry 2021
View Article Online

CrystEngComm Paper
Published on 03 June 2021. Downloaded on 8/11/2021 8:41:07 PM.

Fig. 11 FPs and %contribution of various non-covalent interactions in homoleptic [(a) complexes 1 and (b) 2]; and heteroleptic [(c) complexes 3
and (d) 4] Cu(II) complexes.

This journal is © The Royal Society of Chemistry 2021 CrystEngComm, 2021, 23, 4344–4369 | 4357
View Article Online

Paper CrystEngComm

Table 2 The enrichment ratios (ERs) for the β-diketone Cu(II) complexes

Complex 1 Complex 2
H C F O S Cu H C F O S Cu
a a
H 21 Contacts 38.7 Contacts (%)
(%)
C 24.2 2.2 11.2 6
F 30.8 5.4 2 23.3 7 0.8
O 9.2 2.4 9.5
S 1.2 0.6 3.5
Cu 1
Sx 53.7 18.7 20.4 5.8 0.9 0.5 62.45 15.1 15.95 4.75 1.75
Random contacts Random contacts
H 28.84 39.00
C 20.08 3.50 18.86 2.28
F 21.91 7.63 4.16 19.92 4.82 2.54
O 6.23 2.17 2.37 5.93 1.43 1.52
S 0.97 2.19
Published on 03 June 2021. Downloaded on 8/11/2021 8:41:07 PM.

Cu
Enrichment ratiob Enrichment ratiob
H 0.73 0.99
C 1.20 2.63
F 1.41 1.17 1.45
O 1.48 1.11 1.60
S 1.24 1.60
Cu

Complex 3 Complex 4
Atoms H C F O N Cu H C F O N Cu
a a
H 32.3 Contacts (%) 31.5 Contacts (%)
C 25.4 2.7 14.5 8.4
F 20.9 2.4 3.2 19.9 1.3
O 8 1.9 12.5 3.5 2.2 0.6
N 0.6 1.2 1.00 1.5
Cu 0.1 1.3 1.20 1.9
Sx 59.8 18.8 14.85 4.95 0.9 0.7 56.05 19.75 11.7 9.7 1.25 1.55
Random contacts Random contacts
H 35.76 31.42
C 22.48 3.53 22.13 3.90
F 17.76 5.58 2.21 13.11 4.62 1.37
O 5.92 1.86 1.47 10.87 3.83 2.27
N 1.08 1.40
Cu 0.84 1.73
Enrichment ratiob Enrichment ratiob
H 0.90 1.00
C 1.13 0.76 0.66 2.15
F 1.18 1.45 1.52
O 1.35 1.02 1.15 0.91 0.97
N
Cu
a
Values are obtained from CrystalExplorer. b
The enrichment ratios were not computed when the “random contacts” were lower than 0.9%.

to a relatively higher value of their contribution to the total representing the two close points in the crystal structure
Hirshfeld surface compared to 3. The chalcogen O⋯C and landscape. Hence, interaction energy calculations were
O⋯F contacts have shown the propensity to form in structure performed for the homoleptic (1 and 2) and heteroleptic
4, while in complex 3, only the O⋯C contacts are enriched (3 and 4) solvatomorphs to evaluate the energetics
with an ER of unity. associated with the various inter-contacts in the crystal
packing. The interaction topologies of the complexes are
compared systematically, and their results were visualized
3.4 Interaction energy and energy framework analysis via energy frameworks. Herein the calculated energy is for
The analysis of the various physical characterizations of the molecule–molecule interactions of a specific molecular
quasi-isostructural polymorphs shows no explicit group and not only as specific atom–atom contacts with
differences between them, suggesting that they might be any directional specificity.60,61

4358 | CrystEngComm, 2021, 23, 4344–4369 This journal is © The Royal Society of Chemistry 2021
View Article Online

CrystEngComm Paper

Analysis of the energy framework corresponding to packing. Interestingly, both columbic and dispersion energy
complexes 1 and 2 shows a comparable architecture in their frameworks exhibit similarity in both polymorphs 1 and 2,
molecular interaction topologies (Fig. 12a and b). A 3-D with dominating dispersion components. Further, the net
energy framework along the a-axis shows the layer-like total interaction energy of a shell of neighboring complex
molecular crystal packing arrangement. However, a variation molecules around the central symmetry independent
in the cylinders' dimension (columns, pillars and crossbars) molecule was evaluated. The neighboring-shell interaction
in the frameworks is noticed. The cyclic pair of DMSO solvent energies for complexes 1 and 2 were tabulated (Table 3). The
molecules with an R22(8) inversion dimer shows the most analysis of interaction topology substantiates the relatively
relevant molecular interactions with a larger contribution of stronger and more anisotropic nature in quasi-isostructural
the electrostatic component. In the crossbars that connect polymorph 1 compared to 2.
the hexagonal lattice points of the framework, the dispersion The energy frameworks of pseudopolymorphic complexes 3
energy dominantly contributes to stabilizing the crystal and 4 are depicted in Fig. 12c and d, which reveal the
Published on 03 June 2021. Downloaded on 8/11/2021 8:41:07 PM.

Fig. 12 Packing modes and molecular topology of homoleptic [(a) complexes 1 and (b) 2]; and heteroleptic [(c) complexes 3 and (d) 4] Cu(II)
complexes visualized through energy framework diagrams (Eele, Edis and Etot) for a cluster of nearest-neighbor molecules.

This journal is © The Royal Society of Chemistry 2021 CrystEngComm, 2021, 23, 4344–4369 | 4359
View Article Online

Paper CrystEngComm

Table 3 Components of various interaction energies in kJ mol−1 for homoleptic and heteroleptic β-diketone Cu(II) complexes

Coulombic Polarization Dispersion Repulsion Total interaction energy


Complex 1 −165.68 −42.05 −445.45 178.50 −474.69
Complex 2 −226.72 −75.12 −301.74 155.87 −447.72
Complex 3 −29.04 −12.30 −168.21 47.11 −162.44
Complex 4 −47.99 −8.59 −179.38 62.60 −173.36

remarkable difference in their molecular interaction rhombus-shaped architecture for complex 3. The involvement
topologies. The frameworks are assembled in an interlocked of different cylinders' size in the construction of the framework
Published on 03 June 2021. Downloaded on 8/11/2021 8:41:07 PM.

Fig. 13 3D NCI plots and 2D scatter maps for homoleptic [(a) complexes 1 and (b) 2]; and heteroleptic [(c) complexes 3 and (d) 4] Cu(II) complexes
with an isosurface value of 0.5. The surfaces are coloured according to a blue-green-red scale based on values of sign(λ2)ρ(r).

4360 | CrystEngComm, 2021, 23, 4344–4369 This journal is © The Royal Society of Chemistry 2021
View Article Online

CrystEngComm Paper

indicates the stabilization of packing via various strong and their surroundings; herein, interactions are analysed based
weak interactions. The molecular pairs related through on the electron density distribution's topology. The metal
π-interactions show higher energy, and the relatively lower chelating bonding type and its strength are closely related to
energy is subsidized by C–H⋯F, F⋯F and di-hydrogen ρ and the sign of ∇ 2ρ(r) at the bond critical point. The
interactions. In complex 4, significant molecular interactions reduced density gradient-based NCI indexed model
are observed between the acetic acid and complex molecule (isosurface = 0.05) and 2-D scatter plots provide detailed
with a larger contribution of the electrostatic component. information about non-covalent interactions present in the
Further, π–π and other weak non-covalent interactions with pseudopolymorphic complexes.62–64
dominant dispersion energy components also contributed to The interactions between the fragments of homoleptic
the molecular packing. Complex 3 shows similarity in the complexes 1 [Cu(TFNB)2DMSO] and 2 [Cu(TFNB)2(DMSO)2]
electrostatic and dispersion energy framework, whereas in are identified; the red discs in the middle of the chelating
complex 4, a substantial difference is noticed (Fig. 12c and d). rings and the middle of the naphthyl rings indicate the effect
The above analysis brings out the quantitative interaction of ring strain (Fig. 13a and b). Strong electrostatic
energy values of the central molecule with the neighboring- interactions between the solvent molecule (DMSO) and Cu(II)
shell molecular fragments in the homoleptic and heteroleptic metal atom in 1 and 2 are shown as blue-green disks. Weak
Published on 03 June 2021. Downloaded on 8/11/2021 8:41:07 PM.

pseudopolymorphs. The results revealed a close similarity in intramolecular dihydrogen (H⋯H) contacts, C–H⋯O
the packing framework with slightly varied interaction hydrogen bonding interactions between the chelating oxygen
energies in complexes 1 and 2, whereas in complexes 3 and atom and naphthyl –CH group, and C–H⋯F contacts formed
4, a remarkable difference is exhibited by molecular pairwise by the CF3 moiety are identified as small green discs in both
interactions to construct distinct interaction topology with complexes. In complex 2, the hydrogen bonding interaction
different energies. between the methyl group of the coordinating solvent and
metal chelating atom is also present. The topological
properties of the electron density in the BCPs confirm the
3.5 QTAIM analysis coordination bonding nature of the apically bridged Cu–O
To understand the supramolecular organization of a complex bond in both forms, and it is validated by the higher Eint
in the crystalline phase, the non-covalent interactions are values. A relatively stronger metal ion–solvent coordination
investigated using the QTAIM molecular graph. The strength (Cu–O) is observed for complex 1 with an Eint of 57.49 kcal
of contacts depends on the nature of interacting atoms and mol−1 (Table 4). The low-gradient spike (low density) lying at

Table 4 The electron density [ρ(r)] and its Laplacian [∇ 2ρ(r)], kinetic energy density [G(r)], potential energy density [V(r)], energy density [H(r)] and
interaction energy (Eint) for bonding and non-bonding interactions of β-diketone Cu(II) complexes

Bonding interaction ρ(r) (a.u.) ∇ 2ρ(r) (a.u.) G(r) (a.u.) V(r) (a.u.) H(r) (a.u.) Eint (kcal mol−1)
Complex 1 Cu1–O1 0.0730 0.5006 0.1229 −0.1207 0.0021 158.44
Cu1–O2 0.0715 0.4909 0.1199 −0.1171 0.0028 153.72
Cu1–O3 0.0722 0.4961 0.1214 −0.1189 0.0025 156.08
Cu1–O4 0.0723 0.4947 0.1212 −0.1187 0.0024 155.82
Cu1–O5 0.0384 0.2277 0.0504 −0.0438 0.0065 57.49
H3⋯H6 0.0121 0.0469 0.0094 −0.0071 0.0023 9.32
C15–H15⋯O3 0.0027 0.0102 0.0020 −0.0015 0.0005 1.96
Complex 2 Cu1–O2 0.0690 0.0471 0.1144 −0.1110 0.0034 145.71
Cu1–O1 0.0694 0.4746 0.1153 −0.1120 0.0032 147.02
Cu1–O3 0.0303 0.1505 0.0338 −0.0300 0.0037 39.38
H3⋯H6 0.0116 0.0455 0.0091 −0.0069 0.0022 9.05
C10–H10⋯O1 0.0074 0.0248 0.0053 −0.0044 0.0008 5.77
C15–H15⋯O2 0.0043 0.0164 0.0033 −0.0026 0.0007 3.41
Complex 3 Cu1–O1 0.0478 0.3807 0.0884 −0.0816 0.0067 107.12
Cu1–N1 0.0505 0.3399 0.0828 −0.0807 0.0021 105.93
Cu1–O2 0.0463 0.3335 0.0769 −0.0704 0.0064 92.41
Cu1–O3 0.0534 0.4233 0.1000 −0.0942 0.0057 123.66
Cu1–Oi 0.0308 0.1396 0.0319 −0.0289 0.0029 37.93
C15–H15⋯O3 0.0067 0.0306 0.0061 −0.0047 0.0014 6.16
H3⋯H6 0.0101 0.0454 0.0086 −0.0058 0.0027 7.61
C3⋯C21 0.0019 0.0058 0.0011 −0.0008 0.0003 1.05
Complex 4 Cu1–O1 0.0530 0.3733 0.0893 −0.0854 0.0039 112.10
Cu1–N1 0.0613 0.3472 0.0903 −0.0939 −0.0035 123.26
Cu1–O2 0.0523 0.4069 0.0963 −0.0909 0.0053 119.32
Cu1–O3 0.0559 0.4305 0.1033 −0.0990 0.0042 129.96
Cu1–O5 0.0164 0.1023 0.0219 −0.0183 0.0036 24.02
H3⋯H6 0.0103 0.0464 0.0087 −0.0059 0.0028 7.74
C15–H15⋯O3 0.0080 0.0360 0.0072 −0.0055 0.0017 7.22

This journal is © The Royal Society of Chemistry 2021 CrystEngComm, 2021, 23, 4344–4369 | 4361
View Article Online

Paper CrystEngComm

negative values (−0.02 to 0) indicates the attractive non- with the experimental data and an overlay of the optimized
covalent interactions. The spike in the region −0.03 to −0.04 electronic structure with the crystal structure is displayed in
corresponds to the stronger and more stabilizing interactions Fig. S3.† Root mean square deviations (RMSDs) of 0.372 Å
attributed to metal ion–solvent coordination. The red and 0.017 Å are observed for complexes 1 and 2. The 3D
coloured spikes reflect the steric repulsion due to the non- plots of the three highest occupied and lowest unoccupied
bonded overlap at the ring centres in the region (+0.01 to molecular orbitals of the pseudopolymorphs 1 and 2 are
+0.02) in complexes 1 and 2. The blue coloured spike displayed in Fig. 14 (red and green colored areas in the
evidences the stronger Cu1–O5 coordination in complex 1. MOs indicate the positive and negative regions). The frontier
Besides in complex 2, the dominance of van der Waals molecular orbital (FMO) plots show the spatial distribution
interactions is shown by the green coloured region. and behaviour of the electron system in the molecule. The α
The QTAIM molecular graph and 2-D scatter plots of and β MOs of complex 1 show a remarkable variation in
heteroleptic solvatomorphs 3 and 4 show the significant electron density spatial distribution. In the α MOs of 1, the
difference in the strength of intramolecular interactions HOMO is localized on naphthyl rings and the LUMO is
(Fig. 13c and d). Dihydrogen H⋯H, C–H⋯O and C–H⋯F located over the whole molecule. Meanwhile in the β MOs,
interactions (green coloured discs) are observed in both the HOMO is distributed on the naphthyl ring and the
Published on 03 June 2021. Downloaded on 8/11/2021 8:41:07 PM.

complexes. Further, in dinuclear complex 3, two monomer LUMO is delocalized on the coordination sphere (Fig. 14a).
units are connected via a bridged coordination bond with an The HOMO–LUMO energy gap (Eg) is a critical parameter to
Eint of 37.93 kcal mol−1 (green discs) and stabilized by determine the complex's molecular electrical transport
chalcogen C⋯O and C⋯C (π⋯π) interactions. In complex 4, properties. The Eg results from molecular charge transfer
the lattice solvent interacts with the complex via strong O– from the proficient electron donor to the acceptor group
H⋯O contacts and metal–oxygen (Cu⋯O) interaction is also help to evaluate the redox potential. The energy gap of the α
observed. Furthermore, the interactions are also identified by and β MOs of complex 1 is 3.7307 and 2.8226 eV,
the framework of Bader's QTAIM method and presented. The respectively, showing the significant HOMO–LUMO energy
topological properties of the electron density in the bond difference for the β-MOs. The energy gap of the α and β
critical points suggest that in solvates (complex 1 and 2), the MOs of complex 2 is 4.0001 and 3.2548 eV, respectively
solvent (DMSO) is weakly bound to Cu(II) with a high ionic (Fig. 14b). Relatively lower Eg values are observed for the cis
character, whereas in 4, hydrogen bonding is responsible for conformation (1) than the trans (2) isomer. The complexes'
the position of the acetic acid solvent. The strength of global reactive parameters are given in Table 5, which
chelating coordination bonds in homoleptic complexes 1 and reveals that complex 1 is softer and more reactive than 2
2 is relatively stronger than that in heteroleptic complexes 3 with higher electronegativity.
and 4. The metal ion–solvent (Cu1–O5) coordination in Complexes 3 and 4 are pseudopolymorphic mixed ligand
complex 1 is stronger than the μ-phenoxide bridged (Cu1–Oi) Cu(II) complexes, in which 3 is dinuclear, and 4 is a
coordination in complex 3. mononuclear unit with one extra solvent (acetic acid)
molecule in the lattice. The electronic structure's geometry
optimization reveals that the dinuclear complex (3) is in a
3.6 Density functional theory calculations singlet state and the mononuclear complex (2) is in a double
3.6.1 Molecular geometry optimization. Quantum state. The optimized ground state energy of structures 3 and
computational investigations were performed to advance the 4 is −3322.86 and −1890.67 Hartree, respectively. The
knowledge of the electronic structure of quasi-isostructural superposition of optimized and X-ray single crystal structures
polymorphic complexes. Density functional theory is shown in Fig. S6† with an RMSD of 0.736 and 0.28 Å. The
calculations were used for the ground state optimization to frontier Kohn–Sham molecular orbitals of complexes 3 and 4
achieve the complexes' possible accurate electronic structure with the orbital energy level diagram (Fig. 15) disclose the
conformation. The reliability of the DFT functionals (B3LYP distinct nature of charge distribution on the molecular
and LanL2DZ) is verified by comparing them with the single orbital with different energies. In complex 3, the HOMO
crystal structure results. The electronic structure of the Cu(II) (−5.3821 eV) and LUMO (−4.5174 eV) are mainly delocalized
pseudopolymorphic complexes is reasonably well predicted on the metal chelate rings and 8HQ ligand with a small
with the coordination sphere. The results reveal that B3LYP difference in their energies (Fig. 15a). The most significant
is more consistent for all the calculations from the HOMO–LUMO energy gap of 0.8648 eV is observed for
perspective of geometry conformation.65,66 complex 3 which signifies that the Cu(II) ion has strong
Experimentally, the homoleptic Cu(II) complexes (1 and 2) potential to accept electrons from the bridged phenolate
exist in two different (cis and trans) conformations. The oxygen and chelating atoms. In the α MOs of 4, the HOMO is
energetic difference between the conformations (1 and 2) localized on the 8HQ ring, and the LUMO is majorly located
was computed by geometry optimization in the doublet over the TFNB ligand. Meanwhile in the β MOs, the HOMO is
state. The ground state energy of the optimized electronic distributed on the 8HQ ring and the LUMO is delocalized on
structures is −2726.68 and −3279.93 Hartree for 1 and 2. The the metal coordination sphere (Fig. 15b). The energy gap of
calculated geometrical parameters are in good agreement the α and β MOs of complex 4 is 3.3769 and 2.5138 eV,

4362 | CrystEngComm, 2021, 23, 4344–4369 This journal is © The Royal Society of Chemistry 2021
View Article Online

CrystEngComm Paper
Published on 03 June 2021. Downloaded on 8/11/2021 8:41:07 PM.

Fig. 14 Frontier Kohn–Sham molecular orbitals of homoleptic β-diketone complexes 1 (a) and 2 (b) with energy level diagrams.

Table 5 The energy values and chemically reactive descriptors of the Cu(II) complexes

Complex 1 Complex 2 Complex Complex 4


Parameter Alpha Beta Alpha Beta 3 Alpha Beta
EHOMO (eV) −6.4442 −6.4439 −6.1982 −6.1794 −5.3821 −5.8932 −5.8616
ELUMO (eV) −2.7135 −3.6213 −2.1981 −2.9247 −4.5174 −2.5162 −3.3478
Energy gap (eV) 3.7307 2.8226 4.0001 3.2548 0.8648 3.3769 2.5138
Ionization energy (eV) 6.4442 6.4439 6.1982 6.1794 5.3821 5.8932 5.8616
Electron affinity (eV) 2.7135 3.6213 2.1981 2.9247 4.5174 2.5162 3.3478
Electronegativity (eV) 4.5789 5.0326 4.1982 4.5521 4.9498 4.2047 4.6047
Chemical potential (eV) −4.5789 −5.0326 −4.1982 −4.552 −4.9498 −4.2047 −4.6047
Global hardness (eV) 1.8653 1.4113 2.0000 1.6274 0.4324 1.6885 1.2569
Global softness (eV−1) 0.5361 0.7086 0.5000 0.6145 2.3127 0.5923 0.7956
Electrophilicity index (eV) 5.6199 8.9729 4.4061 6.3665 28.3310 5.2354 8.4348

respectively. However, the Eg is decreased in heteroleptic Fig. 16a and b show the MEP surface of quasi-isostructure
Cu(II) complexes compared to β-diketone based homoleptic polymorphs 1 and 2. In complex 1, the negative electrostatic
complexes 1 and 2. The reactivity, softness and potential well is situated over the apically coordinated oxygen
electronegativity of the molecules are increased upon the (O5) and chelating (O1 and O3) atoms, and the major
formation of a heteroleptic complex relative to homoleptic nucleophilic region is located around the methyl group (–CH3)
complexes (Table 5). of the DMSO molecule that is involved in the hydrogen bonded
3.6.2 Molecular electrostatic potential (MEP). The network. However, a slightly enriched negative region around
molecular electrostatic potential map of the Cu(II) complexes the electron withdrawing trifluoromethyl group (–CF3) and
based on the electron density distribution helps to interpret the positive values around the naphthyl ring are also observed. In
experimental trends of the structure such as the bonding complex 2, a relatively less concentrated negative region is
geometry and electronic properties of the molecule. To noticed, which is delocalized around the apically coordinated
understand the electrophilic and nucleophilic active sites of the oxygen atoms and other parts of the complex show positive
complexes, three dimensional MEP surfaces were generated. values. Further, a less negative region around the –CF3 group
The red, green and blue colours indicate the most negative, zero and an overall predominated nucleophilic region are observed
electrostatic potential and most positive region, respectively. in complex 2. In complexes 3 and 4, similar kinds of

This journal is © The Royal Society of Chemistry 2021 CrystEngComm, 2021, 23, 4344–4369 | 4363
View Article Online

Paper CrystEngComm
Published on 03 June 2021. Downloaded on 8/11/2021 8:41:07 PM.

Fig. 15 Frontier Kohn–Sham molecular orbitals of heteroleptic complexes 3 (a) and 4 (b) with energy level diagrams.

Fig. 16 Three-dimensional MEP plots for homoleptic [(a) complexes 1 and (b) 2]; and heteroleptic [(c) complexes 3 and (d) 4] Cu(II) complexes.

4364 | CrystEngComm, 2021, 23, 4344–4369 This journal is © The Royal Society of Chemistry 2021
View Article Online

CrystEngComm Paper

Table 6 The natural atomic charges and valence electron configurations of Cu(II) complexes

Atom Charge Valence electron configuration Atom Charge Valence electron configuration
Complex 1 Complex 2
Cu1 1.104 [core] 4S(0.25)3d(9.30)4p(0.35) Cu1 1.075 [core] 4S(0.24)3d(9.28)4p(0.40)
O1 −0.668 [core] 2S(1.67)2p(4.98)3p(0.01) O1 −0.951 [core] 2S(1.82)2p(5.12)3p(0.01)
O2 −0.672 [core] 2S(1.67)2p(4.98)3p(0.01) O2 −0.642 [core] 2S(1.66)2p(4.96)3p(0.01)
O3 −0.668 [core] 2S(1.67)2p(4.98)3p(0.01) O3 −0.662 [core] 2S(1.67)2p(4.98)3p(0.01)
O4 −0.668 [core] 2S(1.67)2p(4.98)3p(0.01)
O5 −1.000 [core] 2S(1.82)2p(5.16)3p(0.01)
Complex 3 Complex 4
Cu1 0.909 [core] 4S(0.25)3d(9.51)4p(0.21)5p(0.12) Cu1 0.864 [core] 4S(0.15)3d(4.36)4p(0.13)
O1 −0.684 [core] 2S(1.67)2p(5.01)3p(0.01) O1 −0.292 [core] 2S(0.83)2p(2.45)
N1 −0.492 [core] 2S(1.32)2p(4.16)3p(0.01) N1 −0.208 [core] 2S(0.65)2p(2.05)4p(0.01)
O2 −0.634 [core] 2S(1.66)2p(4.96)3p(0.01) O2 −0.300 [core] 2S(0.83)2p(2.46)
O3 −0.648 [core] 2S(1.67)2p(4.97)3p(0.01) O3 −0.342 [core] 2S(0.84)2p(2.50)
Published on 03 June 2021. Downloaded on 8/11/2021 8:41:07 PM.

distributions are noticed. A small negative potential is localized formation of the Cu(II) complex with a significant
around the TFNB chelating atoms and –CF3 group, whereas the contribution of 2s and 2p orbitals.
positive region is distributed on 8HQ and the naphthyl ring NBO results reveal that in homoleptic complexes 1 and 2,
(Fig. 16c and d). Further, the MEP of complex 4 featured a the chelating oxygen atoms of the TFNB ligand transfer similar
slightly intense red coloured region that is due to the presence charges and the apically coordinated monodentate (DMSO)
of nucleophilic acetic acid in the lattice. ligand donates a slightly different charge to the central metal
3.6.3 Natural bonding orbital (NBO) analysis. NBO ion to form the stable metal complex. Table 7 summarizes the
analysis gives the most accurate natural Lewis structure that bonding and anti-bonding interactions in the structure
includes the highest possible percentage of the electron obtained from the second-order perturbation theory analysis of
densities and provides insight into the nature of the the Fock matrix based on NBO. In the ground state structure,
complex's electronic structure and bonding interactions. The the chelating oxygen atoms with lone electron pairs act as the
strong intramolecular hyperconjugation interactions of the major donor orbital for the complexes' molecular interaction. In
lone pair, σ and π electron systems lead to the stabilization complex 1, the lone pair–π interaction from the donor orbital of
of bonding or anti-bonding orbitals. The natural charges and Cu1–O4 with Cu1 (s: 0.18%; p: 92.07%; d: 0.52) and O4 (p:
valence electron configuration of the atoms in the 99.93%; d: 0.07) to the acceptor orbital of C18 (p: 99.96%; d:
coordination sphere of all the complexes are tabulated 0.04) is the most stabilizing hyperconjugative interaction with
(Table 6). The changes in the charge on the Cu2+ cationic an energy of 75.12 kcal mol−1 (Table 7). In complex 2, σ–lone
centre and the charge value of all the ligating atoms indicate pair interaction from the donor orbital of O1–C2 with O1 (s:
the sharing of electron density to the Cu(II) ion for the 0.01%; p: 99.92%, d: 0.07%) and C2 (p: 99.90%; d: 0.10) to the
formation of the metal complex. In the above coordination acceptor orbital of C4 (s: 0.03%; p: 99.91%; d: 0.06%) is
complexes, the Cu(II) ion coordinates to the ligand through observed with an energy of 85.88 kcal mol−1 (Table 8).
the 3d orbitals with smaller contributions of 4s and 4p In heteroleptic complexes 3 and 4, the chelating oxygen
orbitals. Besides, all the donor atoms participate in the atoms of both ligands (TFNB and 8HQ) transfer similar

Table 7 Second-order perturbation theory analysis of the Fock matrix in NBO basis for homoleptic Cu(II) complex 1

No. Type Donor Occupancy Type Acceptor Occupancy E (2) (kcal mol−1) Ei − Ej F(i, j)
1 π Cu1–O4 0.803 LP*(1) C18 0.3865 75.12 0.17 0.158
2 π C16–C17 0.8648 LP*(1) C18 0.3865 29.61 0.16 0.099
3 π C12–C11 0.8643 LP*(1) C8 0.4934 22.68 0.15 0.093
4 π C20–C21 0.8728 LP*(1) C22 0.4935 19.56 0.16 0.087
5 LP (2) O3 0.8579 LP*(6) Cu1 0.1226 13.92 0.55 0.112
6 LP (2) O1 0.8595 LP*(6) Cu1 0.1226 13.91 0.55 0.112
7 σ C13–H13 0.9897 σ* Cu1-O4 0.0771 13.49 0.14 0.057
8 LP (2) O2 0.8615 LP*(6) Cu1 0.1226 13.46 0.56 0.111
9 LP (2) O1 0.8595 LP*(5) Cu1 0.3252 13.07 0.22 0.075
10 LP (2) O4 0.8626 LP*(6) Cu1 0.1226 13.06 0.56 0.109
11 LP (2) O4 0.8626 LP*(5) Cu1 0.3252 11.28 0.23 0.071
12 LP (2) O3 0.8579 LP*(5) Cu1 0.3252 11.21 0.22 0.069
13 LP (2) O2 0.8615 LP*(5) Cu1 0.3252 10.16 0.23 0.068
14 σ C14–C13 0.9902 σ* Cu1–O4 0.0771 9.02 0.81 0.112

E (2): energy of hyperconjugative interaction. E(i) − E(j): energy difference between (i) donor and (j) acceptor NBO orbitals. F(i, j): Fock matrix
element between (i) and (j) NBO orbitals.

This journal is © The Royal Society of Chemistry 2021 CrystEngComm, 2021, 23, 4344–4369 | 4365
View Article Online

Paper CrystEngComm

Table 8 Second-order perturbation theory analysis of the Fock matrix in NBO basis for homoleptic Cu(II) complex 2

No. Type Donor Occupancy Type Acceptor Occupancy E (2) (kcal mol−1) Ei − Ej F(i, j)
1 σ* O1–C2 0.8115 LP*(1) C4 0.3901 85.88 0.15 0.16
2 π C14–C13 0.8677 LP*(1) C8 0.4973 44.23 0.04 0.067
3 LP (3) O2 0.8035 σ* O1–C2 0.8115 21.12 0.02 0.037
4 π C10–C9 0.8655 LP*(1) C8 0.4973 20.06 0.15 0.088
5 π C5–C6 0.8528 LP*(1) C4 0.3901 16.08 0.15 0.072
6 LP (2) O1 0.8692 LP*(6) Cu1 0.1213 13.03 0.49 0.101
7 LP (2) O2 0.8669 LP*(6) Cu1 0.1213 12.98 0.48 0.1
8 LP (2) O2 0.8669 LP*(5) Cu1 0.3104 11.98 0.21 0.071
9 LP (2) O1 0.8692 LP*(5) Cu1 0.3104 11.59 0.22 0.071
10 LP (2) O1 0.8692 LP*(8) Cu1 0.0697 11.14 0.65 0.111
11 LP (2) O2 0.8669 LP*(7) Cu1 0.0734 11.13 0.62 0.108
12 LP (1) O1 0.966 LP*(8) Cu1 0.0697 10.76 0.82 0.121
13 σ* O1i–C2i 0.8115 LP*(1) C4i 0.3901 10.64 1.44 0.175

E (2): energy of hyperconjugative interaction. E(i) − E(j): energy difference between (i) donor and (j) acceptor NBO orbitals. F(i, j): Fock matrix
element between (i) and (j) NBO orbitals.
Published on 03 June 2021. Downloaded on 8/11/2021 8:41:07 PM.

Table 9 Second-order perturbation theory analysis of the Fock matrix in NBO basis for heteroleptic dinuclear complex 3

Q1 Type Donor Occupancy Type Acceptor Occupancy E (2) (kcal mol) Ei − Ej F(i, j)
1 LP (3) O2 1.5652 LP*(1) C4 0.778 174.84 0.16 0.163
2 LP (3) O1 1.6368 LP*(5) Cu1 1.5594 61.12 0.02 0.043
3 π C3–C2 1.7347 LP*(1) C4 0.7782 58.48 0.16 0.099
4 LP (2) O3 1.7762 LP*(6) Cu1 0.2386 30.51 0.56 0.117
5 LP (1) N1 1.7835 LP*(6) Cu1 0.2386 27.95 0.54 0.110
6 LP (2) O2 1.8097 LP*(6) Cu1 0.2386 25.86 0.56 0.109
7 LP (2) O1 1.8191 LP*(7) Cu1 0.1186 25.02 0.69 0.119
8 LP (2) O1i 1.8200 LP*(7) Cu1i 0.1185 24.98 0.69 0.119
9 LP (1) N1 1.7835 LP*(8) Cu1 0.1122 23.69 0.63 0.112
10 LP (2) O2 1.8097 LP*(8) Cu1 0.1122 22.44 0.65 0.110
11 LP (1) O3 1.9311 LP*(7) Cu1 0.1186 19.50 0.87 0.117
12 LP (1) O1i 1.9110 LP*(9) Cu1 0.0886 16.07 0.66 0.092
13 LP (1) O1 1.9110 LP*(9) Cu1i 0.0884 15.93 0.66 0.092
14 LP (1) O1 1.9110 LP*(7) Cu1 0.1186 15.81 0.8 0.101
15 LP (1) O1i 1.9110 LP*(7) Cu1i 0.1185 15.39 0.8 0.100
16 LP (2) O1 1.8191 LP*(6) Cu1 0.2386 14.27 0.58 0.083
17 LP (2) O3 1.7762 LP*(5) Cu1 1.5594 12.74 0.13 0.064
18 LP*(6) Cu1 0.2386 LP*(7) Cu1i 0.1185 12.26 0.11 0.076
19 LP (2) O1 1.8191 LP*(9) Cu1i 0.0884 10.63 0.55 0.070

E (2): energy of hyperconjugative interaction. E(i) − E(j): energy difference between (i) donor and (j) acceptor NBO orbitals. F(i, j): Fock matrix
element between (i) and (j) NBO orbitals.

charges, and the –N donor atom shares a slightly different 4. Conclusion


charge to the central metal ion to form the stable Cu(II)
complex. In complex 3, the chelating atoms (O and N) of the We have demonstrated the structural diversity of
ligands are the dominant donor orbitals that stabilize the coordination complexes induced by the crystallization solvent
molecule with the major lone pair molecular interactions. via transformations in the coordination sphere and
Meanwhile π–π, lone pair–π, and lone pair–lone pair supramolecular self-assembly of crystal motifs. Homoleptic
interactions are the major stabilizing molecular interactions in (β-diketone) Cu(II) complex 1 is the isostructure of previously
complex 4. The stabilizing hyperconjugative interaction in reported complex (2) which exhibits a cis and trans
complex 3 is between the donor orbital of O2 (s: 0.02%, p: configuration, respectively, and is achieved by the PSM
99.99%, d: 0.07%) and the acceptor orbital of C4 (p: 99.94%, d: method. Complex 1 shows a five coordinated distorted square
0.06%) with an energy of 174.84 kcal mol−1 (Table 9). In pyramidal geometry, whereas complex 2 exhibits a six
complex 4, π*–π* hyperconjugative interaction from the donor coordinated octahedral geometry around the metal centre.
orbital of C5–C6 with C5 (s: p: 99.97%, d: 0.02%) and C6 (p: The primary structural motifs that establish the net
99.94%; d: 0.05) to the acceptor orbital of C14–C13 with C14 (p: supramolecular architecture are linked by C–H⋯O, C–H⋯F,
99.95%; d: 0.05%) and C13 (p: 99.94%; d: 0.06%) is observed C–H⋯S, C⋯O (tetrel) and π⋯π interactions in complex 1,
with an energy of 89.13 kcal mol−1 (Table 10). whereas in 2, C–H⋯O contacts have governed the packing of

4366 | CrystEngComm, 2021, 23, 4344–4369 This journal is © The Royal Society of Chemistry 2021
View Article Online

CrystEngComm Paper

Table 10 Second-order perturbation theory analysis of the Fock matrix in NBO basis for heteroleptic mononuclear complex 4

Q2 Type Donor Occupancy Type Acceptor Occupancy E (2) (kcal mol−1) Ei − Ej F(i, j)
1 π* C5–C6 0.1487 π* C14–C13 0.1075 89.13 0.01 0.077
2 LP (1) C18 0.5005 π* N1–C23 0.3057 73.31 0.09 0.115
3 σ Cu1–O3 0.8050 LP*(1) C4 0.3846 72.88 0.17 0.157
4 LP*(1) C15 0.4452 π* N1–C23 0.3057 43.47 0.11 0.1
5 π* N1–C23 0.3097 π* C22–C21 0.1664 34.44 0.06 0.082
6 LP (3) O2 0.8202 π* C2–C3 0.1552 34.21 0.3 0.13
7 LP (1) C7 0.5041 π* C5–C6 0.1486 32.32 0.14 0.104
8 LP (1) C18 0.5005 π* C16–C17 0.1149 29.53 0.14 0.104
9 π C2–C3 0.8703 LP*(1) C4 0.3846 28.64 0.16 0.098
10 π N1–C23 0.8546 LP*(1) C15 0.4452 28.18 0.2 0.112
11 LP*(1) C12 0.4917 π* C11–C10 0.1226 27.34 0.14 0.1
12 π C16–C17 0.8478 LP*(1) C15 0.4452 27.05 0.14 0.093
13 LP (1) C7 0.5041 π* C8–C9 0.1205 26.89 0.15 0.1
14 LP (1) N1 0.8981 LP*(6) Cu1 0.1360 21.73 0.59 0.145
15 π C5–C6 0.8461 LP*(1) C4 0.3846 21.35 0.14 0.078
16 LP (2) O1 0.8944 π* C22–C21 0.1664 21.3 0.32 0.107
Published on 03 June 2021. Downloaded on 8/11/2021 8:41:07 PM.

17 LP*(1) C15 0.4452 π* C16–C17 0.1149 21.24 0.16 0.098


18 π C11–C10 0.8626 LP*(1) C12 0.4917 20.99 0.15 0.089
19 LP*(1) C4 0.491 π* C2–C3 0.1552 19.7 0.13 0.087
20 LP (1) N1 0.8312 LP*(5) Cu1 0.3787 19.65 0.18 0.085
21 LP (3) O1 0.8364 LP*(5) Cu1 0.3787 19.1 0.15 0.076
22 LP (2) O2 0.8524 LP*(6) Cu1 0.1360 17.8 0.6 0.132
23 LP (2) O3 0.8606 LP*(6) Cu1 0.1360 15.48 0.62 0.124

E (2): energy of hyperconjugative interaction. E(i) − E(j): energy difference between (i) donor and (j) acceptor NBO orbitals. F(i, j): Fock matrix
element between (i) and (j) NBO orbitals.

the molecules in the crystalline phase. The energy gap of completion of single crystal structural studies, data analysis
2.8632 eV for complex 1 unveils that the cis conformer is and all the quantum computational studies.
chemically more active compared to the trans conformer. In
the TFNB and 8HQ mixed ligand Cu(II) complex, the solid Conflicts of interest
state modification leads to the transformation from a
dinuclear (3) into a mononuclear (4) pseudopolymorphic The authors declare that they have no known competing
form. The dinuclear complex is stabilized by significant financial interests or personal relationships that could have
dihalogen C–F⋯F–C interactions, whereas in the appeared to influence the work reported in this article.
mononuclear complex, the lattice solvent plays a crucial role
in the construction of the supramolecular architecture via O– Acknowledgements
H⋯O and C–H⋯O contacts. In-depth analysis of the strength The authors are thankful to the DST-FIST, National Single
and nature of non-covalent interactions by structural and Crystal Diffractometer Facility, DoS in Physics, CPEPA, IOE
computational methodology revealed the balance between and DST-PURSE, Vijnana Bhavan, University of Mysore,
molecular interaction topology and structural conformation Mysuru. The author Mahesha would like to thank the DST-
in isostructures and pseudopolymorphs. This work revealed KSTePS Govt. of Karnataka for providing the fellowship.
that the SCCSE method is probably more common and could
be applied as a general PSM route for homoleptic complexes References
to prepare new crystal forms of the complexes. In heteroleptic
complexes with two or more chelating ligands, SCCSE rarely 1 J. D. Dunitz, The Crystal as a Supramolecular Entity, ed. G. R.
occurs; in this case, anion exchange and lattice solvent Desiraju, Perspectives in Supramolecular Chemistry, Wiley,
incorporation would be the preferable route to obtain more Chichester, 1996.
crystal forms of the complex. 2 H. J. Schneider and H. Durr, Frontiers in Supramolecular
Organic Chemistry and Photochemistry, VCH, 1991.
Author contributions 3 J. W. Steed and J. L. Atwood, Supramolecular Chemistry, John
Wiley and Sons, 2013.
The research work was designed by NKL and Mahesha. The 4 M. Liu, L. Zhang and T. Wang, Chem. Rev., 2015, 115, 7304–7397.
manuscript was prepared by Mahesha and it was critically 5 A. B. Descalzo, R. Martinez-Manez, F. Sancenon, K. Hoffmann
edited and revised by NKL. Synthesis of the complex was and K. Rurack, Angew. Chem., Int. Ed., 2006, 45, 5924–5948.
carried out by CSK, Mahesha and PM. MKH and KJP 6 J. M. Lehn, Angew. Chem., Int. Ed. Engl., 1990, 29, 1304–1319.
contributed to the synthesis and spectroscopic 7 M. M. Smulders, I. A. Riddell, C. Browne and J. R. Nitschke,
characterization. NKL and Mahesha contributed to the Chem. Soc. Rev., 2013, 42, 1728–1754.

This journal is © The Royal Society of Chemistry 2021 CrystEngComm, 2021, 23, 4344–4369 | 4367
View Article Online

Paper CrystEngComm

8 L. J. Chen and H. B. Yang, Acc. Chem. Res., 2018, 51, 36 G. M. Sheldrick, Acta Crystallogr., Sect. A: Found. Crystallogr.,
2699–2710. 2008, 64, 112–122.
9 R. Chakrabarty, P. S. Mukherjee and P. J. Stang, Chem. Rev., 37 G. M. Sheldrick, Acta Crystallogr., Sect. C: Struct. Chem.,
2011, 111, 6810–6918. 2015, 71, 3–8.
10 M. Han, D. M. Engelhard and G. H. Clever, Chem. Soc. Rev., 38 O. V. Dolomanov, L. J. Bourhis, R. J. Gildea, J. A. Howard
2014, 43, 1848–1860. and H. Puschmann, J. Appl. Crystallogr., 2009, 42,
11 T. R. Cook, Y. R. Zheng and P. J. Stang, Chem. Rev., 339–341.
2013, 113, 734–777. 39 A. L. Spek, Acta Crystallogr., Sect. D: Biol. Crystallogr.,
12 T. R. Cook and P. J. Stang, Chem. Rev., 2015, 115, 7001–7045. 2009, 65, 148–155.
13 X. Wu, D. J. Young and T. A. Hor, J. Mol. Eng. Mater., 2015, 3, 40 C. F. Macrae, I. Sovago, S. J. Cottrell, P. T. Galek, P. McCabe,
1540004. E. Pidcock and P. A. Wood, J. Appl. Crystallogr., 2020, 53,
14 V. Vrdoljak, B. Prugovecki, D. Matkovic-Calogovic, T. Hrenar, 226–235.
R. Dreos and P. Siega, Cryst. Growth Des., 2013, 13, 41 M. J. Turner, J. J. McKinnon, S. K. Wolff, D. J. Grimwood,
3773–3784. P. R. Spackman and D. Jayatilaka, CrystalExplorer17,
15 V. Vrdoljak, G. Pavlovic, T. Hrenar, M. Rubcic, P. Siega, R. University of Western Australia, 2017.
Published on 03 June 2021. Downloaded on 8/11/2021 8:41:07 PM.

Dreos and M. Cindric, RSC Adv., 2015, 5, 104870–104883. 42 M. A. Spackman and D. Jayatilaka, CrystEngComm, 2009, 11,
16 C. P. Li and M. Du, Chem. Commun., 2011, 47, 5958–5972. 19–32.
17 P. P. Bag, R. R. Kothur and C. M. Reddy, CrystEngComm, 43 C. Jelsch, K. Ejsmont and L. Huder, IUCrJ, 2014, 1,
2014, 16, 4706–4714. 119–128.
18 A. N. Usoltsev, T. S. Sukhikh, A. S. Novikov, V. R. Shayapov, 44 C. F. Mackenzie, P. R. Spackman, D. Jayatilaka and M. A.
D. P. Pishchur, I. Korolkov and S. A. Adonin, Inorg. Chem., Spackman, IUCrJ, 2017, 4, 575–587.
2021, 60, 2797–2804. 45 E. Espinosa, E. Molins and C. Lecomte, Chem. Phys. Lett.,
19 W. X. Li, J. H. Gu, H. X. Li, M. Dai, D. J. Young, H. Y. Li and 1998, 285, 170–173.
J. P. Lang, Inorg. Chem., 2018, 57, 13453–13460. 46 T. Lu and F. Chen, J. Comput. Chem., 2012, 33, 580–592.
20 G. Mukherjee and K. Biradha, Chem. Commun., 2012, 48, 47 W. Humphrey, A. Dalke and K. Schulten, J. Mol. Graphics,
4293–4295. 1996, 14, 33–38.
21 P. Deria, J. E. Mondloch, O. Karagiaridi, W. Bury, J. T. Hupp 48 M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria,
and O. K. Farha, Chem. Soc. Rev., 2014, 43, 5896–5912. M. A. Robb, J. R. Cheeseman and X. Li, Gaussian 16, 2016.
22 J. Liu, X. P. Zhang, T. Wu, B. B. Ma, T. W. Wang, C. H. Li 49 R. D. Dennington, T. A. Keith and J. M. Millam, GaussView 6.
and X. Z. You, Inorg. Chem., 2012, 51, 8649–8651. 50 R. Vargas, J. Garza and A. Cedillo, J. Phys. Chem. A,
23 W. Q. Zhang, W. Y. Zhang, R. D. Wang, C. Y. Ren, Q. Q. Li, Y. P. 2005, 109, 8880–8892.
Fan and Y. Y. Wang, Cryst. Growth Des., 2017, 17, 517–526. 51 N. S. Abdel-Kader, S. A. Abdel-Latif, A. L. El-Ansary and A. G.
24 S. Rat, K. Ridier, L. Vendier, G. Molnar, L. Salmon and A. Sayed, New J. Chem., 2019, 43, 17466–17485.
Bousseksou, CrystEngComm, 2017, 19, 3271–3280. 52 A. W. Addison, T. N. Rao, J. Reedijk, J. van Rijn and
25 G. R. Desiraju, J. Am. Chem. Soc., 2013, 135, 9952–9967. G. C. Verschoor, J. Chem. Soc., Dalton Trans.,
26 S. K. Seth, I. Saha, C. Estarellas, A. Frontera, T. Kar and 1984, 1349–1356.
S. Mukhopadhyay, Cryst. Growth Des., 2011, 11, 53 L. Yang, D. R. Powell and R. P. Houser, Dalton Trans.,
3250–3265. 2007, 955–964.
27 P. Panini and D. Chopra, New J. Chem., 2015, 39, 8720–8738. 54 D. G. Evans and J. C. A. Boeyens, Acta Crystallogr., Sect. B:
28 K. Gholivand, K. Farshadfar, S. M. Roe, M. Hosseini and A. Struct. Sci., 1989, 45, 581–590.
Gholami, CrystEngComm, 2016, 18, 7104–7115. 55 D. A. Safin, M. G. Babashkina, K. Robeyns and Y. Garcia,
29 A. Zabardasti, H. Afrouzi, A. Kakanejadifard and Z. Jamshidi, RSC Adv., 2016, 6, 53669–53678.
J. Sulfur Chem., 2017, 38, 249–263. 56 G. Resnati, E. Boldyreva, P. Bombicz and M. Kawano, IUCrJ,
30 M. K. Hema, C. S. Karthik, K. J. Pampa, H. M. Manukumar, 2015, 2, 675–690.
P. Mallu, I. Warad and N. K. Lokanath, Polyhedron, 57 P. Manna, S. K. Seth, A. Das, J. Hemming, R. Prendergast, M.
2019, 168, 127–137. Helliwell and S. Mukhopadhyay, Inorg. Chem., 2012, 51,
31 H. Xu, W. Chen, P. Zhan and X. Liu, MedChemComm, 3557–3571.
2015, 6, 61–74. 58 K. Kaabi, K. Klai, E. Wenger, C. Jelsch, F. Lefebvre and C. B.
32 P. Sanphui and G. Bolla, Cryst. Growth Des., 2018, 18, Nasr, Acta Crystallogr., Sect. C: Struct. Chem., 2020, 76,
5690–5711. 572–578.
33 Mahesha, M. K. Hema, C. S. Karthik, K. J. Pampa, P. Mallu 59 A. Di Santo, H. Perez, G. A. Echeverria, O. E. Piro, R. A.
and N. K. Lokanath, New J. Chem., 2020, 44, 18048–18068. Iglesias, R. E. Carbonio and D. M. Gil, RSC Adv., 2018, 8,
34 Mahesha, M. K. Hema, C. S. Karthik, K. J. Pampa, P. Mallu 23891–23902.
and N. K. Lokanath, Polyhedron, 2020, 185, 114571. 60 M. J. Turner, S. P. Thomas, M. W. Shi, D. Jayatilaka
35 C. C. S. Rigaku, Expert 2.0 r15, Rigaku Corporation, Tokyo, and M. A. Spackman, Chem. Commun., 2015, 51,
Japan, 2011. 3735–3738.

4368 | CrystEngComm, 2021, 23, 4344–4369 This journal is © The Royal Society of Chemistry 2021
View Article Online

CrystEngComm Paper

61 J. Granifo, R. Gavino, S. Suarez and R. Baggio, Acta 64 M. A. Martins, M. Horner, J. Beck, A. Z. Tier, A. L. Belladona,
Crystallogr., Sect. C: Struct. Chem., 2019, 75, 1299–1309. A. R. Meyer and C. P. Frizzo, CrystEngComm, 2016, 18,
62 N. K. Nkungli and J. N. Ghogomu, J. Mol. Model., 2017, 23, 3866–3876.
1–20. 65 N. Benarous, A. Cherouana, E. Aubert, P. Durand and S.
63 E. R. Johnson, S. Keinan, P. Mori-Sanchez, J. Contreras- Dahaoui, J. Mol. Struct., 2016, 1105, 186–193.
Garcia, A. J. Cohen and W. Yang, J. Am. Chem. Soc., 66 W. Li, Y. Lyu, H. Zhang, M. Zhu and H. Tang, Dalton Trans.,
2010, 132, 6498–6506. 2017, 46, 106–115.
Published on 03 June 2021. Downloaded on 8/11/2021 8:41:07 PM.

This journal is © The Royal Society of Chemistry 2021 CrystEngComm, 2021, 23, 4344–4369 | 4369

You might also like