Jana Piska,
Ivica Đilovića,
Tomica Hrenara,
Danijela Cvijanovićb,
Gordana Pavlovićc and
Višnja Vrdoljak*a
aUniversity of Zagreb, Faculty of Science, Department of Chemistry, Horvatovac 102a, 10000 Zagreb, Croatia. E-mail: visnja.vrdoljak@chem.pmf.hr
bUniversity of Zagreb, School of Medicine, Department of Chemistry and Biochemistry, Šalata 3, 10000 Zagreb, Croatia
cUniversity of Zagreb, Faculty of Textile Technology, Division of Applied Chemistry, Prilaz baruna Filipovića 28a, 10000 Zagreb, Croatia
First published on 20th October 2020
Synthesis of hydrazones (1a–4a and 1b–4b), quinazolines (3c·MeOH and 3d·MeOH), and hydrazone-Schiff bases (4c and 4d) is achieved by combining suitable aldehydes (2,3- or 2,4-dihydroxybenzaldehyde) with four hydrazides (isonicotinic, nicotinic, and 2- or 4-aminobenzoic acid hydrazide). A suite of approaches for their preparation is described: solution-based synthesis, mechanosynthesis, and solid-state melt reactions. The mechanochemical approach is generally a better choice for the quinazolines, while the solid-state melt reaction is more efficient for derivatives of (iso)nicotinic based hydrazones. Crystalline amine-functionalised hydrazones 4a and 4b undergo post-synthetic modifications in reactions with 3- or 4-pyridinecarbaldehyde vapours to form hydrazone-Schiff bases 4a-3py, 4b-3py, 4a-4py, and 4b-4py. Mechanochemical and vapour-mediated reactions are followed by ex situ powder X-ray diffraction and IR-ATR methods, respectively. The chemometric analysis of these data using principal component analysis provided an insight into the reaction profiles and reaction times. Azines (5a and 5b), achieved from aldehydes and hydrazine, reversibly change colour in response to temperature changes. The structures of all products are ascertained by a combined use of spectroscopic and X-ray diffraction methods. The cytotoxic and antimicrobial activities of all compounds against selected human cancer cell lines and bacterial strains are evaluated.
The most common pathway for the preparation of these hydrazide-based compounds includes the reaction of appropriate hydrazides with different aldehydes or ketones in various organic solvents. Different solution-based synthetic approaches have been tested and combined.16–19 Gudasi et al. described the ineffectiveness in synthesizing hydrazones by condensing pyridine-2-carbaldehyde with 2-aminobenzoylhydrazide.17–19 It was found that reactions of the corresponding hydrazide and 3-methoxybenzaldehyde, in a molar ratio of 1:1, led to the formation of quinazolin-4(3H)-one. Recently, the mechanochemical synthetic route provided an efficient solution for obtaining hydrazones20 and 3-arylideneaminoquinazolin-4(1H)-one compounds, from various aldehydes.21
On account of it, 2,3- and 2,4-dihydroxybenzaldehyde were used as carbonyl forerunners, while isonicotinic and nicotinic acid hydrazides, were used as hydrazides for the hydrazones synthesis. Additionally, 2- and 4-aminobenzoic acid hydrazides were also investigated as compounds that provided classical hydrazones, quinazolinone or hydrazone-Schiff bases via controlled condensation. Conventional and a variety of green synthesis methodologies were used and critically compared (Fig. 1). We aimed to investigate their effect on reaction time, yield, as well as the crystallinity and purity of the prepared compounds. This study was also focused on applicability of the synthetic procedures and their selectivity towards the desired product. The synthetic path was refreshed by employing the “melt synthesis” technique.
Fig. 1 Synthetic procedures applied for the preparation of hydrazones, quinazolines, and hydrazine-Schiff bases. |
Also, post-synthetic condensation with aldehyde vapours was applied to yield hydrazone-Schiff bases. The chemometric study was used for monitoring the starting material conversion by mechanochemical and vapour-mediated synthesis by processing ex situ powder X-ray diffraction and ATR spectra, respectively. Principal component analysis of these data was applied as a data reduction technique where collection of data (diffractograms or spectra) were decomposed in order to provide linearly independent variables. These variables can be plotted in one or two dimensions giving a detailed insight into reaction profiles and allowing identification of the reaction end points. Adequacy of the used method was validated by visual inspection of the final PXRD diffractograms and ATR spectra with the final product and in each case, it was confirmed that reaction finished in the exact time obtained by PCA.
Crystal and molecular structures of hydrazones 1a·MeOH, 1b·H2O, 3a, 3b, 4a, and 4b·MeOH, quinazolinone 3c·MeOH, hydrazone-Schiff base 4c, and azine 5b·H2O were determined by the single crystal X-ray diffraction method, providing proposed investigation more comprehensive. It should be noted that in Cambridge Structural Database (CSD)22 there are few hydrazone structures reported with 2,3- and 2,4-dihydroxybenzaldehyde fragment.23–40 All of the investigated compounds were further characterized by elemental analysis, thermogravimetric and differential scanning calorimetry measurements, IR-ATR and NMR spectroscopy. The solid-state thermochromic changes of azines and influence of the central spacer and functional groups on the responsive properties were also explored.
To complement our study, the cytotoxicity of the compounds was evaluated against HepG2 and THP-1 cells. The compounds were also screened for their antimicrobial activity against S. aureus, E. faecalis, E. coli and M. catarrhalis to evaluate influence of the compound type and substituents on their activities.
Elemental analyses were provided by the Analytical Services Laboratory of the Rudjer Bošković Institute, Zagreb, Croatia. The powder X-ray diffraction data (PXRD) for qualitative phase analysis were collected on a Phillips X'Change powder diffractometer in the Bragg–Brentano geometry using CuKα radiation. The data were collected and visualized using the X'Pert programs Suite.41 IR-ATR spectra were recorded on a Perkin Elmer Spectrum One spectrometer. Thermogravimetric analyses (TGA) were conducted on a Mettler Toledo TG/DSC 3+ Stare System coupled with Thermo Fischer Nicolet iS50 FT-IR with Al2O3 crucibles under nitrogen stream in a temperature range between 25 °C and 400 °C, while the heating rate was adjusted to 10 °C min−1. Differential scanning calorimetry (DSC) measurements were performed under the nitrogen stream on the Mettler-Toledo DSC823e calorimeter with aluminium crucibles, in the temperature range 25–400 °C, with the heating rate 10 °C min−1. The results of both TGA and DSC experiments were evaluated using the Mettler STARe 9.01 software. Nuclear magnetic resonance (NMR) spectra were recorded on a Bruker Avance III HD 400 spectrometer operating at 400 MHz. Compounds were dissolved in dmso-d6 and measured in 5 mm NMR tubes at 298 K with TMS as an internal standard.
The structures were solved by direct methods using SHELXT.43 The refinement procedure by full-matrix least-squares methods based on F2 values against all reflections included anisotropic displacement parameters for all non-H atoms. The positions of H atoms each riding on carbon atoms were determined on stereochemical grounds. The positions of other H atoms were determined from the difference Fourier map but were refined using the riding model. In the asymmetric unit of 3c·MeOH, the methanol molecule exhibits a two-site (57:43) disorder. The restraints on geometrical and displacement parameters (DFIX, DELU, SIMU and ISOR) were applied for that molecule. Refinements were performed using SHELXL-97. The SHELX programs operated within the Olex2 v1.2 (ref. 44) suite. Geometrical calculations and molecular graphics were done with PLATON, MERCURY,45 ORTEP,46 PyMOL47 and POV-Ray.48
In one case the data matrix X was composed of PXRD diffractograms whereas in the other the data matrix was composed of ATR spectra. PCA enables one to find the best linear projections for a high dimensional set of data in the least-squares sense. Scores ti represent projections of the original points on the principal component (PC) direction and can be used for classification or building of probability distributions,49 whereas loadings56 represent the eigenvectors of data covariance (or correlation) matrix and can be used for the identification of variability among the data. The initial development of PCA goes back to Beltrami50 and Pearson,51 whereas the name was introduced by Hotelling.52 More details on PCA can be found in the literature.53 Principal component analysis was used as a dimensionality reduction tool for a set of PXRD diffractograms and for a set of ATR spectra. In each cae PCA was performed using a NIPALS algorithm54 implemented in our own program moonee.55,56
In vitro antibacterial activity of the investigated compounds was evaluated against Gram-positive, namely Staphylococcus aureus (ATCC 13709) and Enterococcus faecalis (ATCC 29212), and Gram-negative, namely Escherichia coli (ECM 1556) and Moraxella catarrhalis (ATCC 23246), bacterial strains by broth microdilution according to the CLSI guidelines.59 Details of the bacterial growth and the method used have been described previously.57 Briefly, the bacteria were seeded in 96-well microplates and incubated overnight before they were treated with varying concentrations of the tested compounds. After the treatment, the lowest concentrations at which no bacterial growth was observed were determined by visual inspection and reported as minimum inhibitory concentration (MIC) values. All experiments were performed in duplicate.
Like in many similar compounds, the keto form dominates in the solid state (CO bond is around 1.2 Å). An extensive survey of the Cambridge Structural Database shows that hydrazones derived from salicylaldehyde derivatives predominantly crystallize in the amide tautomeric form with E configuration. This appears to be the case in all the investigated hydrazones (1a·MeOH, 1b·H2O, 3a, 3b, 4a, and 4b·MeOH): all of them have an intramolecular six-membered heteronuclear resonance-assisted hydrogen bond (Fig. 3). The cis configuration of the O1 atom with respect to N2 is present in all the examined molecules.
In all the crystal structures, molecules are almost planar or slightly inclined in a way to maximize their intermolecular hydrogen bonding potential (the planarity is disrupted in the case of solvated compounds). Planarity enhances the delocalization of π electrons through the spacer unit between two aromatic fragments, regardless of the position of substituents. This can be easily seen from the bond length distribution in the central parts of molecules (which fits well by already mentioned tautomeric form). The other, endocyclic, amino or hydroxyl, bond lengths and angles are comparable to those found in the CSD database for similar structures.
The crystal structures of the prepared hydrazones unveil that in all cases molecules are mutually connected with an extensive hydrogen bonding (Fig. 3, 4, 5, and S1–S6, see ESI†). The pyridine N atoms are connected with hydroxyl H atoms, carbonyl O atoms and amido H atoms are usually employed in hydrogen bonding with solvent molecules.
All solvates lose their solvent upon prolonged standing at room temperature. A comparison of PXRD patterns was used to identify obtained forms, Fig. S7, see ESI†. Structures of 1a and 2a were reported previously (CSD code WAFVEG and WOFYUN, respectively) but their synthesis was performed differently.23,24
Even-though a variety of products can be obtained by a solution-based method, large amounts of alcohol was employed. To minimize the use of solvent, and turn the synthetic protocols towards more sustainable ones: mechanochemical and solid-state melt-synthesis were engaged. The idea was to explore pros and cons of the methods as well as their selectivity towards the desired product.
Fig. 6 PXRD patterns of hydrazones obtained by the solution-based method (blue line) and by the mechanosynthesis (green line). |
In the case of 1b and 2a longer milling was necessary which doesn't follow the sustainable conditions. Besides that, under the prolonged grinding the original PXRD patterns were found to be altered and the amorphization degree increased with milling time. Similarly, PXRD diffractograms of 1a, 2b, 3a, 3b, 4a, and 4b additionally milled for 30 min showed great amount of amorphous phase.
The first idea of the proposed research was to follow the formation of all hydrazones by ex situ PXRD and application of principal component analysis for the mechanochemical synthesis monitoring. PXRD were measured and collected in a data matrix X that was decomposed using the PCA. The duration of the milling was systematically increased up to 60 min to observe the influence of the milling duration on the reaction conversion. Even though it seemed that the PXRD patterns of the samples obtained after a particular reaction time fit well (Fig. S8, see ESI†), PCA revealed the problems in this kind of analysis for monitoring the reaction profiles and detection of reaction end.
Satisfactory results were not obtained in majority of cases due to the several causes: (a) mechanochemical reactions were too fast, (b) amorphization of the reaction components with milling time and reduction in overall quality of the data obtained, and (c) continuation of the reaction after the milling was stopped. In this way, after the reduction of data matrix composed of PXRD diffractograms and plotting the scores values in the dependence on time, the time scale on the ordinate was shifted by an unpredictable amount. This amount of time was different for each point in the reaction because the reaction was continued for different time.
Nevertheless, the PCA was able to provide some additional insight into the reaction mechanisms that we encountered and explained in our previous work.57 For 4a ligand the first principal component described only 41.56% of the total variance whereas the second described 27.57%. This low value of variance for the PC1 and relatively high value of variance for PC2 are indicators of many simultaneous reactions (at least two).57 To confirm that fact, principal component loadings were plotted in 2- and 3-dimensional space and inspected visually. 2D and 3D loadings display a particular symmetrical “butterfly” pattern (Fig. 7a and b).57
Fig. 7 Principal component loadings spanned by (a) three and (b) two principal components calculated for a set of PXRD data collected through mechanochemical synthesis of 4a. |
This symmetrical pattern spanned along the first principal component (PC1) suggests that at least two orthogonal eigenvectors extracted from the original PXRD data are necessary for good description of the reaction profile. Since this arrangement is almost symmetrical along PC1 (with symmetry plane cutting the PC2-ordinate at 0.0) it is reasonable to expect that two or even more reactions are occurring simultaneously. To further confirm this fact, we investigated loadings in three dimensions (Fig. 7a). The pattern indeed looks two-dimensional where each branch of the loadings falls symmetrically on the side thus confirming two reaction steps.
However, 2- or 4-aminobenzhydrazide-based hydrazones were not accessible in this manner, and each time reactions gave a mixture of products. This suggests higher reactivity of the amino groups at the 2- and 4- position under these conditions.
Fig. 8 2,3-Dihydroquinazolinones obtained by the reaction of 2-aminobenzhydrazide and 2,3-dihydroxybenzaldehyde, 3c, and 2,4-dihydroxybenzaldehyde, 3d. Red spheres present OH groups. |
When employing 2-aminobenzhydrazide, the best results were obtained by using LAG mechanochemical method, and disubstituted quinazolin-4(3H)-one products 3c·MeOH and 3d·MeOH were prepared, Fig. 8. The heterocyclic product dominates in both cases, and purity of products was confirmed also by NMR analysis.
It seems, when 1:1 condensation product under solution-based or mechanochemical conditions was formed in the first step, the quinazoline product was not achieved in the next step even if 2,3- or 2,4-dihydroxysalicylaldehyde was added in excess (Scheme S1, see ESI†).
On the other hand, the condensation of 4-aminobenzhydrazide and 2,3- or 2,4-dihydroxybenzaldehyde provided hydrazone-Schiff bases 4c and 4d only by the solution-based method (Fig. 9). The LAG route did not afford pure compounds and according to NMR, a significant quantity of reagents remained post milling. Most probably hydrazone-Schiff base molecules assemble and afterwards disassemble under mechanical force.
Fig. 9 Hydrazone-Schiff bases obtained by the reaction of 4-aminobenzhydrazide and 2,3-dihydroxybenzaldehyde, 4c, and 2,4-dihydroxybenzaldehyde, 4d. Red sphere presents OH group. |
Solutions of quinazolines and hydrazone-Schiff bases at room temperature were stable and no traces of the starting aldehydes, hydrazide or azine (vide infra) were detected as a result of the hydrolysis.60 Their absence is proven by the XRD method.
Compound 3c crystallizes as methanol solvate and its molecular structure contrasts the most among the other molecular structures. The reason emerges from the fact that during synthesis both terminal NH2 groups were condensed with 2,3-dihydroxysalicylaldehyde. The resulting molecule consists of four cyclic fragments, out of four only one is heterocyclic and non-planar, Fig. 10. Since molecules crystallize in the centrosymmetric space group and sp3 hybridized carbon atom is chiral, both stereoisomers can be found in the crystal structure. A relatively similar thing happened in the case with 4c where two salicylaldehyde molecules condensed to both terminal NH2 groups, Fig. 10. Molecules are mutually connected with an extensive hydrogen bonding (Fig. S9 and S10, see ESI†).
A comparison of reaction yields obtained by different approaches is given in Table 1. Besides high yields, the advantages of solid-state procedures also include operational simplicity and use of environmentally benign conditions.
Ligand | Synthetic procedure | ||
---|---|---|---|
Solution-based | Mechanochemistry | Melt synthesis | |
1a | 83 | >99 | >99 |
1b | 92 | — | >99 |
2a | 81 | — | >99 |
2b | 86 | >99 | >99 |
3a | 46 | >99 | — |
3b | 53 | >99 | — |
3c | 76 | >99 | — |
3d | — | >99 | — |
4a | 82 | >99 | — |
4b | 78 | >99 | — |
4c | 91 | — | — |
4d | 81 | — | — |
The proton and carbon NMR chemical shifts of all compounds in dmso-d6 solution were deduced by combined use of one (1H, 13C APT) and two-dimensional NMR techniques (COSY, HMQC and HMBC), Tables S2–S7 and Fig. S11–S22, in ESI.† The 1H NMR spectra exhibited downfield singlet signals in the range from 9.11 to 13.32 ppm attributed to NNH and phenolic OH protons, respectively. The broad OH and NH signals and chemical shift values indicated the presence of hydrogen-bonding interactions. The azomethine CHN proton and carbon signals were observed in the range from 8.29 to 9.00 ppm and 149.39 to 165.64 ppm. In the 13C NMR spectra, the low-field amide CO carbon signals appeared in the range from 160.24 to 165.40 ppm.
Additionally, singlets observed around 6.4 ppm and 5.8 ppm were assigned to –NH2 protons in the spectra of 3a–3b and 4a–4b, respectively. Quinazolines 3c and 3d showed the characteristic doublet peaks at 6.61 and 7.28 ppm, and 6.87 and 7.44 ppm, respectively in the 1H NMR spectra. They were assigned to CH and NH protons of the quinazoline ring. The CH carbon atom of the quinazoline ring was observed around 66 ppm.
Application of principal component analysis to IR-ATR spectral data obtained by reaction monitoring provided a detailed insight into the reaction profiles and the detection of reaction end. These profiles can be well represented using only the 1st principal component (Fig. 11–14).
In each case the PC1 describes more than 70% of the total variance (Table 2). These high values of explained variances in the principal component ensure the proper description of the reaction with only the first principal component. The first component picks up the most prominent changes during reactions. The reaction profiles nicely show the progress and the end of the reaction. For the vapour-mediated synthesis of 4a-3py and 4a-4py the end of the reaction can be estimated to be 100 and 90 min (Fig. 11 and 12), respectively. In the case of 4b-3py and 4b-4py these values were estimated to 70 and 120 minutes (Fig. 13 and 14). These values were confirmed by comparison of the spectra measured in these times with the spectrum of final product. Loadings vectors for monitored reactions are presented on Fig. S23–S26.† Investigation of loadings vectors confirms the creation of products during reactions.
Component | 4a-3py | 4a-4py | 4b-3py | 4b-4py |
---|---|---|---|---|
PC1 | 76.62 | 86.42 | 92.65 | 74.58 |
PC2 | 15.08 | 8.08 | 4.47 | 16.03 |
PC3 | 3.33 | 2.82 | 1.26 | 4.62 |
PC4 | 2.04 | 1.40 | 0.70 | 3.36 |
Fig. 11 Time dependence of PC1 scores calculated for a set of IR-ATR data collected through a synthesis of 4a-3py. |
Fig. 12 Time dependence of PC1 scores calculated for a set of IR-ATR data collected through a synthesis of 4a-4py. |
Fig. 13 Time dependence of PC1 scores calculated for a set of IR-ATR data collected through a synthesis of 4b-3py. |
Fig. 14 Time dependence of PC1 scores calculated for a set of IR-ATR data collected through a synthesis of 4b-4py. |
Investigation of spectral patterns confirmed that reactions were completed in the time predicted by the PC1 reaction profiles. A comparison of the NMR spectra is given in Fig. S27, see ESI†. Singlets observed around 5.8 ppm, assigned to –NH2 protons in the spectra of 4a and 4b, are missing in the spectra of the hydrazone-Schiff bases.
Fig. 15 Compounds obtained by the reaction of azine and 2,3-dihydroxybenzaldehyde, 5a, and 2,4-dihydroxybenzaldehyde, 5b. Red sphere presents OH group. |
TG-FTIR analysis of compounds 4a–4d, as well as 5a and 5b, indicates absorption bands characteristic for ammonia in the IR spectra during decomposition of the mentioned compounds caused by heating. On the other hand, the TG-FTIR spectra of the compounds 1a–3a and 1b–3b did not provide results that could be easily interpreted.
To elucidate the thermal behaviour of 5a, several series of cyclic DCS experiments were conducted. The thermograms, during several heating runs from 25 °C to 225 °C, did not exhibit any peaks and therefore phase transformations could be excluded. It is, therefore, reasonable to assume that the gradual colour changes could be caused by changes in the intermolecular interactions.61
In all crystal structures, molecules are almost planar, and a similar scenario could be assumed also for 5a. We assume that at higher temperatures, the increased unit cell volume and the distance between assembled molecules could influence the charge delocalization.
The colour change of 5b·H2O upon heating up to 300 °C was less pronounced (from yellow to light orange). This is most probably influenced by the distance between molecules in the crystal structure. Thermochromic changes were not observed for hydrazone-Schiff bases, implying the importance of the aldazine –CN–NC– unit towards the external stimuli.
Unlike the tested hydrazones (1a–4a and 1b–4b), these related hydrazones had been reported to exhibit weak to moderate cytotoxicity against THP-1 cells.57,62 Several observations can be highlighted about the factors affecting the cytotoxic activity of hydrazones on THP-1 cells. To specify, introducing the hydroxy group on position 3- and 4- of the salicylaldehyde moiety significantly lowered the cytotoxic activity. Replacing the 3- or 4-methoxy group by a hydroxy one also decreased the hydrazone cytotoxicity. In addition, the cytotoxicity of hydrazones diminished upon introduction of nitrogen into the benzhydrazide moiety regardless of whether the nitrogen is a part of the ring or the amino group.
In general, as seen from the high MIC values (Table S10, see ESI†), the investigated compounds did not inhibit the growth of Gram-positive bacteria. Regarding the Gram-negative bacteria, the majority of tested compounds (excluding the azines) exhibited a weak growth-inhibitory effect on E. coli while most of them showed moderate antibacterial activity against M. catarrhalis. However, the observed activity of the tested compounds on M. catarrhalis was substantially lower compared to the reference antibiotic azithromycin.
The only hydrazone that exhibited broad-spectrum activity was 2a. Interestingly, hydrazone 2b that differs from 2a only in the position of a hydroxy group was found to be inactive against all selected bacterial strains. The influence of the hydroxy group position on the anti-M. catarrhalis activity observed in hydrazones derived from isonicotinic, 2- or 4-aminobenzoic acid hydrazide is not as pronounced as in hydrazones 2a and 2b. Specifically, the anti-M. catarrhalis activity of 4a is four times greater than the activity of 4b while the MIC values for 1a and 1b as well as for 3a and 3b differ even less. Also, a slightly greater inhibitory effect of 4- over 3-hydroxy isomer was observed only for 2-aminobenzoylhydrazones (3a and 3b).
The higher antibacterial activity against Gram-negative bacteria of the 3-hydroxy isomer, when compared to the corresponding 4-hydroxy isomer, was established not only for hydrazones but also for all other investigated types of compounds. Moreover, the azine containing the 3-hydroxy groups (5a) was one of the most efficient tested compounds while the azine with 4-hydroxy groups (5b) did not inhibit the growth of M. catarrhalis (Table S9†).
Quinazolinones 3c and 3d and Schiff bases 4c and 4d, were generally more active against M. catarrhalis than the corresponding hydrazones (3a, 3b, 4a, and 4b, respectively).
Mild reaction conditions are necessary for the post-synthetic modification of 4-aminobenzhydrazones. Furthermore, vapour-mediated synthesis is the superior approach to convert the primary amine group into an imine functionality and to avoid transimination processes.
Mechanochemical and vapour-mediated reactions can be successfully followed by ex situ PXRD and IR-ATR methods thus avoiding the use of a solution-phase spectroscopic analysis and possibility of further reaction occurrence. Implementation of principal component analysis provided a detailed insight into the reaction profiles and detection of reaction end.
The azines exhibited temperature-responsive colour change. The central unit and position of functional groups in molecules affects their response.
Quinazolinones and Schiff bases are generally more active against M. catarrhalis than the corresponding hydrazones. Introducing the hydroxy group on position 3- and 4- of the salicylaldehyde moiety significantly lowered the cytotoxic activity.
Footnote |
† Electronic supplementary information (ESI) available: (1) Analytical and spectral data, (2) view of structures and packing diagrams, (3) crystallographic data, and (4) powder diffraction patterns. Crystallographic data sets for the structures 1a·MeOH, 1b·H2O, 3b, 3c, 4a, 4b·MeOH, 4c, and 5b·H2O. CCDC 1985275–1985283. For ESI and crystallographic data in CIF or other electronic format see DOI: 10.1039/d0ra06845d |
This journal is © The Royal Society of Chemistry 2020 |