Unnati
Patel
a,
Bhavesh
Parmar
bc,
Parth
Patel
ad,
Abhishek
Dadhania
*ab and
Eringathodi
Suresh
*bc
aDepartment of Chemical Sciences, P. D. Patel Institute of Applied Sciences, Charotar University of Science and Technology, Changa-388 421, Gujarat, India. E-mail: abhishekdadhania.bt@charusat.ac.in
bAnalytical and Environmental Science Division and Centralized Instrument Facility, CSIR-Central Salt and Marine Chemicals Research Institute, G. B. Marg, Bhavnagar-364 002, Gujarat, India. E-mail: esuresh@csmcri.res.in
cAcademy of Scientific and Innovative Research (AcSIR), Ghaziabad-201 002, India
dInorganic Materials and Catalysis Division, CSIR-Central Salt and Marine Chemicals Research Institute, G. B. Marg, Bhavnagar-364 002, Gujarat, India
First published on 30th September 2020
Two mixed ligand MOFs, {[Zn2(5NO2-IP)2(L)2](H2O)}n (ADES-1) and {[Cd2(5NO2-IP)2(L)2(H2O)4](L)(H2O)(CH3OH)6}n (ADES-2) (where 5NO2-IP = 5-nitroisophthalate and L = (E)-N′-(pyridin-3-ylmethylene)nicotinohydrazide), have been synthesized and characterized via various analytical methods, including SXRD analysis. SXRD analysis revealed that both MOFs adopt an identical two-dimensional network topology. ADES-1 with bulk phase purity has been synthesized via a conventional reflux method and its potential applications as a functional material for reversible dye adsorption and as an efficient heterogeneous catalyst for the Biginelli reaction were exploited. ADES-1 exhibited the rapid and efficient adsorption of both cationic and anionic dye molecules with a slight preference for methyl violet (86%) compared to the other dyes that were investigated. A separation study involving a mixture of dyes and the application of ADES-1 as a column filler for the separation of organic dyes from water are also demonstrated. Moreover, the catalytic performance of ADES-1 towards a multi-component Biginelli reaction to yield 3,4-dihydropyrimidin-2(1H)-ones (DHPM) derivative has been studied. Finally, studies of the utilization of ADES-1 for dye adsorption and its catalytic activity revealed good stability and reusability, enabling it to be a potential functional material, and possible dye adsorption and Biginelli reaction mechanisms are also proposed.
Dyes as industrially important molecules are among the most abundant and frequently observed effluents which are composed of hazardous organic compounds and toxic substances that incessantly pollute our natural water resources. MOFs with the characteristic features mentioned earlier have been particularly exploited as a potential material for the adsorption/separation of hazardous dye molecules and catalytic applications. MOFs are advanced porous materials and their organic linkers can be functionally decorated to tune the adsorbate–MOF supramolecular interactions for the efficient adsorptive removal of hazardous pollutants from water, including dyes. Different supramolecular interactions, such as electrostatic, stacking, hydrogen bonding, and hydrophobic interactions are commonly proposed for the adsorptive mechanism.21–30 The tuneable porosity and framework charge associated with MOFs also play an important role in the adsorption and separation of neutral, anionic or cationic dye molecules.31–36 The chemical stability of MOF in solvents is crucial for the aqueous phase adsorption of dye molecules for practical applications. On the other hand, high surface areas together with regularly distributed active sites enable MOFs to be excellent candidates for catalytic applications. Among catalytic reactions, the Biginelli reaction, the three-component reaction between aldehyde, urea, and ethyl acetoacetate, is one of the most efficient ways to synthesize dihydropyrimidinone derivatives (DHPMs). The growing interest in this reaction is mainly due to Biginelli products (DHPMs) having therapeutic and pharmacological properties.37–39
We have been involved in the design and synthesis of multidimensional MOFs by a dual ligand strategy utilizing N-donor ligands besides multi-carboxylic acids as well as their applications in the areas of sensing hazardous molecules and catalysis. Recently, we have reported some mixed ligand MOFs as heterogenous catalysts for CO2 sequestration and luminescent MOFs for the selective sensing and detection of pollutant molecules.40–43 In continuation of our ongoing research, herein we report Zn(II)/Cd(II)-based 2D MOFs, {[Zn2(5NO2-IP)2(L)2](H2O)}n (ADES-1) and {[Cd2(5NO2-IP)2(L)2(H2O)4](L)(H2O)(CH3OH)6}n (ADES-2), (where 5NO2-IP = 5-nitroisophthalate and L = (E)-N′-(pyridin-3-ylmethylene)nicotinohydrazide) synthesized by a mixed ligand strategy. Single crystals of both MOFs were harvested by a layering method and bulk phase pure compound ADES-1 was synthesized by a reflux method using the respective metal salt and ligand precursors in an appropriate stoichiometry. Both MOFs were characterized by various analytical methods, including SXRD analysis. The present study outlines the synthesis of two MOFs, their characterization, structural details, and the application of ADES-1 for dye adsorption studies in aqueous media and as a heterogeneous catalyst for the Biginelli reaction to afford dihydroprimidinone derivatives, which are significant in the pharmaceutical industry.
O bands. The difference in antisymmetric and symmetric carbonyl stretching frequencies Δν = 175 and 152 cm−1 indicates the chelating and bidentate coordination mode of carboxylate moieties in both MOFs. In an attempt to measure the thermal behaviour and stability of the MOFs, thermogravimetric analyses (TGA) were performed in the range of 25–800 °C in an inert atmosphere (Fig. S3, ESI†). The TG curve of ADES-1 exhibits a weight loss at 110 °C which corresponds to the expulsion of one lattice water molecule (cal. 3.35%, obs. 3.47%), followed by the gradual collapse of the framework at ∼330 °C. ADES-2 expelled about four methanol and one water molecules present in the lattice in the temperature range 40–150 °C (cal. 14.64%, obs. 14.75%). After the stepwise weight loss with a stable plateau in the range 200–320 °C, decomposition of the encapsulated L followed by the 2D network commences at ca. 340 °C. Surface area and porosity measurement of an activated sample of ADES-1 was performed by gas adsorption experiments at 77 K showing an adsorption–desorption isotherm with marginal hysteresis. The sorption isotherm of both materials is expressed in Fig. S4 and S5 (ESI†). The sorption isotherm of ADES-1 revealed a BET surface area of 10 m2 g−1 indicating the microporous nature of the material. The low surface area and porosity observed in the N2 isotherm may be due to the offset orientation of the alternate 2D networks in ABAB pattern effectively blocking the through channels. The pore size plot and CO2 adsorption isotherm were also measured for ADES-1 and are depicted in Fig. S6 and S7 (ESI†).
space group retaining identical framework topology and the detailed structural discussion is confined mainly to the zinc-based MOF (ADES-1). The asymmetric unit of ADES-1 is composed of one molecule each of the ligand moiety, one independent metal ion along with one water molecule as the solvent of crystallization. As depicted in Fig. 1a, the octahedral coordination around each metal ion is provided by the four carboxylate oxygens from three different 5NO2-IP ligand and pyridyl nitrogens at the axial position from two L moieties. Carboxylate oxygens constitute the square base with the Zn–O bond length ranging from 2.036(3) to 2.359(4) Å and the axial coordination with almost the same Zn–N bond length is provided by L (Zn(1)–N(1) = 2.132 (3) Å; Zn(1)–N(2) = 2.137(4) Å). As mentioned earlier, three 5NO2-IP ligands are involved in coordination with each metal ion and the 5NO2-IP moiety acts as a μ3-type linker.
Thus, 5NO2-IP is involved in the μ3-κ2,κ1,κ1 mode of coordination in which carboxylate oxygens O1 and O2 are involved in the syn–anti mode bridging the neighbouring Zn(II) metals, generating a binuclear [Zn2(COO)2] secondary building unit (SBU), which is further extended as a one-dimensional double chain by chelate coordination of carboxylate oxygens O3 and O4 in connecting the binuclear SBUs (Fig. 1b). The Zn1⋯Zn1 distance within the dimeric unit is 4.04 Å and that between the nearest dimeric clusters of the double chain is 7.55 Å. Interestingly, the nitro group from 5NO2-IP exposed on the periphery of the double chain can probably be involved in supramolecular interaction with analytes and these materials may find application in sensing and adsorption. The double chain oriented along the bc plane is doubly pillared by the N-donor ligand L, generating a double-lined two-dimensional network, as portrayed in Fig. 1c. Axial coordination of the N-donor ligand L results in the [Zn2(5NO2-IP)2]n double chains being separated by a distance of 12.43 Å. The double chain layers are oriented in an offset ABAB manner, engendering a microporous structure with interlayer voids occupied by the lattice water molecules (Fig. 1c and d). Extensive hydrogen-bonding interaction does exist within the two-dimensional network. Supramolecular interactions in the form of hydrogen bonding, involving the carboxylate/nitro group of the zinc-isophthalate chains, amide functionality of the acylhydrazone ligand L and the lattice water molecule is responsible for the offset displacement of the 2D framework. Thus, water hydrogens H8A and H8B as donors to O8 in O–H⋯O interaction with chelated carboxylate oxygen O3 and nitro oxygen O5, as well as O8 as an acceptor in N–H⋯O contact with the amide hydrogen H2A assemble a hydrogen-bonded 3D network in bridging the adjacent sheets via interlayer interactions (Fig. S8, ESI†). In addition to the above-mentioned O–H⋯O and N–H⋯O interaction, weak C–H⋯O intermolecular contact between the pyridyl hydrogen H20 from L and the nitro oxygen O6 of 5NO2-IP stabilize the 3D hydrogen-bonded supramolecular architecture. In the case of ADES-2, the topology of the MOF network is identical and can be labelled as a double-walled rectangular net, as observed in the case of ADES-1. However, the asymmetric unit of ADES-2 occupies an additional half a molecule of L, having a dynamic disorder in which the amide oxygen with half occupancy resides at two positions, imposing a center of symmetry on L (Fig. 2a). Furthermore, L present in the lattice is encapsulated strongly between adjacent [Cd2(COO)2] dimeric clusters of a double-walled rectangular net involving strong π⋯π stacking interaction between the encapsulated pyridyl rings of the lattice and pillared L. Thus, the observed π⋯π stacking distances of the encapsulated ligand moiety are C1g⋯C3g = 3.63 Å and C2g⋯C3g = 3.94 Å, where C1g, C2g and C3g are the centroids of the pyridyl rings N1C9–C13, N4C16–C20 and N6C21–C25, respectively (Fig. S9, ESI†).
The Cd⋯Cd distance observed within the dimeric SBU is 3.97 Å and between the neighbouring clusters is 7.59 Å and the pillaring distance of L with the [Cd2(5NO2-IP)2]n double chain is 14.70 Å (Fig. 2b and c). Hydrogens of the lattice water molecule O9 could not be located from the difference Fourier map and O9 acts as an acceptor by N–H⋯O and C–H⋯O contacts in bridging the neighbouring 2D nets. Thus, amide hydrogens H2C and H11C from the pyridyl ring of L are involved in the intermolecular N–H⋯O and C–H⋯O H-bonding (Fig. S10, ESI†). In addition to the above H-bonding, carboxylate oxygens O2 and O3 are making intermolecular C–H⋯O contacts with the pyridyl hydrogens from L in stabilizing the supramolecular assembly (Fig. 2d). Details of all pertinent hydrogen-bonding interaction for ADES-1 and ADES-2 with symmetry are provided in Table S3 (ESI†). It is worth mentioning here some recent reports based on mixed ligand MOFs involving acylhydrazone Schiff base and aromatic dicarboxylates and their structural and functional studies.44–46
For the adsorption and separation studies, a combination of three cationic dyes, methyl violet (MV), methylene blue (MB), and rhodamine B (RhB) and an anionic dye, methyl orange (MO), were chosen (Scheme 1). As a preliminary check on the adsorption capacity of the MOF material, 100 mg of ADES-1 was added to 10 mL of an aqueous solution of the respective dye (5 × 10−5 M) and allowed to mix thoroughly at room temperature for 3 h under dark conditions by constant stirring. All the dye molecules irrespective of their charge showed a reduction in colour intensity after the stipulated time of 3 h of dye adsorption process in the presence of ADES-1. The reduction in colour strength of the dye molecules after the adsorption process clearly revealed no specific preference for adsorption by ADES-1 based on the nature and charge of the dye molecule (Fig. 3a). All cationic dyes (MV, MB, and RhB) as well as the anionic dye (MO) were adsorbed by ADES-1, which was reflected in naked eye detection as a reduced colour intensity, with a slight preference for MV from this series. The colour of the pristine ADES-1 sample completely changed to violet, blue, orange, or pink as the dye-adsorbed material (ADES-1@dye) in the presence of the respective dyes MV, MB, MO and RhB (Fig. 3b).
UV-vis spectrometry was used in the dye adsorption experiment to monitor and evaluate the decrease in dye concentration (Fig. 3c). The observed drops in dye concentration from the initial set concentration of 5 × 10−5 M were 7.17 × 10−6 M, 9.85 × 10−6 M, 1.98 × 10−5 M, and 2.32 × 10−5 M after 3 h of constant stirring for MV, MB, MO and RhB, respectively. The decreasing adsorption capacity by ADES-1 towards the dye molecules in the present investigation is in the order MV > MB > MO > RhB and the calculated adsorption efficiencies (η) from eqn (S2) (ESI†) for MV, MB, MO and RhB are 86%, 80%, 60% and 54%, respectively, indicating no specific adsorption trend based on the charge of the dye molecule (Fig. 3d). To evaluate the amount of dye adsorbed by ADES-1 in a given time interval, a calibration curve using standard aqueous solutions of the respective dyes was implemented (Fig. S11, ESI†). The calculated adsorption amounts of the respective dyes (qe, mg g−1) from eqn (S1) (ESI†) were found to be 1.75, 1.28, 0.99 and 1.28 mg g−1 for MV, MB, MO and RhB. Upon an increase in the concentration of the dye solution (i.e. from 5 × 10−5 M to 10−4 M) the adsorption efficiency also increased with respective calculated adsorption amounts of 4.03, 3.24, 3.26 and 3.33 mg g−1. In brief, the adsorption experiments clearly indicate that the neutral framework ADES-1 is a potential candidate for the efficient removal of hazardous dye pollutants. As mentioned earlier, the colour of the off-white pristine ADES-1 changed upon dye adsorption and the materials obtained are designated ADES-1@MV, ADES-1@MB, ADES-1@MO and ADES-1@RhB. All these ADES-1@dye materials obtained were further characterized by various analytical methods, such as solid-state UV-Vis, FTIR spectra and TG analysis (Fig. S12–S15, ESI†). The morphology and structural integrity of ADES-1 before and after dye adsorption were studied by FE-SEM and PXRD analysis. As depicted in Fig. 4, FE-SEM micrographs of both pristine MOFs showed their microcrystalline nature and upon MB and MO adsorption they showed agglomeration with ADES-1 generating a spherical morphology in the ADES-1@dye composite.
Adsorption of MV and RhB by ADES-1 also revealed a morphological change with a lower agglomeration rate causing disruption of the microcrystalline nature. Preservation of the framework structure is witnessed from the PXRD data of ADES-1@dye, articulating that the MOF moiety retains its chemical and structural integrity, one of the prime requirements for an adsorption material (Fig. S15, ESI†). Adsorption and selectivity of targeted dyes by developing smart functionally decorated MOF material is a promising area of research not only of academic interest but also considering the functional utility. For example, the recovery of organic dyes from industrial waste by MOF via reversible adsorption is a step towards water preservation and environmental protection.
Among the different pools of dyes used in the present investigation, it is observed that RhB is least adsorbed (54%) by ADES-1 and a separation study involving a mixture of dyes was performed. For this experiment RhB was engaged with the other three dye molecules (MV, MB and MO) in an equimolar ratio (1
:
1, 10 mL) for separate batch studies. To this mixture of organic aqueous dye solution 100 mg of ADES-1 was added, and stirred vigorously for 3 h under dark conditions. The absorption intensity of the respective dyes MV (∼588 nm) MB (∼664 nm), MO (∼464 nm) and RhB (∼556 nm) was monitored with a UV-vis spectrophotometer from the batch of mixtures of dyes for selective adsorption by ADES-1. UV-vis spectra revealed a marginal decrease in absorption intensity of RhB (∼556 nm) but a complete/significant decrease in absorption intensity for MV and MB followed by MO from the mixture of dye solutions (Fig. 5a and b). This is also evidenced by naked eye visualization of the mixed dye solution after 3 h of continuous stirring showing a pink/light orange colour indicating the presence of RhB/RhB + trace of MO in the residual discharged solution (Fig. 5a and b). Spectroscopic and visual detection studies clearly demonstrate the efficiency of ADES-1 for the preferential adsorption of MV from the selected pool of dye molecules explored.
![]() | ||
| Fig. 5 (a) Digital photographs and (b) UV-vis spectral changes presenting the higher adsorption of MV, MB, and MO over RhB by ADES-1. | ||
Inspiration from the dye adsorption results by spectroscopic methods encouraged us to make a more practical application of ADES-1 as a column filler for the separation of organic dyes from water. Column preparation and other details are provided in the Experimental section. Almost complete adsorption of the dyes by ADES-1 was observed when 10−4 M solutions of individual aqueous dye solutions of MV, MB and MO were allowed to pass through the column while only partial adsorption by the column material had taken place in the case of RhB. This clearly echoed the colourless solution collected in the case of MV, MB and MO and the pink solution in the case of RhB after passing through the column (Fig. 6a–d).
Further, to supplement the adsorption results from spectroscopic methods of mixed-dye by ADES-1, similar column experiments were conducted by passing streams of dye combinations (RhB + MV, RhB + MB and RhB + MO). The elute collected after passing through the column showed a pale pink colour, indicating the presence of RhB in the discharge solution, reinforcing the separation as well as the capture of MV, MB and MO in the column filler (Fig. 6e–g).
A probable mechanism for dye adsorption based on the crystal structure analysis has been proposed. The low surface area and porosity of ADES-1 ruled out the possibility of dye molecule adsorption by encapsulation. However, the layered gap between the offset-stacked 2D network can probably accommodate the dye molecule involving various supramolecular as well as electrostatic interactions on the surface of the MOF.51
Crystal structure analysis of ADES-1 revealed that the polar nitro group from 5NO2-IP exposed on the periphery of the double chain can create an anionic cloud and is probably involved in electrostatic interaction with the cationic dyes in the adsorption process. Higher adsorption of MV (86%) and MB (80%) compared to RhB (54%) can be justified due to the diffusion of the bulky RhB being too low to make an effective electrostatic interaction. Interestingly, good adsorption of anionic dye MO (60%) can be attributed to the hydrogen-bonding interaction between the amide hydrogen from L with the sulfonate oxygen of the MO on the surface of the ADES-1 host by accommodating the dye molecule between the 2D layers. Since the adsorbent and adsorbate possess an aromatic moiety, π⋯π stacking also probably plays an important role in the adsorption process. A schematic representation of the probable mechanism for cationic and anionic dyes is depicted in Fig. 7. To reinforce this claim we further performed zeta potential measurements of ADES-1 in aqueous solution. The negative zeta potential (−5.8 mV) observed for ADES-1 in aqueous solution clearly demonstrates that strong electrostatic attractions between cationic dye molecules and the highly negative charged surface of ADES-1 play a vital role in the adsorption process. The desorption process is equally important in the case of reversible adsorption of hazardous molecules as a functional material for practical applications. In the present investigation, dye release experiments were performed by soaking the dye-adsorbed materials ADES-1@dye in methanol as a dye-release medium for 90 minutes. Visual images of the progressive release of MB to methanol and the enhancement in the intensity at 652 nm corresponding to MB are shown in Fig. 8a–c. Release studies performed in a similar way for other dyes also revealed the steady release of the dye molecules over time, as shown in digital photographs (Fig. S16–S18, ESI†).
![]() | ||
| Scheme 2 The Biginelli reaction of aldehyde derivatives with ethyl acetoacetate and urea to produce 3,4-dihydropyrimidin-2(1H)-one (DHPM) derivatives. | ||
To optimize the reaction conditions, the condensation of benzaldehyde, ethyl acetoacetate and urea was chosen as a model reaction. A series of experiments was conducted to examine the influence of different parameters like the amount of catalyst, temperature, solvent and reaction time. The obtained results are summarized in Table 1 for the model reaction. Initially the model reaction was performed at 80 °C under solvent-free conditions with a catalyst loading of 0.005 mmol and the reaction was completed with poor conversion (22.8%) to yield ethyl 6-methyl-2-oxo-4-phenyl-1,2,3,4-tetrahydropyrimidine-5-carboxylate within 10 min (Table 1, entry 1).
| Entry | Catalyst (mmol) | Temp. (°C) | Solvent | Conversion (%) | Time (min) |
|---|---|---|---|---|---|
| a Reaction conditions: benzaldehyde (1.0 mmol), ethyl acetoacetate (1.0 mmol), urea (1.5 mmol), solvent (2.0 mL). b With catalyst ADES-2. | |||||
| 1 | 0.005 | 80 | None | 22.8 | 10 |
| 2 | 0.010 | 80 | None | 73.2 | 10 |
| 3 | 0.015 | 80 | None | 93.6 | 10 |
| 4 | 0.020 | 80 | None | 98.3 | 10 |
| 5 | 0.025 | 80 | None | 93.2 | 10 |
| 6 | 0.020 | 60 | None | 86.5 | 20 |
| 7 | 0.020 | 70 | None | 91.3 | 10 |
| 8 | 0.020 | 90 | None | 98.1 | 10 |
| 9 | 0.020 | 80 | CH2Cl2 | 82.2 | 10 |
| 10 | 0.020 | 80 | Toluene | 76.2 | 120 |
| 11 | 0.020 | 80 | EtOH | 86.7 | 60 |
| 12b | 0.020 | 80 | None | 97.4 | 10 |
To check the effect of catalyst loading, four different experiments were performed with 0.010, 0.015, 0.020 and 0.025 mmol of catalyst under similar conditions (Table 1, entries 2–5). In the presence of 0.020 mmol of catalyst, the conversion observed was 98.3% within the set reaction time of 10 min and a further increase in catalytic amount tended to decrease the conversion. Further, we also examined the influence of temperature and different solvents like CH2Cl2, toluene and EtOH, while keeping the catalyst amount at 0.020 mmol. Notably, at temperatures lower than 80 °C or in the presence of organic solvents the reaction advanced with a reduction in conversion and employing a longer reaction time (Table 1, entries 6–11). A representative reaction by ADES-2 for the model substrate benzaldehyde with ethyl acetoacetate and urea was conducted, which showed similar efficiency (97.8%) for the product conversion (Table 1, entry 12). From the preliminary optimization reaction, the following conditions were set for the product scope with solvent-free reaction using 0.020 mmol of the ADES-1 catalyst at 80 °C with the substrate concentration mentioned in Table 1. To check the scope and generality of the ADES-1 catalyst for the Biginelli reaction, the condensation of different substituted aromatic aldehydes with ethyl acetoacetate and urea was performed under the set optimized reaction conditions and the obtained results are abridged in Table 2. Indeed, all the aldehydes with electron withdrawing and donating groups gave a good product yield within 8 to 12 min. The formation of the desired Biginelli product, dihydropyrimidinone (DHPMs), was confirmed by 1H NMR and 13C NMR spectroscopy (Fig. S19–S32, ESI†). The ADES-1 catalyst showed good recyclability up to 6 cycles showing no appreciable reduction in product yield (Fig. S33, ESI†). After each cycle, the catalyst was separated by filtration and washed with methanol/water (1/1) followed by acetone and dried to regenerate the active catalyst. The recovered catalyst after six cycles revealed the integrity of the framework, showing identical PXRD and FTIR patterns to the pristine sample (Fig. S34 and S35, ESI†). A plausible mechanism for the Biginelli reaction can be elucidated on the basis of the formation of an N-acyliminium ion intermediate similar to the mechanism proposed by Kappe.63 The crystal structure of the ADES-1 catalyst revealed the presence of weakly coordinated chelated carboxylate which can open up and act as a Lewis acidic site and the existence of amide decorated L as a Lewis basic site can probably concomitantly activate the substrates in the formation of an N-acyliminium ion intermediate from the reaction between aldehyde and urea. Subsequently, nucleophilic attack on the N-acyliminium ion by the enol tautomer of ethyl acetoacetate yielded ureide. In the last step, the ureide intermediate undergoes cyclodehydration to generate the desired DHPM product. Therefore, ADES-1 acts as an efficient heterogeneous catalytic material exhibiting an excellent product yield for both electron donating and withdrawing aromatic substrates in the Biginelli reaction at 80 °C under solvent-free mild reaction conditions, which is comparable with some recently reported MOF catalysts.64–68
Footnote |
| † Electronic supplementary information (ESI) available. CCDC 2014940 and 2014941. For ESI and crystallographic data in CIF or other electronic format see DOI: 10.1039/d0qm00611d |
| This journal is © the Partner Organisations 2021 |