0% found this document useful (0 votes)
17 views9 pages

_

The document discusses the synthesis and characterization of a one-dimensional polymeric copper(II)-hydrazone complex, specifically {[Cu(H0.5L)(µ1,3-SCN)]0.5ClO4·0.5MeOH}n. Various spectroscopic techniques, including IR, UV-Vis, and EPR, were employed to analyze the complex, which exhibits weak antiferromagnetic interactions and non-covalent interactions evaluated through DFT calculations. The study highlights the significance of hydrazone ligands in coordination chemistry and their potential for forming unique supramolecular architectures.

Uploaded by

khalifamostafa90
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
17 views9 pages

_

The document discusses the synthesis and characterization of a one-dimensional polymeric copper(II)-hydrazone complex, specifically {[Cu(H0.5L)(µ1,3-SCN)]0.5ClO4·0.5MeOH}n. Various spectroscopic techniques, including IR, UV-Vis, and EPR, were employed to analyze the complex, which exhibits weak antiferromagnetic interactions and non-covalent interactions evaluated through DFT calculations. The study highlights the significance of hydrazone ligands in coordination chemistry and their potential for forming unique supramolecular architectures.

Uploaded by

khalifamostafa90
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 9

Inorganica Chimica Acta 484 (2019) 95–103

Contents lists available at ScienceDirect

Inorganica Chimica Acta


journal homepage: www.elsevier.com/locate/ica

Research paper

Synthesis, structure, physicochemical characterization and theoretical T


evaluation of non-covalent interaction energy of a polymeric copper(II)-
hydrazone complex

Dipali Sadhukhana,b, , Monami Maitia,c, Antonio Bauzád, Antonio Fronterad, Eugenio Garribbae,
Carlos J. Gomez-Garcíaf
a
Department of Chemistry, Jadavpur University, Kolkata 700032, India
b
Centre de Recherche Paul Pascal (CRPP), 115 Avenue du Dr. A. Schweitzer, Pessac 33600, France
c
Department of Chemistry, Narasinha Dutt College, Howrah 711101, West Bengal, India
d
Departament de Química, Universitat de les Illes Balears, Crta. de Valldemossa km 7.5, 07122 Palma de Mallorca (Balears), Spain
e
Dipartimento di Chimica e Farmacia, Università di Sassari, Via Vienna 2, I-07100 Sassari, Italy
f
Instituto de Ciencia Molecular (ICMol), Dpto. Química Inorgánica. Universidad de Valencia, 46980 Paterna, Spain

A R T I C LE I N FO A B S T R A C T

Dedicated to the memory of Late Prof. Samiran One dimensional polymeric copper-hydrazone complex {[Cu(H0.5L)(µ1,3-SCN)]0.5ClO4·0.5MeOH}n (1) has been
Mitra on his first death anniversary, for his synthesized with Cu(ClO4)2·xH2O and N'-(1-(pyridin-2-yl)ethylidene)acetohydrazide (HL) in presence of NaSCN.
versatile contribution to coordination The ligand and the complex have been characterized by several spectroscopic techniques (IR, UV–Vis and EPR),
chemistry and great inspiration to the authors. cyclic voltammetry and the structure of 1 has been determined by single crystal X-ray diffraction. The complex is
Keywords: an infinite one dimensional polymer bridged by thiocyanate. The magneto-structural correlation has been de-
Copper(II)-hydrazone complex termined and the non-covalent interactions present in the molecule have been energetically evaluated by means
1D chain of DFT calculations.
H-bonds
Antiferromagnetism
π-Hole interaction
MEP analysis

1. Introduction In addition, positive electrostatic potentials can also be found on elec-


tron deficient π-systems and these ‘π-holes’ can also attract electron rich
Hydrazone ligands are a special branch of Schiff bases which can species in a directional manner (perpendicularly to the π-system) in the
accommodate with different coordination modes and oxidation states solid state [7–9]. Conventional π-acidic systems are aromatic rings
according to the demand of variable geometries and valences of metal substituted with electron withdrawing substituents. More unconven-
ions in coordination complexes [1,2]. Moreover, they form interesting tional and minimalist π-holes are, for instance, the N atom in nitro-
hydrogen bonded self assembly in metal-free state as well as in metal compounds such as nitromethane and nitroaromatic species [7], the C
complexes, mainly due to the presence of the NeH functionality ad- atom in carbonyl groups (e.g. amides) or cyanide, the S atom in SO3,
jacent to the azomethine (C]N) chromophore [3]. Moreover, the etc. [8].
electron withdrawing NeH group makes imidic C]N bond highly Coordination chemistry of acetyl hydrazone complexes is more
acidic, hence apart from hydrogen bonding the imidic C]N bond of the limited [10–14] than that of its benzoyl analogue [10,11,15–28]. A few
metal-coordinated Schiff base ligand can also interact with electron rich coordination complexes with NiII, CoII, MnIV and ZnII metal ions and
atoms and develop distinct supramolecular interaction. Chemists hydrazone ligands derived from acetic hydrazide and substituted
working in supramolecular chemistry need a deep understanding of phenol-based carbonyl derivatives such as salicylaldehyde, 2-hydro-
intermolecular interactions, such as hydrogen bonding and related in- xyacetophenone, 2,3-dihydroxybenzaldehyde, etc. have been reported
teractions where electron rich entities interact with a ‘σ-hole’ (a posi- [3,29]. Some of these complexes show interesting hydrogen bonding
tive electrostatic potential along the vector of the covalent bond) [4–6]. interactions leading to multi-dimensional supramolecular architectures.


Corresponding author at: Centre de Recherche Paul Pascal (CRPP), 115 Avenue du Dr. A. Schweitzer, Pessac 33600, France.
E-mail address: [email protected] (D. Sadhukhan).

https://doi.org/10.1016/j.ica.2018.09.031
Received 1 January 2018; Received in revised form 31 July 2018; Accepted 9 September 2018
Available online 11 September 2018
0020-1693/ © 2018 Elsevier B.V. All rights reserved.
D. Sadhukhan et al. Inorganica Chimica Acta 484 (2019) 95–103

An efficient π–π overlap eligible for weak magnetic exchange interac- methanol (100 mL). On refluxing the methanolic solution for 5 h a clear
tions is also observed [29]. It has been noticed that the phenol-based solution of the Schiff base was obtained. White crystalline product was
hydrazone ligands are reluctant to acquire pseudohalides and the obtained by removing the solvent under vacuum. Anal. Calcd for
complexes are usually µ-phenoxo bridged dinuclear [29,30] as well as C9H11N3O (FW: 177.2) (%): C, 61.00; H, 6.26; N, 23.71. Found: C,
µ- and µ3-phenoxo bridged tetranuclear clusters [15,16]. Hydrazone 60.59; H, 6.21; N, 23.79. FT-IR bands (KBr, cm−1): νC=N 1578, νC=O
ligands of alkyl/aryl hydrazides and pyridine based carbonyls such as 1678, νN-H(sym) 3080, νN-H(asym) 3194 cm−1. UV–Vis bands (λmax, nm):
pyridine-2-carboxaldehyde, 2-acetylpyridine etc. easily accommodate π-π* 246, n-π* 298. 1H NMR (δ ppm): 2.36 (3H,s), 2.41 (3H,s), 7.31
terminal and bridging halide and pseudohalide ligands in their metal (1H,t), 7.76 (1H,t), 8.07 (1H,d), 8.61 (1H,d).
complexes [31–34].
In order to explore the weak intermolecular interactions of acetyl 2.2.2. Synthesis of {[Cu(H0.5L)(µ1,3-SCN)]0.5ClO4.0.5MeOH}n (1)
hydrazone, we have prepared one dimensional copper(II) complex with Cu(ClO4)2·xH2O (0.370 g, 1 mmol) was dissolved in 20 mL of acet-
acetyl hydrazone [N'-(1-(pyridin-2-yl)ethylidene)acetohydrazide (HL)] onitrile and 10 mL of the methanolic solution of HL (0.180 g, 1 mmol)
and thiocyanate bridging ligand formulated as {[Cu(H0.5L)(µ1,3-SCN)] was added. A solution of NaSCN (0.08 g, 1 mmol) in 1 mL H2O was
0.5ClO4·0.5MeOH}n (1). The complex has been characterized by several added drop wise to the mixture and it was stirred for 1/2 h with
spectroscopic techniques, elemental and electrochemical analyses and warming. The dark green solution was kept in a refrigerator. Green
magnetic susceptibility measurements that reveal a weak anti- prismatic single crystals were obtained after four days. Yield: 0.283 g
ferromagnetic Cu···Cu interaction that can be well reproduced with a (71%). Anal. Calcd for C10.5H12.5Cl0.5Cu1N4O3.5S1 (FW: 364.08) (%): C,
regular chain model with very weak inter-chain interactions. We also 34.64; H, 3.46; N, 15.39; S, 8.81. Found: C, 34.94; H, 3.44; N, 15.69; S,
analyzed the ability of the imidic C]N bond belonging to the metal- 8.90. FT-IR bands (KBr, cm−1): νC=N 1525, νC=O 1600, νN-H(sym + asym)
coordinated Schiff base ligand to interact with electron rich atoms by 2900–3100, νO-H(MeOH) 2800, νClO4 1110–1030, νSCN 2099 cm−1.
means of DFT calculations. UV–Vis bands (MeOH, 10−5 M), λmax(nm)/ε max(M−1 cm−1)): π-π*
251/79800), n-π* 306/25700), LMCT 363/9600).
2. Experimental An outline of the syntheses of the ligand and the complex is depicted
in Scheme 1.
2.1. Materials and physical measurements
2.3. X-ray data collection and refinement
Caution! Perchlorate salts are potentially explosive and should be
used in small quantity with much care. Intensity data were collected using MoKα radiation with a Bruker
All solvents were of reagent grade and used without further pur- ApexII CCD diffractometer at 120 K for 1. Data collection and data re-
ification. Acetic hydrazide, 2-acetylpyridine and sodium thiocyanate duction were performed with SHELX [37]. The structure was solved by
were purchased from Aldrich Chemical Company. Cu(ClO4)2·xH2O was direct method using the program SIR92 [38] and refined with the
prepared by the treatment of basic copper(II) carbonate, CuCO3·Cu program CRYSTALS [39]. Empirical absorption corrections were car-
(OH)2 (AR grade, E. Merck, India) with 60% perchloric acid (AR grade, ried out by multi scan technique [40]. All non hydrogen atoms were
E. Merck, India) followed by slow evaporation on a steam bath. It was refined anisotropically by full-matrix least-squares based on F. All other
then filtered through a glass frit and stored in a CaCl2 desiccator. H atoms were generated geometrically and were included in the re-
The Fourier Transform Infrared spectra were recorded in the range finement by the riding model approximation. The N–H proton (H31)
4000–400 cm−1 on a Perkin-Elmer RX I FT-IR spectrophotometer with has been refined with 0.5 site-occupancy in correlation to the 0.5
solid KBr pellets. The electronic spectra in HPLC grade acetonitrile were stoichiometry of ClO4− per asymmetric unit. The CleO bond lengths
recorded at 300 K on a Perkin-Elmer Lambda 40 (UV–Vis) spectrometer and the OeCleO bond angles of ClO4− moiety have been restrained to
in a 1 cm quartz cuvette in the range 800–200 nm. C, H and N micro- be close to 1.40 Å and 120°, respectively in order to impose a tetra-
analyses were carried out with a Perkin-Elmer 2400 II elemental ana- hedral geometry. The crystallographic data for the complexes are
lyzer. Electrochemical measurements were performed at 20 °C on a summarized in Table 1.
VersaStat-PotentioStat II cyclic voltammeter at different scan rates
varying from 20 to 100 mV s−1 using HPLC grade acetonitrile as solvent 2.4. Theoretical methods
and tetrabutylammonium perchlorate as supporting electrolyte.
Ferrocene was used as internal reference. Platinum and saturated ca- The calculations of the noncovalent interactions were carried with
lomel electrodes (SCE) were the working and the reference electrodes in TURBOMOLE software (version 7.0) [41] using the M06-2X/def2-TZVP
the process, respectively. EPR spectra were recorded from 0 to 1000 mT level of theory. To evaluate the interactions in the solid state, we have
in the temperature range 77–298 K with an X-band (9.15 GHz) Varian used the crystallographic coordinates where only the position of the H-
E−9 spectrometer. The EPR parameters reported in the text were ob- atoms and perchlorate oxygen atoms were optimized. This procedure
tained by simulating the spectra with the computer program Bruker and level of theory have been successfully used to evaluate similar in-
WinEPR SimFonia [35]. The magnetic susceptibility measurements teractions [42–44]. The interaction energies were computed by calcu-
were carried out in the temperature range 2–300 K with an applied lating the difference between the energies of isolated monomers and
magnetic field of 0.5 T on a polycrystalline sample of 1 (with a mass of their assembly. The interaction energies were corrected for the Basis Set
18.74 mg) with a Quantum Design MPMS-XL-5 SQUID susceptometer. Superposition Error (BSSE) using the counterpoise method [45]. Bader’s
The isothermal magnetization was performed on the same sample at 2 K “Atoms in molecules” theory has been used to study the interactions
with magnetic fields up to 5 T. The susceptibility data were corrected discussed herein by means of the AIMall calculation package [46]. The
for the sample holder previously measured using the same conditions molecular electrostatic potential surface (MEPS) calculations have been
and for the diamagnetic contributions of the salt as deduced by using performed by means of the SPARTAN software [47].
Pascaĺs constant tables (χdia = –180.33 × 10−6 emu mol−1) [36].
3. Results and discussion
2.2. Synthesis
3.1. Crystal structure of the complex 1
2.2.1. Synthesis of N′-(1-(pyridin-2-yl)ethylidene)acetohydrazide (HL)
The ligand [HL] was prepared by the condensation of acetic hy- Single crystal X-ray diffraction study reveals that the complex {[Cu
drazide (0.74 g, 10 mmol) with 2-acetylpyridine (1.12 mL, 10 mmol) in (H0.5L)(µ1,3-SCN)]0.5ClO4·0.5MeOH}n (1) presents a one dimensional

96
D. Sadhukhan et al. Inorganica Chimica Acta 484 (2019) 95–103

Scheme 1. Synthesis of the hydrazone ligand, its tautomeric equilibrium and the synthesis of the complex 1.

Table 1
Crystal structure parameters of 1.
Empirical formula C10.5H12.5Cl0.5CuN4O3.5S

Formula weight 364.08


Crystal system Orthorhombic
Space group Pnma
a (Å) 15.9478(8)
b (Å) 23.3266(12)
c (Å) 7.6380(3)
α = β = γ (deg) 90
V (Å3) 2841.4(2)
Z 8
Dcalc (Mg m−3) 1.70
µ (mm−1) 1.794
F(0 0 0) 1484
θ range (deg) 3–26
Total data 9742
Obs data [I > 3σ(I)] 2179
R1[I > 3σ(I)] 0.0585
wR2[I > 3σ(I)] 0.0606
GoF 1.1395
Rint 0.048
Residuals (e.Å−3) 2.34, −1.55

polymeric structure where the metal-ligand assembly is propagated by


single end-to-end SCN− bridges. Fig. 1. Perspective view of the asymmetric/repeating unit of the complex 1.
The asymmetric unit of 1 is shown in Fig. 1 and the related bond
lengths and angles are listed in Table 2. The sum formula suggests that a also supports that the ligand shows different valence states such as
ClO4− unit belongs to the coordination spheres of two symmetry neutral-ketonic state [48] as well as monoanionic enolic state [49].
equivalent Cu(II) (symmetry code: 1.5−x, 1−y, 0.5 + z). The anionic Cu1 is penta-coordinated (avoiding the very long axial coordination
contribution from one bridging SCN− ligand and 0.5 ClO4− counter ion by ClO4−) with a distorted square pyramidal geometry with Addison
to each asymmetric unit suggests that the ligand is in equilibrium be- parameter 0.22 [50]. The tridentate hydrazone ligand is essentially
tween its neutral-ketonic (HL) and anionic-enolic (L−) form to coun- planar and coordinates three equatorial positions around the central
terbalance the charge of the complex as depicted in Scheme 1. Previous metal with formation of two five membered chelate rings via the amido
reports of end-to-end thiocyanato bridged polymeric complexes of HL oxygen (O1), pyridyl (N1) and imine (N2) nitrogen atoms with

97
D. Sadhukhan et al. Inorganica Chimica Acta 484 (2019) 95–103

Table 2 O (amide) stretching vibration is observed at 1678 cm−1. The NeH


Selected bond distances (Å) and bond angles (deg) in 1. (amide) bond vibrations are observed as an asymmetric stretching band
Atoms Distance (Å) Atoms Distance (Å) at 3194 cm−1 and a symmetric one at 3080 cm−1. The four adjacent
pyridine ring hydrogens give rise to wagging vibration bands in the
region 1469–1342 and at 783 cm−1 [53]. The in plane and out of plane
i
Cu1 -N1 2.017(4) Cu1-O1 1.989(4)
Cu1-N2 1.928(4) Cu1-S1 2.793(2)
CeC bending vibrations appear respectively at 624 and 411 cm−1 [54].
Cu1-N4 1.906(5)
The νC=N and the νC=O signals in the complex shift to lower fre-
Atoms Angle (°) Atoms Angle (°) quencies: 1525 cm−1 and 1600 cm−1, respectively. The OeH stretching
N1-Cu1-N2 80.03(17) N2-Cu1-O1 80.14(16)
N1-Cu1-N4 99.42(19) N2-Cu1-N4 172.8(2)
vibration of methanol appears around 2800 cm−1. The NeH bands are
N1-Cu1-O1 159.77(16) N4-Cu1-O1 99.76(18) observed as wide multiplets in the region 3100–2900 cm−1 possibly
N1-Cu1-S1ii 90.73(12) N4-Cu1-S1 99.74(17) due to the involvement of that moiety in hydrogen bonding and equi-
N2-Cu1-S1 89.47(12) O1-Cu1-S1 93.02(13) librium between HL and L−. 1 shows characteristic sharp νSCN bands at
2100 cm−1. Finally, 1 presents the characteristic multiplet bands in the
(i) 1.5−x, 1−y, 0.5 + z; (ii) 1.5−x, 1−y, −0.5 + z.
region 1110–1030 cm−1 corresponding to the ClO4− anions [55]. All
these characteristic vibrational bands in the complex give important
N1eCu1eN2 and O1eCu1eN2 bite angles 80.01° and 80.15°, respec-
structural information of the ligand in 1 in full agreement with its
tively. The fourth equatorial site is occupied by the nitrogen end (N4) of
crystal structure.
the µ1,3-bridging SCN− ligand. The bulkier sulphur end (S1) of a second
SCN− ligand occupies the axial position. The axial Cu1eS1 bond
(2.7935(15) Å) is noticeably longer than the equatorial ones 3.3. Cyclic voltammetry
(1.906(5)–2.017(4) Å) even after considering the large covalent radius
of S, in consistency with other single thiocyanato bridged Cu(II) com- Cyclic voltammetry was performed with reference to SCE using
plexes [51]. The cis angles (80.03(17)–99.76(18)°) vary widely from the ferrocene as internal standard. The representative cyclic voltammogram
ideal 90° value and also the trans angles (159.77(16) and 172.8(2)°) of HL and 1 are shown in Fig. 4. For HL, during the cathodic scan two
deviate significantly from 180° due to the steric constraints of the consecutive reduction peaks are observed at Ep,c = −0.590 and
chelating ligand. The Cu⋯Cu separation along the chain is 6.077 Å −0.907 V, possibly due to the reduction of the carbonyl and the azo-
which is in good agreement with previous reports [52]. methine groups, respectively. The oxidation peak of azomethine moiety
The end-to-end (µ1,3-) bridging thiocyanato ligands are out of the appears at −0.654 V in the anodic scan whereas the carbonyl reduction
equatorial plane and interconnect adjacent metal-ligand assemblies at is irreversible. At more positive potential (0.909 V) another peak is
alternating basal-apical positions into a one-dimensional array along observed, possibly due to the irreversible oxidation of the pyridine ring.
the c axis (Fig. 2). The SCN moiety is almost linear with a S1eC10eN4 The peak positions are affected upon metal coordination as observed in
bond angle of 177.4(5)°, similar to other previously reported single, the voltammogram of 1. The ligand-based reductions occur at −0.572
end-to-end thiocyanato bridged polymeric complexes [51]. and −0.748 V for carbonyl and azomethine reductions, respectively,
The packing diagram (Fig. 3) suggests that a mirror plane passes whereas in the reverse sweep the quasi-reversible azomehine oxidation
through the C11eO5 bond of the CH3OH molecule and renders a and irreversible pyridine ring oxidation occur at −0.520 and 0.375 V,
symmetrical distribution of H3 and H52 over two positions (with respectively. During the cathodic scan two consecutive reductive re-
symmetry code: x, 0.5−y, z), with 0.5 site occupancy. The Cl atom also sponses associated to the metal ions appear at Ep,c = 0.143 and
falls on the same plane which imposes the symmetry on O2. Adjacent −0.367 V for Cu2+ → Cu+ and Cu+ → Cu0 processes, respectively. In
1D chains running along the c-axis are H-bonded via amide nitrogen the anodic scan a very sharp signal, corresponding to the Cu0 →
(N3) donor to the methanol oxygen (O5) acceptor with the parameters Cu2+ + 2e− process, appears at Ep,a = −0.237 V followed by a weak
d(N3eH31) = 0.83 Å; d(H31⋯O5) = 1.93 Å; d(N3⋯O5) = 2.709(8) signal at Ep,a = −0.017 V, corresponding to the Cu0 → Cu+ + e− pro-
Å; < (N3eH31eO5) = 157°. cess [56].

3.4. EPR spectroscopy


3.2. Fourier transform IR spectroscopy
EPR spectra of polycrystalline sample of 1 were recorded at 77 and
The solid state FT-IR spectrum of the hydrazone Schiff base ligand 298 K (Fig. 5). The spectra are axial, with gz ≫ gx ∼ gy > 2.0023, but
shows a sharp band for C]N (imine) stretching at 1578 cm−1. The C] the resonances are not completely resolved at 298 K (trace a of Fig. 5)

Fig. 2. One-dimensional polymer of 1 propagated along the crystallographic c axis.

98
D. Sadhukhan et al. Inorganica Chimica Acta 484 (2019) 95–103

Fig. 3. Molecular packing of complex 1 along the ab plane.

due to the intercenter exchange and dipolar interactions between the structure of 1 in the solid-state breaks down in coordinating solvents,
paramagnetic ions of the lattice [57]. The resolution improves sig- where mononuclear complexes are formed, as observed for other
nificantly lowering the temperature. At 77 K the linewidth reduces polynuclear Cu(II) species [60-70] and, in particular, for polymeric
considerably and two g values were measured (gz = 2.254 and gx = gy NCS-bridged Cu(II) compounds [60,71]. The fact that the A|| values
2.088) with a weak resonance, not detected at RT, being observed at ca. measured in DMF, DMSO or CH3OH are comparable suggests that the
220 mT (Fig. 5b). The order of the g values indicates a ground state same species is formed in organic solution. Such a species can be ex-
based on the Cu-dx2−y2 orbital and a geometry with an axial symmetry plained postulating the break of the thiocyanate bridge and the for-
such as an elongated square pyramid [58,59]. mation of a square pyramid with a long axial bond formed by per-
EPR spectra of 1 dissolved in DMF, DMSO or CH3OH are char- chlorate or, more probably, by a solvent molecule. Peisach and
acteristic of mononuclear species with the unpaired electron in the Cu- Blumberg found that for a species with N3O equatorial coordination
dx2−y2 orbital. They are reported in Fig. 6, while the spectral parameters and + 1 total charge values of gz in the range 2.26–2.28 and Az in the
are collected in Table 3 and are similar. This means that the polymeric range (160–180) × 10−4 cm−1 are expected, while gz lower and Az

Fig. 4. Cyclic voltammogram of HL (in red) and 1 (in black). (For interpretation of the references to colour in this figure legend, the reader is referred to the web
version of this article.)

99
D. Sadhukhan et al. Inorganica Chimica Acta 484 (2019) 95–103

Fig. 7. Thermal variation of the χmT product per Cu(II) ion for 1 (empty cir-
cles). Solid lines are the best fit to the dimer (red) and regular S = 1/2 chain
models with (purple) and without (blue) inter-chain interactions. (For inter-
Fig. 5. X-band EPR spectra of the polycrystalline complex 1 at (a) 298 K and (b) pretation of the references to colour in this figure legend, the reader is referred
77 K. Diphenylpicrylhydrazyl (dpph) is the standard field marker (g = 2.0036). to the web version of this article.)

a room temperature value of ca. 0.38 emu K mol−1. When the tem-
perature is lowered, χmT remains constant down to ca. 10 K. Below this
temperature χmT shows an abrupt decrease to reach a value of ca.
0.13 emu K mol−1 at 2 K for 1 (Fig. 7). This behaviour is typical of a
S = 1/2 centre with a very weak antiferromagnetic interaction.
The crystal structure of 1 shows the presence of two different ex-
change pathways to explain this very weak antiferromagnetic interac-
tion. On one hand, there is a semi-coordinative SeCu bond (with a Cu-S
distance of 2.793(2) Å) established between the Cu(II) centres and the S
atom of a SCN− ligand of a neighbouring unit (Fig. 8a) and, on the
other, there is a S⋯S interaction between neighbouring units with a
S⋯S distance of 3.51 Å (Fig. 8b), shorter than the sum of the van der
Waals radii (3.60 Å, see the theoretical study section). The semi-co-
ordinative CueS bond generates a regular eCueSCNeCue chain
(Fig. 8a), whereas the S⋯S interaction gives rise to isolated Cu-S⋯S-Cu
dimers (Fig. 8b).
Of course, a priori, we cannot discard any of these interactions and,
accordingly, we have used both models to fit the magnetic data of 1 (a
regular S = 1/2 chain and a simple S = 1/2 dimer). Although both
models are very similar and reproduce the general shape of the mag-
netic plot, they fail in the low temperature region (see red and blue
Fig. 6. Anisotropic X-band EPR spectra recorded on the complex 1 dissolved in: lines in Fig. 7). These results point to the simultaneous existence of both
(a) DMF, (b) DMSO and (c) simulated spectrum in DMSO. interactions and, accordingly, we have used two possible models: a
Diphenylpicrylhydrazyl (dpph) is the standard field marker (g = 2.0036). regular chain with inter-chain interactions and a regular dimer with
inter-dimer interactions. The inter-chain and inter-dimer interactions
Table 3 are reproduced with the molecular field approximation [73]. Although
EPR parameters of 1 in several organic solvents. these two models significantly improve the magnetic fit, the best fit is
obtained with the first one: a regular S = 1/2 chain with inter-chain
Solvent gz Az (10−4 cm−1) gx = gy Ax = Ay (10−4 cm−1)
interactions. This model reproduces very satisfactorily the magnetic
DMSO 2.286 149 2.061 14 properties of 1 in the whole temperature range with g = 2.005,
DMF 2.287 148 2.061 14 J = −5.6 cm−1 and zJ′ = −1.8 cm−1, where z is the number of
CH3OH 2.287 149 2.062 15 neighbouring interacting chains (purple lines in Fig. 7). The exchange
coupling found in compound 1 agrees with the expected ones since both
exchange pathways (the long CueS bond and the intermolecular S⋯S
significantly larger are predicted for a neutral species [72]. Therefore, it
interaction) are expected to give rise to very weak antiferromagnetic
is plausible that the stoichiometry of the mononuclear species is the
couplings since they are connecting apical/apical or basal/apical po-
monocationic [Cu(HL)(NCS)(Solvent)], with Solvent = DMF, DMSO or
sitions, precluding an effective overlap of the dx2–y2 SOMO orbitals.
CH3OH axially bound to Cu. The weak axial coordination of the solvent
would account also for the reduction of Az (∼150 × 10−4 cm−1, see
Table 3) compared to the value expected for a square planar structure.
3.6. Theoretical study

3.5. Magnetic study Firstly, we have analysed the existence of π-acidic regions in the π-
system of the Schiff base ligand using the Molecular Electrostatic
The thermal variations of the χmT product per Cu(II) ion for 1 shows Potential (MEP) mapped onto the van der Waals surface (using a square

100
D. Sadhukhan et al. Inorganica Chimica Acta 484 (2019) 95–103

Fig. 8. Two possible spin interaction pathways: (a) a regular Cu-SCN-Cu chain and (b) a discrete Cu-S⋯S-Cu dimer.

(bluish isocontour, Fig. 9). More interestingly, the electrostatic poten-


tial over the C atom of the imidic C]N bond is even more positive than
the potential at the metal centre, thus highlighting its ability to interact
with electron rich atoms. The most positive potential regions are lo-
cated in the molecular plane along the NeH bonds (∼80 kcal/mol).
In 1 the chains are interconnected by supramolecular forces in
several directions yielding to the final 3D solid state architecture. In one
of these directions (see Fig. 10a) the S atom of the SCN− ligand is si-
tuated over the C atom of the imidic double bond (π-hole interaction),
in agreement with the MEP surface shown in Fig. 9. Remarkably, the
S⋯C distance is shorter than the sum of van der Waals radii, thus
emphasizing the importance of the π-hole interaction. We have com-
puted the interaction energy of a finite model of the chains (see
Fig. 10b) that is very large in absolute value (ΔE1 = −62.7 kcal/mol)
due to strong electrostatic effects (see MEP surface in Fig. 9) and the
contribution of additional H-bonding interactions between the per-
chlorate anion and the H-atoms of the methyl group (see blue dashed
lines in Fig. 10b) along with other long range van der Waals interac-
Fig. 9. MEP open surface of the Schiff base ligand coordinated to the Cu(II) tions. Moreover, there is also a chalcogen bonding interaction between
centre using two SCN− counter-ions. The energies at some points of the surface the S atoms of the pseudohalide (S⋯S distance also shorter than the
are indicated.
sum of van der Waals radius). We have used the Bader’s theory of
“atoms in molecules” (AIM) to characterize the π-hole interactions in-
pyramidal geometry and two SCN− ligands to counter-balance the volving the Schiff base ligand described above for 1. A bond critical
metal charge). The MEP surface shows that in the Schiff base ligand point (CP) and a bond path connecting two atoms is an unambiguous
coordinated to Cu(II) the MEP values over the π-systems are positive evidence of interaction [74]. A partial view of the AIM distribution of

Fig. 10. (a) Fragment of the X-ray structure of 1. (b) Theoretical model used to estimate the π-hole interaction. Distances in Å. (c) Distribution of bond and ring
critical points (red and yellow spheres, respectively) and bond paths for the model of compound 1. (For interpretation of the references to colour in this figure legend,
the reader is referred to the web version of this article.)

101
D. Sadhukhan et al. Inorganica Chimica Acta 484 (2019) 95–103

critical points and bond paths computed for 1 is shown in Fig. 10c. It [18] Y. Li, Z. Yang, M. Zhou, Y. Li, J. He, X. Wang, Z. Lin, RSC. Adv. 7 (2017)
reveals that each S⋯π–hole interaction is characterized by a bond CPs 41527–41539.
[19] S. Banerjee, A. Ray, S. Sen, S. Mitra, D.L. Hughes, R.J. Butcher, S.R. Batten, Inorg.
(red sphere) and a bond path connecting the S atom of the pseuohalide Chim. Acta 361 (2008) 2692–2700.
ligand anion to the imidic C atom, thus confirming the existence of the [20] S. Banerjee, S. Sen, S. Basak, S. Mitra, D.L. Hughes, C. Desplanches, Inorg. Chim.
interaction. The S⋯S interaction is also characterized by a bond CPs Acta 361 (2008) 2707–2714.
[21] B. Samanta, J. Chakraborty, S. Shit, S.R. Batten, P. Jensen, J.D. Masuda, S. Mitra,
(red sphere) and a bond path interconnecting the S atoms. Inorg. Chim. Acta. 360 (2007) 2471–2484.
[22] M. Sutradhar, A.J.L. Pombeiro, Coord. Chem. Rev. 265 (2014) 89–124.
4. Conclusion [23] M. Sutradhar, E.C.B.A. Alegria, T. Roy Barman, F. Scorcelletti, M. Fátima, C. Guedes
da Silva, A.J.L. Pombeiro, Mole. Catal. 439 (2017) 224–232.
[24] M. Sutradhar, E.C.B.A. Alegria, K.T. Mahmudov, M. Fátima, C. Guedes da Silva,
We have synthesized a thiocyanato bridged copper complex of a A.J.L. Pombeiro, RSC Adv. 6 (2016) 8079–8088.
pyridine based acetohydrazide ligands. As ClO4− anion is shared by the [25] K.T. Mahmudov, M. Fátima, C. Guedes da Silva, M. Sutradhar, M.N. Kopylovich,
F.E. Huseynov, N.T. Shamilov, A.A. Voronina, T.M. Buslaeva, A.J.L. Pombeiro,
coordination sphere of two Cu(II) ion, the ligand is in equilibrium be-
Dalton Trans. 44 (2015) 5602–5610.
tween its neutral ketonic form and anionic enolic form. The Cu(II) [26] Non-Covalent Interactions in the Synthesis and Design of New Compounds. John
complex is weakly antiferromagnetically coupled via the SCN− bridges Wiley & Sons, Inc., Hoboken, NJ, 2016.
which are not efficient exchange pathways due to their out of plane [27] E. Viñuelas-Zahínos, F. Luna-Giles, P. Torres-García, A. Bernalte-García, Polyhedron
28 (2009) 1362–1368.
basal-apical coordination. The solution state composition of the com- [28] N. Filipović, H. Borrmann, T. Todorović, M. Borna, V. Spasojević, D. Sladić,
plex as predicted by EPR spectroscopy is essentially a mononuclear I. Novaković, K. Andjelković, Inorg. Chim. Acta 362 (2009) 1996–2000.
cationic species coordinated by solvent [Cu(HL)(NCS)(Solvent)]. We [29] D. Sadhukhan, A. Ray, G. Pilet, C. Rizzoli, G.M. Rosair, C.J. Gomez García,
S. Signorella, S. Bellu, S. Mitra, Inorg. Chem. 50 (2011) 8326–8339.
have also described the existence of unconventional S⋯π-hole inter- [30] A. Ray, C. Rizzoli, G. Pilet, C. Desplanches, E. Garribba, E. Rentschler, S. Mitra, Eur.
actions in the solid state packing of this compound. The energetic J. Inorg. Chem. 20 (2009) 2915–2928.
features of several interactions have been studied by means of DFT [31] A. Datta, K. Das, Y.-M. Jhou, J.-H. Huang, H.M. Lee, Acta Crystallogr., Sect. E 66
(2010) m1271.
calculations and characterized using the Bader’s theory of atoms in [32] A. Datta, B. Jui-Hsien Huang, J. Machura, Chem. Cryst. 42 (2012) 691–696.
molecules. Finally, the acidity of the imidic bond in the Schiff base [33] A. Datta, K. Das, Y.-M. Jhou, J.-H. Huang, H.M. Lee, Acta Crystallogr., Sect. E 67
ligand coordinated to Cu(II) has been also demonstrated using MEP (2011) m123.
[34] K. Das, C. Sinha, A. Datta, E. Garribba, M. Fondo, F.A. Mautner, R.C. Fischer, J.
surface analysis. This interaction might be rationally exploited in Chem. Res. 36 (2012) 722–725.
crystal engineering and supramolecular chemistry. [35] WINEPR SimFonia, version 1.25, Bruker Analytische Messtechnik GmbH, Karlsruhe,
(1996).
[36] G.A. Bain, J.F. Berry, J. Chem. Educ. 85 (2008) 532–536.
Acknowledgements
[37] G.M. Sheldrick, SHELXS-97, SHELXL-97, Programs for Crystal Structure Analysis,
University of Göttingen, Germany, 1997.
DS is thankful to Prof. Rodolphe Clérac and Centre de Recherche [38] A. Altomare, G. Cascarano, C. Giacovazzo, A. Guagliardi, M.C. Burla, G. Polidori,
Paul Pascal for the X-ray facilities. The authors thank the MINECO of M. Camalli, J. Appl. Cryst. 27 (1994) 435–436.
[39] P.W. Betteridge, J.R. Carruthers, R.I. Cooper, K. Prout, D.J. Watkin, J. Appl. Cryst.
Spain (projects CTQ2014-57393-C2-1-P and CTQ2017-87201-P, FEDER 36 (2003) 1487.
funds), the Generalitat Valencia (PrometeoII/2014/076) and FFABR [40] J. De Meulenaer, H. Tompa, Acta Crystallographica 19 (1965) 1014–1018.
2017 “Fondo per il finanziamento delle attività base di ricerca” for fi- [41] R. Ahlrichs, M. Bär, M. Häser, H. Horn, C. Kölmel, Chem. Phys. Lett. 162 (1989)
165–169.
nancial support. We also thank the CTI (UIB) for computational facil- [42] D. Sadhukhan, M. Maiti, G. Pilet, A. Bauzá, A. Frontera, S. Mitra, Eur. J. Inorg.
ities. Chem. 11 (2015) 1958–1972.
[43] M. Mirzaei, H. Eshtiagh-Hosseini, Z. Bolouri, Z. Rahmati, A. Esmaeilzadeh,
A. Hassanpoor, A. Bauza, P. Ballester, M. Barceló-Oliver, J.T. Mague, B. Notash,
Appendix A. Supplementary data A. Frontera, Cryst. Growth Des. 15 (2015) 1351–1361.
[44] P. Chakraborty, S. Purkait, S. Mondal, A. Bauzá, A. Frontera, C. Massera, D. Das,
Supplementary data to this article can be found online at https:// CrystEngComm. 17 (2015) 4680–4690.
[45] S.F. Boys, F. Bernardi, Mol. Phys. 19 (1970) 553–566.
doi.org/10.1016/j.ica.2018.09.031.
[46] AIMAll (Version 13.05.06), Todd A. Keith, TK Gristmill Software, Overland Park KS,
USA, 2013.
References [47] Spartan ’10, v. 1.1.0, Wavefunction Inc, Irvine CA. www.wavefun.com.
[48] A. Datta, P.-H. Liu, J.-H. Huang, E. Garribba, M. Turnbull, B. Machura, C.-L. Hsu,
W.-T. Chang, A. Pevec, Polyhedron 44 (2012) 77–87.
[1] D.K. Kolmel, E.T. Kool, Oximes and hydrazones in bioconjugation: mechanism and [49] K. Das, A. Datta, C. Sinha, J.-H. Huang, E. Garribba, C.-S. Hsiao, C.-L. Hsu, Chem.
catalysis, Chem. Rev. 117 (2017) 10358–10376. Open 1 (2012) 80–89.
[2] L.A. Tatum, X. Su, I. Aprahamian, Simple hydrazone building blocks for compli- [50] A.W. Addison, T.N. Rao, J. Reedijk, J.V. Rijn, G.C. Verschoor, J. Chem. Soc., Dalton
cated functional materials, Acc. Chem. Res. 47 (2014) 2141–2149. Trans. (1984) 1349–1356.
[3] D. Sadhukhan, A. Ray, G. Pilet, G.M. Rosair, E. Garribba, S. Mitra, Bull. Chem. Soc. [51] J. Carranza, J. Sletten, F. Lloret, M. Julve, Polyhedron 28 (2009) 2249–2257.
Jpn. 8 (2011) 764–777. [52] P. Talukder, A. Datta, S. Mitra, G. Rosair, M.S. El Fallah, J. Ribas, Dalton Trans.
[4] T. Clark, WIREs Comput. Mol. Sci. 3 (2013) 13–20. (2004) 4161–4167.
[5] P. Politzer, J.S. Murray, T. Clark, Phys. Chem. Chem. Phys. 15 (2013) [53] R.A. Krause, N.B. Colthup, D.H. Busch, J. Phys. Chem. 65 (1961) 2216–2219.
11178–11189. [54] C.W. Frank, L.B. Rogers, Inorg. Chem. 5 (1966) 615–622.
[6] P. Politzer, J.S. Murray, P. Lane, Int. J. Quantum Chem. 107 (2007) 3046–3052. [55] K. Nakamoto, Infrared and Raman Spectra of Inorganic and Coordination
[7] A. Bauza, T.J. Mooibroek, A. Frontera, Chem. Commun. 51 (2015) 1491–1493. Compounds: Theory and Applications in Inorganic Chemistry, Part A 5th ed., John
[8] A. Bauza, T.J. Mooibroek, A. Frontera, Chem. Phys. Chem. 16 (2015) 2496–2517. Willey and Sons Inc., New York, 1997, p. 83.
[9] A. Bauza, T.J. Mooibroek, A. Frontera, Chem. Rec. 16 (2016) 473–487. [56] J. Losada, I. del Peso, L. Beyer, Inorg. Chim. Acta 321 (2001) 107–115.
[10] A. Ray, S. Banerjee, R.J. Butcher, C. Desplanches, S. Mitra, Polyhedron 27 (2008) [57] A.C. Rizzi, N.I. Neuman, P.J. González, C.D. Brondino, Eur. J. Inorg. Chem. (2016)
2409–2415. 192–207.
[11] S. Sen, C.R. Choudhury, P. Talukder, S. Mitra, M. Westerhausen, A.N. Kneifel, [58] B.J. Hathaway, D.E. Billing, Coord. Chem. Rev. 5 (1970) 143–207.
C. Desplanches, N. Daro, J.P. Sutter, Polyhedron 25 (2006) 1271–1490. [59] E. Garribba, G. Micera, J. Chem. Ed. 83 (2006) 1229–1232.
[12] S. Pal, S. Pal, Eur. J. Inorg. Chem. 23 (2003) 4244–4252. [60] M.A. Ali, A.H. Mirza, R.J. Fereday, R.J. Butcher, J.M. Fuller, S.C. Drew, L.R. Gahan,
[13] S.C. Chan, L.L. Koh, P.-H. Leung, J.D. Ranford, K.Y. Sim, Inorg. Chim. Acta 236 G.R. Hanson, B. Moubaraki, K.S. Murray, Inorg. Chim. Acta. 358 (2005)
(1995) 101–108. 3937–3948.
[14] E.W. Ainscough, A.M. Brodie, A.J. Dobbs, J.D. Ranford, J.M. Waters, Inorg. Chim. [61] G.A. van Albada, M.E. Quiroz-Castro, I. Mutikainen, U. Turpeinen, J. Reedijk, Inorg.
Acta 267 (1998) 27–38. Chim. Acta 298 (2000) 221–225.
[15] S. Banerjee, S. Mondal, S. Sen, S. Das, D.L. Hughes, C. Rizzoli, C. Desplanches, [62] W.A. Alves, R.H. de Almeida Santos, A. Paduan-Filho, C.C. Becerra, A.C. Borin,
C. Mandal, S. Mitra, Dalton Trans. 34 (2009) 6849–6860. A.M. Da Costa Ferriera, Inorg. Chim. Acta 357 (2004) 2269–2278.
[16] S. Banerjee, S. Sen, J. Chakraborty, R.J. Butcher, C.J. Gómez-García, R. Puchta, [63] A. Koval, M. Sgobba, M. Huisman, M. Lüken, E. Saint-Aman, P. Gamez, B. Krebs,
S. Mitra, Aust. J. Chem. 62 (2009) 1614–1621. J. Reedijk, Inorg. Chim. Acta 359 (2006) 4071–4078.
[17] J. Chakraborty, S. Thakurta, G. Pilet, D. Luneau, S. Mitra, Polyhedron 28 (2009) [64] S. Thakurta, J. Chakraborty, G. Rosair, J. Tercero, M.S. El Fallah, E. Garribba,
819–825. S. Mitra, Inorg. Chem. 47 (2008) 6227–6235.

102
D. Sadhukhan et al. Inorganica Chimica Acta 484 (2019) 95–103

[65] S. Thakurta, P. Roy, G. Rosair, C.J. Gómez-García, E. Garribba, S. Mitra, Polyhedron Polyhedron 52 (2013) 963–969.
28 (2009) 695–702. [70] S. Saha, A. Sasmal, C.R. Choudhury, C.J. Gomez-Garcia, E. Garribba, S. Mitra,
[66] C.P. Pradeep, S.K. Das, Polyhedron 28 (2009) 630–636. Polyhedron 69 (2014) 262–269.
[67] S. Thakurta, C. Rizzoli, R.J. Butcher, C.J. Gómez-García, E. Garribba, S. Mitra, [71] G.A. Bowmaker, C. Di Nicola, F. Marchetti, C. Pettinari, B.W. Skelton, N. Somers,
Inorg. Chim. Acta 363 (2010) 1395–1403. A.H. White, Inorg. Chim. Acta 375 (2011) 31–40.
[68] A. Ray, S. Mitra, A. Dehno Khalaji, C. Atmani, N. Cosquer, S. Triki, J.M. Clemente- [72] J. Peisach, W.E. Blumberg, Arch. Biochem. Biophys. 165 (1974) 691–708.
Juan, S. Cardona-Serra, C.J. Gómez-García, R.J. Butcher, E. Garribba, D. Xu, Inorg. [73] C.J. O’Connor, Prog. Inorg. Chem. 29 (1982) 203–283.
Chim. Acta 363 (2010) 3580–3588. [74] R.F.W. Bader, Atoms in Molecules – A Quantum Theory, Oxford University Press,
[69] S. Shit, M. Nandy, G. Rosair, M. Salah El Fallah, J. Ribas, E. Garribba, S. Mitra, Oxford, U.K., 1990.

103

You might also like