Download as pdf or txt
Download as pdf or txt
You are on page 1of 18

Anion-directed structural diversification in four new Cd(II)complexes of a flexible polyether-based

dicarboxylic acid†

Sakshi and Sadhika Khullar*


Department of Chemistry, Dr B R Ambedkar National Institute of Technology, GT Road bypass
Jalandhar, Punjab-144011, INDIA
E-mail: khullars@nitj.ac.in

ABSTRACT

In this work, we report the chemistry of an unexplored flexible polyether based dicarboxylic acid (3,6,9-
trioxaundecanedioic acid, H2toua) with cadmium salts containing different counter anions. With CdCl 2 or
Cd(ClO4)2 in a 1:1 ratio, it remarkably forms a discrete neutral compound [Cd(H2toua)Cl2].H2O (1)and a
polymeric ionic compound, {[Cd(Htoua)(H2O)](ClO4)}n(2), respectively, as determined by the single
crystal X-ray diffraction. In both cases, Cd(II) is heptacoordinate withall basal donor atoms are from the
ligand (three ether oxygen and two oxygen atoms from the fully intact or mono-deprotonated carboxylic
acid groups, respectively) but with a major difference in axial (trans to each other) coordinated atom types
- two chlorines in 1 while one water molecule and a bridging carboxylate oxygen in 2.Unlike 1, the
carboxylate group of Htoua bridges between the Cd(II) centers for the formation of a 1D polymeric chain
in 2. Furthermore, Hirshfeld surface analysis of the various intermolecular interactions enables
quantifiable contributions to crystal packing, revealing the key differences in the interactions
encountered by 1 and 2. Using Cd(OAc)2 as the metal salt, it is possible to make [Cd(toua)].H2O (3),due
to the formation of easily removable volatile by-productAcOH. On the other hand, a nitrate analogue of 2,
i.e., {[Cd(Htoua)(H2O)](NO3)}n(4) is isolated if Cd(NO3)2 is used. All these compounds were
characterized by a number of spectroscopic and microscopic methods. Their thermal behaviour and
luminescence properties were also determined.
Keywords: 3,6,9-trioxaundecanedioic acid, Hirshfeld analysis, Coordination polymer, Cd(II)
complexes.

1. Introduction
Both discrete and polymeric coordination compounds are the backbone of new developments in inorganic
chemistry for their intriguing structural features and for diverse applications in catalysis, luminescence,
optics, biomedical, electronic and magnetic devices [1-5]. Furthermore, in recent years their use in energy
and environment sectors has boosted the exploration of designing and developing new materials
comprised of various metal ions and newly-designed or readily available organic ligands [6,7]. Depending
on the desired applications, the structural features and physico-chemical properties of such discrete and
polymeric compounds are custom-made and tuned for a judicious combination of metal ions and the
organic ligands. On the other hand, the role of supramolecular interactions (hydrogen bonding, π-π
interactions, etc.) that take the discrete molecules into supramolecular coordination complexes, and 1D
and 2D coordination polymers into higher dimensional (2D and 3D) supramolecular coordination
architectures are also important. In exploring the possibilities for such compounds, divalent, trivalent and
tetravalent metal ions are primarily combined with suitable ligands with N-/P- and/or O-donor atoms
based on the SHAB principles. Similarly, the synthesis methods that range from room temperature to
reflux to hydro-/solvothermal have been the most common. However, the preference for the room
temperature synthesis has a few advantages that include green chemistry approaches, easy to scale up,
reproducibility, and simplicity.
For the consideration of metal ions in making such compounds, the coordination number and the
geometry around the metal center primarily dictate the outcome. With a vast accessibility and
convenience, the divalent metal ions are explored more than the trivalent and tetravalent metal ions.
Among the divalent metal ions, the coordination chemistry of Cd(II) is unique for its large radius and d 10
electronic configuration. It exhibits coordination numbers from 4 to 8 and can form complexes with
geometries varying from tetrahedral to dodecahedral with an easy occurrence of severe distortion from
ideal polyhedron geometry [8]. For choosing the ligands with N/O donor atoms, carboxylates are the most
explored ones. While a combination of neutral N-donor (amines, pyridyl, pyrazole, imidazole, etc.) and
anionic O-donor atoms (carboxylates, alkoxides, phenoxides, etc.) in separate ligands or in the same
ligands is the majority, the utilization of ligands with both anionic and neutral O-donor atoms is less
common. Additionally, a huge portion of the literature is filled with rigid ligands as the formation of
complexes is more feasible. Thus, if the chosen ligands with N-/O-donor atoms have flexible linkages
between them, these provide some properties unseen in complexes with rigid ligands. It has been
observed that the complexity arises in dealing with flexible ligands. Similarly, the diverse binding
(monodentate, chelated or bridging bidentate) modes of the carboxylate groups also provides structural
diversity in the resultant compounds [9]. Finally, the reaction parameters such as temperature, anion in
the metal salt precursors, pH and solvent play a crucial role in the final product formation. For a proper
understanding of the effect of each of these parameters, researchers have explored with numerous
examples. The selected examples showing the effects of solvent and time [10], the role of ancillary
ligands [11,12], diversity of carboxylate binding [13-15], and the flexible linkages between donor atoms
[16,17], and the effect of anions [18-23] are relevant to this work. The carboxylate ligands with a
heteroatom in the rings (furan dicarboxylic acid and thiophene dicarboxylic acid are the most common)
and with a heteroatom connecting the two or three carboxylates (aliphatic or aromatic) like in 4,4’-
oxydibenzoic acid (OBA) [24-27] have been explored. On the other hand, the use of -O-(CH 2)n-O-CH3 (n
= 2, 3 or 4) linked to the phenyl ring of a dicarboxylate linker for the functionalization of the resultant
CPs has also been demonstrated [28]. To the best of our knowledge, a multitopic carboxylate ligand with
a polyether linkage between the carboxylate groups has not been explored for making new coordination
complexes [29,30].
With the above background, the scope of the present work was defined with an unexplored flexible
polyether-based heteroatom containing dicarboxylic acid, 3,6,9-trioxaundecanedioic acid (H 2toua) as
shown in Fig. 1. It is expected that the hydrophilic nature of H 2touawill enhance the polarity and will
make the coordination complexes more soluble in polar solvents. Furthermore, the presence of ether
linkages will provide greater linkage flexibility. Herein, we report the room temperature synthesis and
structural characterization of four Cd(II) complexes, [Cd(H2toua)Cl2].H2O (1),
{[Cd(Htoua)(H2O)](ClO4)}n(2), [Cd(toua)].H2O (3), and {[Cd(Htoua)(H2O)](NO3)}n(4), where Cd(II)
salts containing different anions, such as Cl -, ClO4-, OAc-,andNO3-, respectively, were used as the starting
materials.Their full characterization by elemental analysis, FTIR and UV-vis spectroscopy, and
electrospray ionization mass spectroscopy (ESI-MS), Field-emission scanning electron microscopy
(FESEM) and X-ray energy dispersive spectroscopy (EDS). Solid state structures of 1 and 2 determined
by single crystal X-ray diffraction indicated that a mere change of counter anion from chloride to
perchlorate provided a discrete compound (1) vs one dimensional coordination polymer (2). Interestingly,
the ether O atoms in the ligand participated in the coordination to the metal center.
Fig. 1. Structure and features of linker (H2toua) used in this study.

2. Experimental

2.1. Materials and instrumentation

All reagents and chemicals were used as received from commercial sources, without any further
purification. All reactions were carried out under aerobic conditions.
Caution! In this study Cadmium (II) perchlorate was used, which is explosive in nature. However, handle
such salts with special care.
Elemental analysis (C, H, N) was carried out using a ThermoFinnigan CHNS and Perkin-Elmer 2400
series II CHNS analysersat IIT Bombay and IISER Kolkata, respectively. High Resolution Mass
Spectrometry (HRMS) data were measured by XEVO G2-XS QTOF instrument for the 50-2000 amu
range with ESI ion source at IIT Ropar. The solid-state UV−Vis spectra were recorded using an Agilent
Technologies Cary-5000 UV-Vis-NIR spectrophotometer. FT-IR spectra (KBr pellet, 400−4000 cm−1)
were measured on a Perkin-Elmer Spectrum I spectrometer. Thermogravimetric analysis was carried out
using a TA Shimadzu instrument. Field emission scanning electron microscopy (FESEM) experiments,
including energy dispersive X-ray spectroscopy (EDX) for mapping of elements, were performed on a
JEOL instrument; each sample was well dispersed in methanol, drop casted in a silicon wafer, dried and
coated with gold using a working distance of 15-22 mm and a voltage of 3-5 kV. A RigakuUltima IV
diffractometer equipped with a 3kW sealed tube (voltage and current were 40 kV and 40 mA,
respectively) for Cu Kα X-ray radiation and a DTex Ultra detector was used to collect and analyze the
powder X-ray diffraction. The data were collected over an angle range of 5° to 70° with a scanning speed
of 1° per minute with 0.01° steps.
For the single crystal X-ray structure determination, crystals of 1 and 2 mounted on a goniometer head
were evaluated and their full data were collected using a Kappa APEX II diffractometer, equipped with a
CCD detector (with the crystal-to-detector distance fixed at 60 mm) and sealed-tube monochromated
MoKα radiation and interfaced to a PC, through the program APEX2 [31]. In each case, three sets of
frames of data were collected with 0.30° steps in ω and an exposure time of 20 s within a randomly
oriented region of reciprocal space surveyed to the extent of 1.3 hemispheres. Using the program SAINT
[31] integration of the data, fit of reflection profiles, and calculation of F 2 and σ(F2) for each reflection
were done. Further correction to the data for the Lorentz and polarization effects was also applied. For the
processing of data that included determination of space group, application of an absorption correction
(SADABS) [31], merging of data, and generation of files necessary for solution and refinement was done
by the subroutine XPREP [31]. Finally, the structure solution and refinement of 1 and 2 were conducted
using SHELX97 [32]. Final crystallographic parameters and basic information pertaining to data
collection and structure refinement for 1 and 2 are summarized in Table 1. All figures were drawn using
Mercury V 3.0 [33], and hydrogen bonding parameters were generated using Platon [34]. The final
positional and thermal parameters of the non-hydrogen atoms for all structures are listed in the CIF files.
The Hirshfeld surfaces (HSs) of compound 1 and 2 have been carried out based on electron distribution of
the molecules and are calculated as the sum of spherical atom electron densities [35,36]. HSs are unique
for the investigating molecule and a set of spherical atomic electron densities. The normalized contact
distance (dnorm) is generated based on de, diand the van der Waals radii of the atom where de and di are
defined as the distance from the point to the nearest nucleus external and internal to the surface,
respectively [37]. The dnorm is calculated by the equation

𝑑 −𝑟 𝑑 −𝑟
𝑑 = +
𝑟 𝑟
The 2D fingerprint plots are generated by using the de and di parameters, which elucidate the
percentage dominance of particular interaction in developing the supramolecular architecture
[38,39]. The HSs mapped with various properties and 2D fingerprint plots are generated by using
the program CrystalExplorer [40].

2.2. Synthesis of [Cd(H2toua)Cl2].H2O (1)

In a 10 mL round bottom flask (RBF), 80.6 mg (0.4 mmol) of CdCl 2.H2O and 89 mg (0.4 mmol) of
H2toua were dissolved in 4 mL methanol. A colorless precipitate appeared immediately and the reaction
mixture was further stirred for 2 h at room temperature. The product was filtered, washed with methanol
and air-dried. Yield: 133 mg (95%). Anal.Calcd. (%) for C8H16CdCl2O8 (MW 423.51): Calc. C, 22.69; H,
3.81. Found: C, 22.33; H, 3.60. Selected FTIR peaks (KBr, cm -1): 3395,2943, 1710,1665, 1454, 1074,
884,658. UV-Visible (λmax, water): 204 nm (Eg 4.8 eV). Its single crystals were obtained from theslow
evaporation of a solution of 1 in MeOH: CH3CN (1:1).

2.3. Synthesis of {[Cd(Htoua)(H2O)](ClO4)}n (2)

In a 10 mL RBF, 94.8 mg (0.288 mmol) of Cd(ClO4)2.H2O and 64 mg (0.288 mmol) H2toua were
dissolved in 3 mL methanol. The clear solution was stirred for 3 h at room temperature. The product was
isolated via evaporation under vacuum. Yield: 64.4 mg (50%). Anal.Calcd. (%) for C 8H15CdClO12 (MW
451.05): Calc. C, 21.30; H, 3.354. Found: C, 20.97; H, 3.42. Selected FTIR peaks (KBr, cm -1): 3514,
2927, 1737, 1621, 1435, 1242, 1080, 623. UV-Visible (λ max, water): 206 nm (Eg = 5.5 eV). Its single
crystals were grown by the slow evaporation of an aqueous solution of 2.

2.4. Synthesis of [Cd(toua)].H2O (3)

In a 10 mL RBF, 85.9 mg (0.254 mmol) of Cd(OAc)2.6H2O and 56.4 mg (0.254 mmol) of H2toua were
dissolved in 2 mL methanol. After stirring for 3 h at room temperature, the solvent was removed via
evaporation under vacuum and the solid was furthertreated with a 1:1 acetonitrile and toluene mixture to
remove the acetic acid by-product. Yield: 92.4 mg (82.6%). Anal.Calcd. (%) for C 8H14CdO8 (MW
350.603): Calc. C, 27.41; H, 4.02. Found: C, 26.17; H, 3.86. Selected FTIR peaks (KBr, cm -1): 3484,
2927, 1582, 1420, 1327, 1111, 902, 678, 608. UV-Visible (λ max, water): 207 nm (Eg = 5.6 eV).

2.5. Synthesis of {[Cd(Htoua)(H2O)](NO3)}n (4)


In a 10 mL RBF, 72.1 mg (0.2336 mmol) of Cd(NO3)2.4H2O and 51.9 mg (0.2336 mmol) H2toua were
dissolved in 3 mL methanol. The clear solution was stirred for 3 h at room temperature. The product was
isolated via evaporation under vacuum. Yield: 60.2 mg (62%). Anal.Calcd. (%) for C 8H15CdNO11 (MW
413.616): Calc. C, 23.23; H, 3.66; N, 3.39. Found: C, 22.72; H, 3.52; N, 3.74. Selected FTIR peaks (KBr,
cm-1): 3381, 3180, 1680, 1586, 1424, 1307, 1242, 1076, 897, 680. UV-Visible (λ max, water): 212 nm (Eg =
5.3 eV).

3. Results and discussion

3.1. Synthesis and spectroscopic characterization

Scheme 1 summarizes the synthesis of compounds 1-4 by employing different Cd(II) salts and H2toua in
a 1:1 ratio in methanol at room temperature. In case of 1, the ligand does not get deprotonated. On
changing from CdCl2 to Cd(ClO4)2, one of the carboxylic acid groups gets deprotonated in 2. Compound
1 is a neutral discrete molecule, whereas compound 2 is an ionic compound where the cation is a 1D
coordination polymer with a bridging carboxylate group of Htoua as confirmed by the single crystal X-
ray diffraction studies. On the other hand, compound 3 is neutral where the ligand is fully deprotonated
due to the formation of acetic acid by-product that is completely removed using an azeotropic mixture of
acetonitrile and toluene in a 1:1 ratio. Compound 4 is a nitrate analog of 2 and thus contains one partially
deprotonated ligand. Compound 4 was made to use it for comparing the CdO nanostructures with those
made from 1 and 2. These compounds were extensively characterized by numerous spectroscopic and
microscopic methods as described below. Single crystal structures of 1 and 2 provided the structural
diversity. Several attempts to obtain a suitable single crystal of 3 were unsuccessful; however, its
structure can be correlated to the Mn analogue reported decades ago [29]. On the other hand, the structure
of 4 is expected to be similar to that of 2 as only the counter anions are different.

Table 1
Crystallographic data and structure refinement parameters for 1 and 2.
Compound 1 2
Chemical formula C9H17CdCl2O7 C8H15CdClO12
Formula weight 423.51 451.05
Temperature (K) 296 296
Wavelength (Å) 0.71073 0.71073
Crystal system monoclinic Orthorhombic
Space group P21/c Cmc21
a/Å 14.307(2) 23.174(6)
b/Å 8.1354(13) 9.411(3)
c/Å 14.609(2) 13.660(4)
α/° 90 90
β/° 117.770(2) 90
γ/° 90 90
Volume (Å3) 1504.5(4) 2979.1(14)
Z 4 8
Density (g/cm3) 1.87 2.011
Absorption coefficient (mm-1) 1.834 1.705
F(000) 840 1792
θ range for data collection 1.61 to 24.90 2.336 to 24.803
Reflections collected 6082 5887
Independent reflections 2594 1948
Reflections with I> 2σ(I) 2414 1860
Rint 0.0161 0.0243
No. of parameters 185 217
Goodness-of-fit on F2 0.987 1.051
R1= 0.0161, wR2= R1= 0.0201, wR2=
Final R1/wR2 (I>=2σ (I))a,b
0.0417 0.0451
R1= 0.0180, wR2= R1= 0.0217, wR2=
R1/wR2 (all data)a,b
0.0432 0.0462
Largest diff. peak/hole / e Å-3 0.289/-0.239 0.333/-0.371
Flack parameter 0.07(5)
a
R1 = Σ||Fo| − |Fc||/Σ|Fo|. bwR2 = [Σw(Fo2 − Fc2)2/Σw(Fo2)2]1/2, where, w = 1/[σ2(Fo2) + (aP)2 +
bP], P = (Fo2 + 2Fc2)/3.

[Cd(H2toua)Cl2].H2O (1)
(Yield: 95%)

CdCl2.H2O
MeOH, RT

Cd(NO3)2.4H2O Cd(ClO4)2.H2O
{[Cd(Htoua)(H2O)](NO3)}n (4) H2toua {[Cd(Htoua)(H2O)](ClO4)}n (2)
MeOH, RT MeOH, RT
(Yield: 62%) (Yield: 50%)

Cd(OAc)2.6H2O
MeOH, RT

[Cd(toua)].H2O (3)
(Yield: 83%)

Scheme 1. Syntheses of 1-4 at room temperature.

The FTIR spectra of 1-4 are recorded in the solid state using KBr pellets (Figs. S1-S4). A band centered
at 3395 cm-1 for 1, 3514 cm-1 for 2, 3484 cm-1 for 3, and 3381 cm-1 for 4 is due to the O-H stretching
frequency of the coordinated/lattice water molecules. All 1-4 also have a band around ~2900 cm-1, which
indicates can be assigned for the alkane stretching of C-H bond. A band at 1710 cm -1, 1737 cm-1 and 1680
cm-1 observed in 1, 2 and 4, respectively, corresponds to the stretching frequency of the carbonyl group of
the carboxylic acid in the ligand. The asymmetric and symmetric stretching frequencies for the
coordinated carboxylate groups appear at 1621/1435 cm -1 (2), at 1582/1327 cm-1 (3) and 1586/1424 cm-
1
(4), respectively. The difference between asymmetric and symmetric stretching frequencies is 186 cm -
1
for 2, 255 cm-1 for 3, and 162 cm-1 for 4. These values indicate that the carboxylate group binds to the
metal ion in a bidentate chelating mode in 2 and 4 (160-200 cm-1) and in a monodentate fashion in 3 (>
230 cm-1). The selected FTIR peaks for 1-4 are listed in Table S1.

3.2 Optical and photoluminescence properties


The optical and electronic properties of 1-4 are studied using UV-visible spectroscopy from 200 to 800
nm. All the compounds are colorless; therefore, these exhibited no absorption in the visible region and
mainly showed absorption only in the UV region (Fig. S5). The band gap energy (E g ) for 1-4 is
calculated from the Tauc plot of (αhν)2 vs hν using the relationship (αhν)2 = Ed(hν - Eg), where α is the
absorption coefficient, hν = photon energy, and Ed is a constant; the intercept of the near-edge region
gives the band gap value [41]. The band gap value for 1-4 is found to be in between 5.5-6.0 eV (Fig. S6)
indicating that these materials are neither semiconductor (~1 eV) nor insulator (>9 eV) [41].
As reported in the literature for complexes of d10 metal ions and carboxylates, [42-47], 1-4 are expected to
exhibit solid state luminescence at room temperature. With an excitation at a wavelength of 220 nm, the
emission was observed at 290 nm for 1, 288 nm for 2, 304 nm for 3 and288 nm for 4(see Fig. 2). The
peaks occurring in the range of 280-304 nm is due to the π-π*transition. Apart from this, the emission
peaks observed at 390 nm for 1, 360 nm for 2, 370 nm for 3 and 360 nm for 4 are due to the n-π*
transition.

Fig. 2 Photoluminescence spectra of 1-4. Excitation wavelength is 220 nm.

3.3 Thermal stability


In order to study the thermal stability of 1-4 as a function of temperature, their thermogravimetric analysis
(TGA) was carried out between 30 and 500 °C under dinitrogen atmosphere (Fig. 3). Compound 1 shows
an initial loss of one chlorine (calc.: 9.9%, found: 8.4%) between 30-108 °C, after which it is found to be
stable up to 150 °C. Compound 2 shows an initial loss of a water molecule and carbon dioxide from the
Htoua linker (calculated: 13.7%, found: 13.6%) between 30-105 °C. After the initial loss, its stability up
to 270 °C is correlated to its polymeric structure. Compound 3 shows an initial loss of a water molecule
and carbon dioxide (calculated: 18.13%, found: 18.74%) between 30-195 °C. Similar to 2, it is also stable
up to 230 °C. Compound 4 shows a weight loss of 39.6% (cal. 40.6%) between 50-200 °C corresponding
to a water molecule, nitrate and carboxylate moieties.

Fig. 3. TGA scans of 1-4.


3.4 Mass spectrometry studies
With the help of ESI mass spectrometry, it was possible to confirm that 1-4 retained their identity in the
solution state (Figs. S7-S10). The characteristic peaks obtained are listed in Table S2. The most
prominent peaks observed are m/z 370.96 (1) and 334.9/332.9/331.9 (2-4) assignable to {Cd(H2toua)Cl}
and {Cd(Htoua)} species, respectively (where Htoua = C8H13O7). The most intense peak m/z 245.534
assigned to {Cd(C5H9O4)} was observed in all compounds due to the loss of part of H2toua molecule.
Specifically, in 3 and 4, the most intense peak m/z 261.0 assigned to {Cd(C 5H10O4)} species was due to
the loss of one ethereal oxygen along with carboxylic group. The second intense peak are m/z 275.5
observed for {Cd(C5H9O4)} species was due to the loss of a carboxylic moiety in 3 and 4.

3.5 Surface analysis

In order to demonstrate the morphological diversity in 1-4 by changing the counter anion in the metal salt
starting materials, their FESEM images were collected and analyzed (Fig. 4). With chloride as a counter
anion for 1, its morphology is found to be rod-like with a thickness of 1.79-1.96 µm. On the other hand, 2
has an aggregate spherical shape morphology. Interestingly, 3 exhibits a porous cage type morphology. In
case of 4, agglomerated particles were obtained. Table 2 summarizes the results for 1-4. The morphology
obtained for 1-4 clearly indicates that the counter anion of the starting metal salts has a great influence on
their morphology [48]. The energy dispersive X-ray analysis (EDX) spectra and elemental mapping of 1-4
are shown in Fig S11-S14. In addition to chemical stoichiometry, the existence of each element present in
these compounds is confirmed.

Table 2
Variation in the Morphology of 1-4.

S. Compound Morphology
No.
1. [Cd(H2toua)Cl2].H2O
Rod-shape
2. {[Cd(Htoua)(H2O)](ClO4)}n
Aggregate Spherical
3. [Cd(toua)].H2O
Porous Cage
4. [Cd(Htoua)(NO3)(H2O)]n
Agglomerated particle

3.6 Single crystal structure description

Compound 1 crystallized in the monoclinic space group P21/c. The Cd(II) ion in 1 is heptacoordinate
surrounded by five oxygen atoms from H2toua, and two chlorine atoms. The geometry around the Cd is
thus a distorted pentagonal bipyramidal with the axial sites occupied by chlorine atoms and equatorial
sites occupied by oxygen atoms of H2toua as shown in Fig. 5. In this structure, H2toua is acting as a
pentadentate ligand using three ethereal oxygens and one oxygen from each of the carboxylic acid group.
The most striking feature in this structure is that H2toua binds to the Cd(II) with carboxylic acid groups
intact, which is very rare. Furthermore, the involvement of five oxygen atoms of H 2toua generates a girdle
around the Cd, where four five-membered chelate rings are present. The axial Cd-Cl distances are
2.5183(6) and 2.4567(6) Å. The Cd-Oether distances are in the range from 2.3953(13) to 2.5127(13) Å.
These values are similar to those found in coordination complexes of Cd(II) containing similar donor
atoms [49].
The Cd-Ocarb distances are 2.3761(14) and 2.4280(13) Å. These values fall within a wide range of
distances (2.279(4)-2.728(5) Å) found in Cd(II)- carboxylate complexes [50]. The O-Cd-O angles
involving five-membered rings range from 65.25(4) to 69.23(5)°, which indicates subtle deviation from
the ideal 72° in a pentagonal system. On the other hand, the open Ocarb-Cd-Ocarb angle (O3-Cd-O7) is
91.23(5)°. The basal oxygen atoms from the diacid are almost coplanar.

Fig.4. FESEM images of (a) rod like (1), (b) aggregate spherical (2), (c) porous cage (3) and (d)
agglomerate particles (4).

Fig. 5. Coordination environment around the Cd(II) center in 1.

The lattice water molecule (O8) present in 1 plays a crucial role in forming a supramolecular assembly
through strong hydrogen bonding (Fig. 6). It acts as both donor and acceptor for such bonding. Using
both hydrogens it connects two molecules of 1 via O-H…Cl interaction(H15…Cl1 and H16…Cl2
distances: 2.349(10) Å and 2.350(10) Å, respectively; O8-H15…Cl1 and O8-H16…Cl2 angles: 178 (3)°
and 174 (3)°, respectively) in one direction. On the other hand, it connects the uncoordinated OH group of
the carboxylic acid of the third molecule of 1in another direction:O6-H5…O8 (H5…O8 distance: 1.87(3)
Å; O6- H5…O8 angle: 169 (4)°. This arrangement is propagated throughout the supramolecular
assembly.
Fig. 6. Hydrogen-bonded network in 1.

Compound 2 crystallized in the orthorhombic space group Cmc21. In this case, the linker H2toua is mono
deprotonated and thus there is one perchlorate anion present (over two positions) in the asymmetric unit.
Like 1, Cd(II) ion in 2 is heptacoordinate with a distorted pentagonal bipyramidal geometry having an O7
environment. Six out of these oxygen atoms come from the Htoua and one from the coordinated water
molecule. Unlike 1, the axial donor atoms in 2 are an oxygen from the water molecule and one oxygen of
the carboxylate group that bridges between two Cd centers in its polymeric structure (Fig. 7). Thus, the
Htoua ligand acts a hexadentate ligand. With five equatorial oxygen atoms that are almost coplanar, a
girdle comprised of four five-membered ring is formed around the Cd center. The Cd-O ether distances are
in the range from 2.343(3) to 2.477(4) Å. The Cd-Ocarbdistances are 2.373(4) Å for the unprotonated
carboxylic acid, and 2.247(4) Å (bridging one) and 2.338(3) Å (monodentate one) for the carboxylate
oxygens.

Fig. 7. Representation of1D coordination polymerof 2 (perchlorates are omitted for clarity); selected part
is expanded with atom labelling to show the repeat unit with a focus on the coordination environment
around the Cd(II) center.
The Cd-Owaterdistance is 2.245(4) Å. These values fall within a wide range of Cd–O bonds in the six- and
seven-coordinated structures reported for Cd(II) coordination polymers [51,52]. The O-Cd-O angles
involving five-membered rings range from 66.28(12) to 68.42(12)°, which indicates subtle deviation
from the ideal 72° in a pentagonal system. On the other hand, the open O carb-Cd-Ocarb angle (O1-Cd-O6)
is 90.92(14)°. The axial water O8 on the Cd(II) center is hydrogen bonded to the oxygen atoms of the
perchlorate anion (O8-H8a…O9a (H8a…O9a distance: 1.840 Å; O8-H8a…O9a angle: 166.43°; (O8-
H8b…O12 (H8b…O12 distance: 2.142 Å; O8-H8b…O12 angle: 174.08°)such that the two layers of the
1D polymeric units in 2 are linked to form higher order supramolecular assembly as shown in Fig. 8.

Fig. 8. Supramolecular network in 2 through hydrogen bonding (hanging contacts are deleted for clarity).

3.7 Hirshfeld surface analysis of 1 and 2

The distinct features of 1 and 2 with respect to the molecular geometries of the linker (H2toua)
and the pattern of the non-covalent interactions encouraged us to quantify the contribution of the
interactions. In this context, we have generated the Hirshfeld surfaces [35,56] (HSs) of the title
compounds, that have been mapped over dnorm in the range from −0.500 (red) to 1.200 (blue) Å
(see Fig. 9). The strength of the intermolecular interaction is taken by using color codes where the
red regions represent the shorter contacts (negative dnorm values) and blue regions represent the
longer contacts (positive dnorm values), while the white regions indicate the contacts close to the
van der Waals limit by mapping the dnorm functions.
In both compounds, the large circular depression (deep red) is appeared due to strong hydrogen
bonding interactions (O···H / Cl···H) and the absence of C ̶ H⋅⋅⋅π and π⋅⋅⋅π interactions represent
the blue regions. However, white spots over linker moieties are indicative of the C⋅⋅⋅H and H⋅⋅⋅H
interactions.

Fig. 9. Hirshfeld surfaces mapped with dnorm (left), shape index (middle), and fragment patches
(right) for compounds (a) 1and (b) 2.

The contacts that are responsible in the crystal packing are estimated with respect to their
contribution to the crystal structure. The scattered points of the fingerprint plots [38,39] evidence
all interactions involved within the structures. To quantify each individual contact, we have
decomposed the full-fingerprint plots in unique visual mode. All major contacts that are involved
within the structures of the title compounds are quantified accordingly and are included in Fig. 10.
The very strong O⋯H/H⋯O interactions in both compounds are evidenced by the sharp tips in the
region (de = di≃ 1.0Å) contributes 33.3% in 1 and 61.5% in 2. This difference in contribution is
due to the presence of two coordinated Cl atoms in 1, whereas one water molecule is coordinated
in 2. The Cl⋯H/H⋯Cl interactions in 1 contribute 29.5%, as evidenced by the spike in the region
(de = di≃ 1.4Å). In both compounds, the H⋯H interactions represent the significant contribution
to the HSs of 1 (32.3%) and 2 (25.8%). Due to the trimming of polymeric chain of 2, it shows the
contribution of 4.3% and 2.0% for C···O/O···C and Cd⋯O contacts, respectively. These are the
non-covalent interactions that contribute the most to the packing of the title compounds.

Fig. 10. Full and decomposed 2D-fingerprint plots corresponding to various contacts involved
within the structure of compound 1 and 2.

4. Conclusion

In summary, we explored the chemistry of a flexible polyether based dicarboxylic acid (H 2toua), where
fourCd(II) salts containing different anions, such as Cl -, ClO4-, OAc-, and NO3-, were utilized to make
structurally diverse [Cd(H2toua)Cl2].H2O (1),{[Cd(Htoua)(H2O)](ClO4)}n(2),[Cd(toua)].H2O (3) and
[Cd(Htoua)(NO3)(H2O)]n (4), respectively.Their spectroscopic and structural characterization by
numerous analytical techniques allowed us to understand their formation and properties for
furtherexploitation. Single crystal X-ray structures of 1 and 2indicated that the formation of a discrete
compound vs one dimensional coordination polymer, respectively, was due to a mere change of counter
anion from chloride to perchlorate in the starting metal salt. Further structural analysis of 1 and 2 was
carried out by 3D Hirshfeld surface analysis and 2D fingerprint plots. Their spectroscopic and thermal
properties allowed us to utilize these compounds for further chemistry. The first series of compounds of
H2toua reported hereprovides an avenue to synthesize several newcompounds of many other metal
centers. Furthermore, compounds 1 and 2 (and derivatives of other metal ions) can be used as
metalloligands to explore the formation of new homo and heteronuclear clusters. Such work is underway
in our laboratory for their applications in various fields.

Declaration of Competing Interest

There are no competing interests to declare.


Acknowledgements

This was supported by the TEQIP-III, NIT Jalandhar. Ms. Sakshi is thankful to Ministry of Education
(MOE), India for a research fellowship. Authors acknowledge the help of Ms. Rupinder Kaur and Dr.
Datta Markad for their kind help.

References

[1] Chuanqi Zhang, Chao Chen , Song Liu, Ning Zhang, Modulated assembly of interdigitated 1D
isolated tube and 2D highly undulating layer based on N-donor ligand and multidentate
carboxylic acid, Inorganic Chemistry Communications 20 (2012) 18–22.
https://doi.org/10.1016/j.inoche.2012.02.004
[2] S. Khullar, S. K. Mandal, Structural Diversity of Mn(II) Complexes with Acetylene
Dicarboxylate and Hexadentate Ancillary Ligands under Ambient Conditions: Effect of
Methylene Chain Length on Coordination Architectures, Cryst. Growth Des. 13 (2013) 3116-
3125, https://doi.org/10.1021/cg400507n.
[3] J. W. Uebler, A. L. Pochodylo, R. J. Staples, R. L. LaDuca, Control of Self-Penetration and
Dimensionality in Luminescent Cadmium Succinate Coordination Polymers via Isomeric
Dipyridylamide Ligands, Cryst. Growth Des. 13 (2013) 2220–2232,
https://doi.org/10.1021/cg4003302.
[4] X. Z. Guo, S. S. Chen, W. D. Li, S. S. Han, F. Deng, R. Qiao, Y. Zhao, Series of Cadmium(II)
Coordination Polymers Based on a Versatile Multi-N-Donor Tecton or Mixed Carboxylate
Ligands: Synthesis, Structure, and Selectively Sensing Property, ACS Omega 4 (2019) 11540–
11553, https://doi.org/10.1021/acsomega.9b01108.
[5] C. Hu, U. Englert, Coordination polymers based on cadmium(ii) pyridine complexes: synthesis,
range of existence, and structure, CrystEngComm 4 (2002) 20-25,
https://doi.org/10.1039/B110208G.
[6] T. Qiu, Z. Liang, W. Guo, H. Tabassum, S. Gao, R. Zou, Metal–Organic Framework-Based
Materials for Energy Conversion and Storage, ACS Energy Lett. 5 (2020) 520-532,
https://doi.org/10.1021/acsenergylett.9b02625.
[7] A. E. Baumann, D. A. Burns, B. Liu, V. Sara Thoi, Metal-organic framework functionalization
and design strategies for advanced electrochemical energy storage devices, Commun. Chem. 2
(2019) 86-100, https://www.nature.com/articles/s42004-019-0184-6.
[8] F.-A. Li, W.-C. Yang, S-Z. Bai, A 3D cadmium coordination polymer with 2-fold
interpenetration structure and helical chains based on biphenyl-4-hydroxyl-3,3′-bicarboxylic acid:
Synthesis and crystal structure and properties, Inorganic Chemistry Communications 46 (2014)
65–68, https://doi.org/10.1016/j.inoche.2014.04.035
[9] X.-L. Tong, D.-Z.Wang, T.-L.Hu, W.-C. Song, Y.Tao, X.-H. Bu, Zinc and Cadmium
Coordination Polymers with Bis(tetrazole) Ligands Bearing Flexible Spacers: Synthesis, Crystal
Structures, and Properties, Cryst. GrowthDes. 9 (2009) 2280–
2286, https://doi.org/10.1021/cg8010629.
[10] L.Y. Lu, X.W. Tao, F.Y. Chen, A. L. Cheng, Q.S. Xue, E.Q. Gao, A series of new sulfone-
functionalized coordination polymers: Fascinating architectures and efficient fluorescent sensing
of nitrofuran antibiotics, J. Solid State Chem. 301 (2021) 122251,
https://doi.org/10.1016/j.jssc.2021.122251.
[11] X.M. Lin, T.T. Li, Y.W. Wang, L. Zhang, C.Y. Su, Two ZnII Metal-Organic Frameworks
with Coordinatively Unsaturated Metal Sites: Structures, Adsorption, and Catalysis, Chem. Asian
J. 7 (2012) 2796-2804, https://doi.org/10.1002/asia.201200601.
[12] Y.L. Wu, Y.T. Yan, T.S. Liang, H.H. Liang, G. Yang,X.L.Su, X.H. He, M.T. Khan, Y.Y.
Wang, Synthesis of two new Cd(II)-MOFs based on different secondary building units with
highly selective gas sorption for CO2/CH4 and luminescent sensor for Fe3+ and Cr2O72- ions, J.
Solid State Chem. 285 (2020) 121258, https://doi.org/10.1016/j.jssc.2020.121258.
[13] X. Guo, J. Xu, J. Sun, X. Chen, L. Wang, Y. Fan, Three layer-structured cadmium
coordination polymers based on flexible 5-(4-pyridyl)-methoxylisophthalic acid: rapid synthesis and
luminescence sensing†, CrystEngComm 21 (2019) 1001-1008,
https://doi.org/10.1039/C8CE01524D
[14] L.-L. Qu, Y.-L. Zhu, J. Zhang, Y.-Z. Li, H.-B. Du, X.-Z. You, Structural diversity and
properties of coordination polymers built from a semi-rigid tetradentenate carboxylic acid†,
CrystEngComm 14 (2012) 824-831, https://doi.org/10.1039/C1CE05931A
[15] S. Tripathi, G. Anantharaman, Architectures varying from discrete molecular units to 2-
dimensional coordination polymers and photoluminescence behavior of zinc and cadmium comprising
an anionic zwitterion of rigid 4,5-dicarboxy-1,3-dimethyl-1H-imidazolium iodide†, CrystEngComm 17
(2015) 2754-2768, https://doi.org/10.1039/C4CE02215G.
[16] A. K. Gupta, K. Tomar, P. K. Bharadwaj, Structural diversity of Zn(II) based coordination
polymers constructed from a flexible carboxylate linker and pyridyl co-linkers: fluorescence sensing of
nitroaromatics†, New J. Chem. 41 (2017) 14505-14515, https://doi.org/10.1039/C7NJ02651J.
[17] S. Khullar, S. K. Mandal, Effect of Spacer Atoms in the Dicarboxylate Linkers on the
Formation of Coordination Architectures—Molecular Rectangles vs 1D Coordination Polymers:
Synthesis, Crystal Structures, Vapor/Gas Adsorption Studies, and Magnetic Properties, Cryst.
Growth Des. 14 (2014) 6433-6444, https://doi.org/10.1021/cg501284y.
[18] R. D. Marchenko; T. S. Sukhikh, A.A. Ryadun, A. S. Potapov, Synthesis, Crystal
Structure, and Luminescence of Cadmium(II) and Silver(I) Coordination Polymers Based on 1,3-
Bis(1,2,4-triazol-1-yl)adamantine, Molecules 26 (2021) 5400-5410,
https://doi.org/10.3390/molecules26175400.
[19] M. Gupta, D. De, K. Tomar, P. K. Bharadwaj, From Zn(II)-Carboxylate to Double-Walled
Zn(II)-Carboxylato Phosphate MOF: Change in the Framework Topology, Capture and
Conversion of CO2, and Catalysis of Strecker Reaction, Inorg. Chem. 56 (2017) 14605-14611,
https://doi.org/10.1021/acs.inorgchem.7b02443.
[20] G.-C. Xu, Q. Hua, T. Okamura, Z.-S. Bai, Y.-J. Ding, Y.-Q. Huang, N. Ueyama,
Cadmium(II) coordination polymers with flexible tetradentate ligand 1,2,4,5-tetrakis(imidazol-1-
ylmethyl)benzene: anion effect and reversible anion exchange property†, CrystEngComm 11 (2009)
261–270, https://doi.org/10.1039/B813220H.
[21] Z.-X. Li, T.-L. Hu, H. Ma, Y.-F. Zeng, C.-J. Li, M.-L. Tong, X.-H. Bu, Adjusting the
Porosity and Interpenetration of Cadmium(II) Coordination Polymers by Ligand Modification:
Syntheses, Structures, and Adsorption Properties, Cryst. Growth Des. 10 (2010) 1138–1144,
https://doi.org/10.1021/cg900980y.
[22] X.-D. Chen, J.-H. Guo, M. Du, T. C. W. Mak, Copper and cadmium coordination
polymers of di-3-pyridinylmethanone: Subtle anion effect on the dimensionality of Cd(II)
frameworks, Inorganic Chemistry Communications 8 (2005) 766–768,
https://doi.org/10.1016/j.inoche.2005.06.001
[23] H.-Y.Bai, J.-F. Ma, J. Yang, Y.-Y. Liu, Hua-Wu, J.-C. Ma, Effect of Anions on the Self-
Assembly of Cd(II)-Containing Coordination Polymers Based on a Novel Flexible
Tetrakis(imidazole) Ligand, Cryst. Growth Des. 10 (2010) 995–1016,
https://doi.org/10.1021/cg901332m.
[24] S. A. A. Razavi, M. Y. Masoomi, T.Islamoglu, A. Morsali, Y. Xu, J. T. Hupp, O. K. Farha,
J. Wang, P. C. Junk, Improvement of Methane–Framework Interaction by Controlling Pore Size
and Functionality of Pillared MOFs, Inorg. Chem. 56 (2017) 2581–2588,
https://doi.org/10.1021/acs.inorgchem.6b02758.
[25] A. A. García-Valdivia, A. Zabala-Lekuona, G. B. Ramírez-Rodríguez, J. M. Delgado-
López, B. Fernández,J.Cepeda, A. Rodríguez-Diéguez, 2D-Coordination polymers based on 1H-
indazole-4-carboxylic acid and transition metal ions: magnetic, luminescence and biological properties†,
CrystEngComm. 22 (2020) 5086-5095, https://doi.org/10.1039/D0CE00544D.
[26] Q. Cheng, L. Qin, C. Ke, J. Zhou, J. Lin, X. M. Lin, G. Zhang, Y. P. Cai, Four new Zn(II)
and Cd(II) coordination polymers using two amide-like aromatic multi-carboxylate ligands: synthesis,
structures and lithium–selenium batteries application†, RSC Adv. 9 (2019) 14750–14757,
https://doi.org/10.1039/C9RA02163A.
[27] B. Li, J. Wang, J.Gao, Y. Yu, D. Ma, A semi-rigid tricarboxylate ligand based Co(II)
coordination polymer: construction and applications in multiple sensing†, New J. Chem. 44 (2020)
3664-3671, https://doi.org/10.1039/C9NJ05766H.
[28] S. Henke, R. Schmid, J.D. Grunwaldt, R.A. Fischer, Flexibility and Sorption Selectivity in
Rigid Metal–Organic Frameworks: The Impact of Ether-Functionalised Linkers, Chem. Eur. J. 16
(2010) 14296 – 14306, https://doi.org/10.1002/chem.201002341.
[29] S. McCann, M. McCann, M.T.Casey, M. Devereux, V. McKee, P. McMichael, J.
G.McCrea, Manganese(II) complexes of 3,6,9-trioxaundecanedioic acid (3,6,9-tddaH 2): X-ray
crystal structures of [Mn(3,6,9-tdda) (H2O)2]·2H2O and {[Mn(3,6,9-tdda)(phen)2·3H2O]·EtOH}n,
Polyhedron 16 (1997) 4247–4252, https://doi.org/10.1016/S0277-5387(97)00233-7.
[30] M. Miyazaki, Y. Shimoishi, H. Miyata, K. Tôei, The reaction of dicarboxylic acids
containing ether linkages with alkaline earth metals, J. Inorg.Nucl. Chem. 36 (1974) 2033–2038,
https://doi.org/10.1016/0022-1902(74)80718-9.
[31] APEX2, SADABS, SAINT, Bruker AXS Inc: Madison WI, USA 2008,
[32] G. M. Sheldrick, A Short History of SHELX, ActaCrystallogr. Sect. A: Found.
Crystallogr. 64 (2008) 112, https://doi.org/10.1107/S0108767307043930.
[33] C. F.Macrae; I. J. Bruno, J. A. Chisholm, P. R. Edgington, P. McCabe, E. Pidcock, L.
Rodriguez-Monge, R. Taylor, J. Van De Streek, P. A. Wood, Mercury CSD 2.0 - New Features
for the Visualization and Investigation of Crystal Structures. J. Appl. Crystallogr. 41 (2008) 466–
470, http://dx.doi.org/10.1107/S0021889807067908.
[34] A. L. Spek, PLATON version 1.62, University of Utrecht, 1999,
[35] J. J. McKinnon, A. S. Mitchelland, M. A. Spackman, Hirshfeld Surfaces: A New Tool for
Visualising and Exploring Molecular Crystals, Chem. – Eur. J. 4 (1998) 2136-2141,
https://doi.org/10.1002/(SICI)1521-3765(19981102)4:11%3C2136::AID-
CHEM2136%3E3.0.CO;2-G.
[36] M. A. Spackmanand D. Jayatilaka, Hirshfeld surface analysis, CrystEngComm 11 (2009) 19-
32, https://doi.org/10.1039/B818330A.
[37] F. L. Hirshfeld, Bonded-atom fragments for describing molecular charge densities, Theor.
Chim. Acta 44 (1977) 129-138, https://doi.org/10.1007/bf00549096.
[38] J. J. McKinnon, D. Jayatilaka, M. A. Spackman, Towards quantitative analysis of
intermolecular interactions with Hirshfeld surfaces, Chem. Commun. 267 (2007) 3814-3816,
https://doi.org/10.1039/B704980C.
[39] M. A. Spackman, J. J. McKinnon, Fingerprinting intermolecular interactions in molecular
crystals†, CrystEngComm 4 (2002) 378–392, https://doi.org/10.1039/B203191B.
[40] P. R. Spackman, M. J. Turner, J. J. McKinnon, S. K. Wolff, D. J. Grimwood, D.
Jayatilaka, M. A. Spackman, CrystalExplorer : a program for Hirshfeld surface analysis,
visualization and quantitative analysis of molecular crystals, J. Appl. Cryst. 54 (2021) 1006-
1011, https://doi.org/10.1107/S1600576721002910.
[41] B. D. Viezbicke, S. Patel, B. E. Davis, E. Benjamin, D. P. Birnie, III, Evaluation of the
Tauc method for optical absorption edge determination: ZnO thin films as a model system
Physica Status Solidi 252 (2015) 1700-1710, https://doi.org/10.1002/pssb.201552007.
[42] M. O’Keeffe, Design of MOFs and intellectual content in reticular chemistry: a personal view†,
Chem. Soc. Rev. 38 (2009) 1215-1217, https://doi.org/10.1039/B802802H.
[43] M. J. Moloto, N.Revaprasadu, G. A. Kolawole, P. O’Brien, M. A. Malik, M.Motevalli,
Synthesis and X-Ray Single Crystal Structures of Cadmiym(II) Complexes:
CdCl2[CS(NHCH3)2]2 and CdCl2(CS(NH2)NHC6H5)4-Single Source Precursors to CdS
Nanoparticles, E-Journal of Chem. 7 (2010) 1148-1155, https://doi.org/10.1155/2010/561498.
[44] T. C. Stamatatosa, E. Katsoulakoua, V. Nastopoulosa, C. P. Raptopouloub, E. Manessi-
Zoupaaand S. P. Perlepesa, Cadmium Carboxylate Chemistry: Preparation, Crystal Structure, and
Thermal and Spectroscopic Characterization of the One-dimensional Polymer
[Cd(O2CMe)(O2CPh)(H2O)2]n, Z. Naturforsch 58b (2003) 1045-1054,
https://doi.org/10.1515/znb-2003-1103.
[45] R. Das, P. Pachfule, R. Banerjee, P. Poddar, Metal and metal oxidenanoparticle synthesis from
metal organic frameworks (MOFs): finding the border of metal and metal oxides†, Nanoscale 4 (2012)
591-599, https://doi.org/10.1039/C1NR10944H.
[46] D. Z. Husein, R. Hassanien, M. Khamis, Cadmium oxide nanoparticles/graphene composite:
synthesis, theoretical insights into reactivity and adsorption study†, RSC Adv. 11 (2021) 27027-27041,
https://doi.org/10.1039/D1RA04754J.
[47] M. Nasrullah, F. Z. Gul, S. Hanif, A. Mannan, S. Naz, J. S. Ali, M. Zia, Green and
Chemical Syntheses of CdO NPs: A Comparative Study for Yield Attributes, Biological
Characteristics, and Toxicity Concerns, ACS Omega 5 (2020) 5739–5747,
https://dx.doi.org/10.1021/acsomega.9b03769.
[48] F. Afkhami, A. Khandar, G. Mahmoudi, R. Abdollahi, A. Gurbanov, A. Kirillov,
Sonochemical Synthesis of Cadmium(II) Coordination Polymer Nanospheres as Precursor for
Cadmium Oxide Nanoparticles, Crystals 9 (2019) 199, https://doi.org/10.3390/cryst9040199.
[49] S. Khullar, S. K. Mandal, Non-hydrothermal synthesis, structural characterization and
thermochemistry of water soluble and neutral coordination polymers of Zn(II) and Cd(II): precursors for
the submicron-sized crystalline ZnO/CdO†, RSC Adv. 4 (2014) 39204–39213,
https://doi.org/10.1039/C4RA03928A.
[50] M. Fernández, A. Martínez, J. C. Hanson, A. J. Rodriguez, Nanostructured Oxides in
Chemistry: Characterization and Properties, Chem. Rev. 104 (2004) 4063–4104,
https://doi.org/10.1021/cr030032f.
[51] H. S. Wang, W. Shi, J. Xia, H. B. Song, H. G. Wang, P. Chen, Six-, seven- and eight-
coordinated Cd(II) ions with N-heterocyclic multicarboxylic acids, Inorg. Chem. Comm. 10
(2007) 856–859, https://doi.org/10.1016/j.inoche.2007.03.024.
[52] J. Q. Liu, J. Wu, T. Wu, A Luminescent Eight-Coordinated 2D Cd(II) Framework Material
with Flexible Multi-Carboxylate Ligand, Synthesis and Reactivity in Inorganic, Metal-Organic,
and Nano-Metal Chemistry 40 (2010) 231-236, https://doi.org/10.1080/15533171003766378.
[53] M. Mohsen, I. Naeem, M. Awaad, H. Tantawy, A. Baraka, A cadmium-imidazole
coordination polymer as solid state buffering material: Synthesis, characterization and its use for
photocatalytic degradation of ionic dyes, J. Solid State Chem. 289 (2020) 121493,
https://doi.org/10.1016/j.jssc.2020.121493.
[54] S. Kumar, A. K. Ojha, B. Walkenfort, Cadmium oxide nanoparticles grown in situ on
reduced graphene oxide for enhanced photocatalytic degradation of methylene blue dye under
ultraviolet irradiation, J.Photochem. and Photobio. B: Biology 159 (2016) 111–119,
https://doi.org/10.1016/j.jphotobiol.2016.03.025.
[55] C. Xu, G. P. Rangaiah, X. S. Zhao, Photocatalytic Degradation of Methylene Blue by
Titanium Dioxide: Experimental and Modeling Study, Ind. Eng. Chem. Res. 53 (2014) 14641–
14649, https://doi.org/10.1021/ie502367x.
[56] H. A. Alrafai, Z. A. Al-Ahmed, M. K. Ahmed, M. Afifi, K. R. Shoueir and A. Abu
Rayyan, The degradation of methylene blue dye using copper-doped hydroxyapatite encapsulated
into polycaprolactone nanofibrous membranes†, New J. Chem. 45 (2021) 16143-16154,
https://doi.org/10.1039/D1NJ01623G.

You might also like