Next Article in Journal
Gentiana lutea Extract Modulates Ceramide Synthesis in Primary and Psoriasis-Like Keratinocytes
Next Article in Special Issue
Proton Conductive Zr-Phosphonate UPG-1—Aminoacid Insertion as Proton Carrier Stabilizer
Previous Article in Journal
Mulberry Fruit Cultivar ‘Chiang Mai’ Prevents Beta-Amyloid Toxicity in PC12 Neuronal Cells and in a Drosophila Model of Alzheimer’s Disease
Previous Article in Special Issue
Metal-Organic Framework-Based Engineered Materials—Fundamentals and Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Coordination Polymers Based on a Biphenyl Tetraphosphonate Linker: Synthesis Control and Photoluminescence

by
Ana D. G. Firmino
1,2,†,
Ricardo F. Mendes
1,†,
Duarte Ananias
1,3,
Jéssica S. Barbosa
1,2,
João P. C. Tomé
2,4 and
Filipe A. Almeida Paz
1,*
1
Department of Chemistry, CICECO—Aveiro Institute of Materials, University of Aveiro, 3810-193 Aveiro, Portugal
2
QOPNA and LAQV-REQUIMTE, Department of Chemistry, University of Aveiro, 3810-193 Aveiro, Portugal
3
Department of Physics, CICECO—Aveiro Institute of Materials, University of Aveiro, 3810-193 Aveiro, Portugal
4
CQE and Departamento de Engenharia Química Instituto Superior Técnico, Universidade de Lisboa Av. Rovisco Pais, no 1, 1049-001 Lisboa, Portugal
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Submission received: 25 March 2020 / Revised: 9 April 2020 / Accepted: 15 April 2020 / Published: 16 April 2020
(This article belongs to the Special Issue Functional Metal-Organic Framework Based Materials)

Abstract

:
In this work, we used the rigid tetrapodal organic linker, [1,1′-biphenyl]-3,3′,5,5′-tetrayltetrakis(phosphonic acid) (H8btp), for the preparation of two lanthanide–organic framework families of compounds: layered [Ln7(H5btp)4(H5.5btp)2(H6btp)2(H2O)12]∙23.5H2O∙MeOH [where Ln3+ = Eu3+ (1Eu) and Gd3+ (1Gd)], prepared using microwave-irradiation followed by slow evaporation; 3D [Ln4(H3btp)(H4btp)(H5btp)(H2O)8]∙3H2O [where Ln3+ = Ce3+ (2Ce), Pr3+ (2Pr), and Nd3+ (2Nd)], obtained from conventional hydro(solvo)thermal synthesis. It is shown that in this system, by carefully selecting the synthetic method and the metal centers, one can increase the dimensionality of the materials, also increasing structural robustness (particularly to the release of the various solvent molecules). Compound 1 is composed of 2D layers stacked on top of each other and maintained by weak π–π interactions, with each layer formed by discrete 1D organic cylinders stacked in a typical brick-wall-like fashion, with water molecules occupying the free space in-between cylinders. Compound 2, on the other hand, is a 3D structure with small channels filled with crystallization water molecules. A full solid-state characterization of 1 and 2 is presented (FT-IR spectroscopy, SEM microscopy, thermogravimetric studies, powder X-ray diffraction and thermodiffractometry). The photoluminescence of 1Eu was investigated.

1. Introduction

Metal-Organic Frameworks (MOFs), or Porous Coordination Polymers (PCPs), are a class of crystalline materials built from the coordination of organic linkers to metal ions or clusters [1]. Among the assorted reasons why there are so many reports in the area of MOFs and PCPs, and why they remain an attractive topic of research, one can point out their tunable chemical structure and/or their multifunctional properties [2]. The ability to design a network with specific pore sizes and shapes is one of the main features why these materials are good candidates for multiple applications in different research areas like gas storage and separation [3,4], catalysis [5,6], sensing [7,8,9,10,11], and drug delivery [12,13], among others [14,15,16,17,18,19,20,21].
While the main research regarding MOFs remains focused on materials derived from d-block metals coordinated to carboxylate-based organic linkers, our research group has focused on metal phosphonates, more specifically, lanthanide phosphonate-based MOFs (LnOFs). LnOFs have gained particular interest, in some part due to the possibility to design new materials by changing the lanthanide radii. In some particular cases, the LnOFs properties can also change [22,23,24]. Recently, Wang and co-workers were able to improve the DNA sensing of LnOF nanosheets by simple change in lanthanide ions [25]. The contraction of ionic radii led to a lower quenching efficiency. Tetraphosphonic organic linkers have gained particular interest in recent years, especially phosphonic acids based on tetraphenylmethane and tetraphyenylsilane building units. [26] While their geometries can pose some challenges regarding its synthetic approaches (mainly due to the high rigidity that these molecules present), it results in their properties improvement. A particular improvement is the stability of these materials: while many MOFs show low stability in even ambient conditions, some of these materials can withstand solutions of concentrated acids (HCl and H2SO4) and even aqua regia [27]. In our group, the rich chemistry of phosphonic acid allowed the preparation of a variety of structures ranging from 1D to 3D networks by, in many cases, slight changes in experimental conditions. With this we were able to synthetize MOFs with a broad range of applications: from photoluminescence [28,29,30], to luminescent thermometers [31], heterogeneous catalyst [32,33] and/or proton conductors [34].
In previous reports we presented and explored the phosphonic acid [1,1′-biphenyl]-3,3′,5,5′-tetrakis(phosphonic acid) (H8btp) as a new tetraphosphonic organic ligand (Scheme 1) [28,31,34]. H8btp afforded a particular series of MOF-type compounds self-assembled from La3+ cations (or mixtures of lanthanides), which then showed not only fascinating structural features, but interesting photoluminescent properties. We also registered, however, that these 2D materials presented a significant drawback (which might also be seen as a disadvantage regarding a future practical application): their thermal stability was very low with the framework being only stable between ambient temperature and ca. 70 °C. In order to try to overcome this disadvantage, we have further explored the chelating ability of H8btp under different synthetic conditions, including a careful selection of the lanthanide radii, and isolated two new isotypical families of LnOFs which we report here:
(i)
layered [Ln7(H5btp)4(H5.5btp)2(H6btp)2(H2O)12]∙23.5H2O∙MeOH [where Ln3+ = Eu3+ (1Eu) and Gd3+ (1Gd)] (prepared using the same self-assembly methodology previously reported for the materials based on La3+ and H8btp, i.e., microwave-irradiation followed by slow evaporation);
(ii)
3D [Ln4(H3btp)(H4btp)(H5btp)(H2O)8]∙3H2O [where Ln3+ = Ce3+ (2Ce), Pr3+ (2Pr), and Nd3+ (2Nd)], obtained from conventional hydro(solvo)thermal synthesis. This approach allowed us to obtain a material with increased dimensionality (3D) in a much less synthesis time.

2. Results and Discussion

2.1. Crystal Structure Details

2.1.1. [Ln7(H5btp)4(H5.5btp)2(H6btp)2(H2O)12]∙23.5H2O∙MeOH (1)

The structure of 1 was unveiled while studying single-crystals of [Eu7(H5btp)4(H5.5btp)2(H6btp)2(H2O)12]∙23.5H2O∙MeOH (1Eu) using X-ray diffraction. The isotypical nature of 1Eu and 1Gd was unequivocally confirmed by powder X-ray diffraction data (Figures S1 and S2 in the Supporting Information), being supported by EDS mapping (Figures S8 and S9 in the Supporting Information). The family crystallizes in the monoclinic space group P21/c, with a large asymmetric unit composed of three and half metal centers and four different H8−xbtpx residues (x = 2, 2.5 and 3). Structural adhesion is overall achieved by a complex network of hydrogen bonding interactions between the phosphonate groups and the several solvent molecules.
The four metal centers have different coordination numbers as depicted in Figure 1. Eu1 is at a center of inversion, with multiplicity of 2, being hexacoordinated to six phosphonate groups, while Eu2 is heptacoordinated to five different phosphonate groups and two coordination water molecules (O2W and O3W are disordered). The remaining two metal centers, Eu3 and Eu4 are also heptacoordinated to two coordination water molecules and five phosphonate groups, each. For the former two coordination environments, the phosphonate groups coordinate with the metal centers with the same k1-O mode of coordination while for the last two, besides this mode of coordination, two phosphonate groups have a k2-O coordination mode. The Eu–O bond lengths were found in the 2.221(7)–2.535(16) Å range, comparable to those reported for other Ln3+-based phosphonate compounds, and the internal O–La–O polyhedral angles are in the 69.5(4)–180.0(2)° range (Table S1 in the Supporting Information).
The phosphonic acid groups have three different protonation degrees: −3, −2.5 or −2. The partial protonation arises from the need to balance the overall crystal charge, and the location of the respective hydrogen atom was carefully selected and placed on O45 (based on proximity of hydrogen bonding acceptor. We further note that the structure also has hydrogen atoms which “jump” between adjacent phosphonate groups, namely between O23 and O30 and between O33 and O38.
As observed for the previously reported [Ln4(H6btp)2(H4btp)2(H8btp)(H2O)16]∙12H2O isotypical family of networks [where Ln3+ = La3+, (La0.9Eu0.1)3+ and (La0.9Tb0.1)3+] [28], 1 contains discrete 1D organic cylinders stacked in a typical brick-wall-like fashion in the bc plane, with solvent molecules occupying the available space in-between these cylinders as depicted in Figure 2. Connections between the organic cylinders are ensured by metallic dimers (inset in Figure 2a), leading to 2D layers placed in the ac plane of the unit cell, stacked on top of each other and maintained by weak π–π interactions between the aromatic rings of adjacent layers [distance of 3.982(6)Å and dihedral angle of 13.3(5)°].
Within these cylinders the organic linkers are stacked on top of each other with an overall rotation angle of ca. 135° along the [100] direction, forming a double-helix-like structural motif as depicted in Figure 3. Besides the connection to the metal centers, each organic residue forms weak π–π interactions between each other, with intercentroid distances ranging from 3.5805(4) to 3.8321(5) Å [dihedral angles between 4.8(4) and 6.1(5)°].
The presence of several phosphonate groups and coordinated and crystallization solvent molecules promote the formation of a complex network of hydrogen-bonding interactions. Within the organic cylinders, such interactions are of the P‒O⋯H‒O‒P type (i.e., inter-phosphonate groups) being rather strong and directional [dD⋯A distances between 2.496(13) and 3.263(13)Å with <(DHA) between 125–176°]. These ensure a close packing of the 2D layers and overall structural integrity of 1 (for more details see Table S2 in the Supporting Information).
It is interesting to note that the four independent organic linkers composing the aforementioned organic cylinder connect, each, to a different number of metal centers (3, 4, 5 and 6), ultimately leading to the 2D compact network which can be more easily analyzed from a pure topological perspective. Figure S13 (in the Supporting Information) depicts the reduction of the network to simple nodes (i.e., the metal centers) and bridges (ensured by the organic linkers). Following the recommendations of Alexandrov at al. [35], who suggested that any moiety (ligand or atoms) connecting more than two metallic centers should be considered as a network node. For this case, both the metal centers and the organic linkers were considered as network nodes. We note that the hydrogen bonds were not included in these calculations. The network of the isotypical family of 1 is, therefore, an octanodal 3,4,5,5,5,5,6,6-connected network with total point symbol {413·62}2{43}2{45·63·82}6{46·88·10}{46}2{49·6}2 as revealed by the software package TOPOS [36]. Database review of the Reticular Chemistry Structure Resource (RCSR) [37] and in EPINET [38] show that this complex topology has not been described yet. We note that the network was attempted to be reduced to more simple nodes such as those described in the paper by Bonneau at al. [39] However, this approach did not significantly reduce the number of nodes ultimately leading instead to a mixture of different node approaches, hence in this manuscript we decided to maintain the classical approach. For this case, the notations {43}, {49·6}, {46} correspond to the three H8−xbtpx residues and {413·62}, {45·63·82}, {46·88·10} corresponds to the metal centers.

2.1.2. [Ln4(H3btp)(H4btp)(H5btp)(H2O)8]∙3H2O (2)

The structure of the isotypical family of compounds [Ln4(H3btp)(H4btp)(H5btp)(H2O)8]∙3H2O [where Ln3+ = Ce3+ (2Ce), Pr3+ (2Pr), and Nd3+ (2Nd)] was elucidated for the La3+-based material. Besides elemental analysis (see Experimental Section), the homogeneity of the bulk samples was studied using powder X-ray diffraction (Figures S1 and S3 in the Supporting Information), FT-IR spectroscopy (Figure S7 in the Supporting Information), and electron microscopy (Figures S10 to S12 in the Supporting Information).
Compound 2 crystallizes in the monoclinic space group P21/c, with the asymmetric unit being composed of four metal centers and three different H8−xbtpx residues (x = 5, 4 and 3). The four metal centers have different coordination numbers (Figure 4): La1 and La2 are both octocoordinated to five phosphonate groups and three coordination water molecules; La3 is nonacoordinated to seven different phosphonate groups and two coordination water molecules; La4 is heptacoordinated to six different phosphonate groups and one coordination water molecule. As expected, and compared with the structure of 1, the overall coordination numbers are higher, thus reflecting the larger ionic radius of the metal centers. The La–O bond lengths were found in the 2.346(5)–2.851(6) Å range while the internal O–La–O polyhedral angles are in the 53.61(15)–59.40(15)° range (see Table S3 in the Supporting Information).
When compared to the previously described layered material, 2 has a smaller number of organic linkers in the asymmetric unit: a total of three, showing distinct protonation degrees −5, −4 and −3. The overall structural connectivity of each linker is distinct: one acts as a hexadentate bridge, connecting six metal centers, while the remaining residues act as a hepta- and octadentate linkers, connecting seven and eight metal centers each. For the former case, the phosphonate groups coordinate to the metal centers by simple k1-O coordination modes, while in the latter two situations there are also k2-O and μ-O,O modes of coordination involved. It is noteworthy to mention that when compared to the structural features of the organic linkers in 1, one can easily infer that the bridges can easily accommodate more metal centers in the structure by increasing the overall connectivity and having, in average, a higher level of deprotonation.
The network of 2 is also composed by organic cylinders distributed in typical brick-wall-like fashion in the ab plane and formed by stacked organic linkers along [001] direction of the unit cell (Figure 5). The organic linkers are stacked, as depicted in Figure 6, on top of each other with a rotation angle of ca. 92° along the [100] direction, forming a double-helix-like structural motif, just as observed in the other related materials. As so, besides the connection with the metallic centers (described below), the organic linkers are further engaged in weak π–π interactions with intercentroid distances ranging between 3.5545(5) and 3.9692(5) Å [dihedral angles between 4 and 9°] (Figure 6).
As for 1, these cylinders are mutually connected by metallic dimers. However, unprecedented tetrameric units are also present for this family of compounds (insets in Figure 5a). Intermetallic distances: for the dimer, La2⋯La2 of 6.0643(6) Å; for the tetramer, La1⋯La3 of 4.0834(4) Å and 4.4179(5) Å. These metallic connections (covalent bonds), allied to very strong and directional hydrogen-bonding interactions between phosphonate groups [dD⋯A distances between 2.521(7) and 2.842(8)Å with <(DHA) between 111 and 172°—see Table S4 in the Supporting Information], are the main reason for the formation of the three-dimensional network of the MOF structure 2 as shown in Figure 5b.
Overall, this family of compounds has a lower number of both coordinated and crystallization solvent molecules. We note that these are also engaged in strong hydrogen bonding interactions with the phosphonate groups lining the inner walls of the channels of the network [dD⋯A distances between 2.553(9) and 3.013(8)Å with <(DHA) between 131 and 161°—see Table S4 in the Supporting Information].
Using the same topological strategy previously described for 1 (again, the hydrogen bonds were not included in the calculation), this new MOF structure is a new hexanodal 5,6,6,7,7,8-connected network with total point symbol {3·412·62}{32·410·69}{34·416·56·62}{37·411·53}2{44·5·6·89}{48·62} (see Figure S14 in the Supporting Information), wherein {3·412·62},{ 32·410·69},{ 34·416·56·62} and {37·411·53}, {44·5·6·89}, {48·62} notations correspond to the metal centers and H8−xbtpx residues, respectively. We note that in this framework La1 and La3 exhibit the same point and extended symbol being, thus, equivalent hence the six-connectivity nature of the network.

2.2. Structural Characterization

2.2.1. Thermogravimetry

The thermal stability of [Gd7(H5btp)4(H5.5btp)2(H6btp)2(H2O)12]∙23.5H2O∙MeOH (1Gd) was investigated between ambient temperature and ca. 800 °C, providing additional information on the hydration level of this family of isotypical compounds. As depicted in Figure S4, the thermogram for compound 1Gd shares many similarities with that of [La4(H6btp)2(H4btp)2(H8btp)(H2O)16]∙12H2O [28], ultimately indicating a comparable thermal behavior. For 1Gd the thermogram does not evidence unequivocal regions in which the mass remains stable. There is a continuous weight loss (more significant at some specific temperatures) over the entire temperature range. This behavior creates additional difficulties for individual assignments of temperature range vs. released residues, although it was foreseeable that this should happen due to the large number of coordination and crystallization solvent molecules present in the structure. One can, nevertheless, confirm the crystallographic studies for this compound. It is possible to discern three main weight losses: the first, between ambient temperature and ca. 100 °C, corresponding to 8.2% of the total weight loss, is attributed to the release of all crystallization solvent molecules from the structure (calculated 8.2%), i.e., the 23.5 water molecules plus the one methanol molecule. Above this temperature, there is an immediate second weight loss until ca. 200 °C. In this temperature range, the coordination water molecules are removed from the structure corresponding to a total weight loss of 4.3% (calculated 3.9%). The final step corresponds to the decomposition of the organic component.
The thermal behavior of the bulk [Ln4(H3btp)(H4btp)(H5btp)(H2O)8]∙3H2O (2) materials were also investigated between ambient temperature and ca. 800 °C. Due to the isotypical nature of the compounds the following paragraph discussion will be solely focused on compound 2Ce (see Figure S5 in the Supporting Information). For 2Ce, one discerns three main weight losses. The first, between ambient temperature and ca. 110 °C, corresponding to 2.4% of the total weight loss, is attributed to the release of all crystallization solvent molecules (calculated 2.4%), i.e., a total of 3 water molecules. We emphasize that this value agrees well with the water content found from the performed crystallographic studies. Above this temperature, there is a second weight loss until ca. 250 °C. In this temperature range, the coordination water molecules are removed from the structure (in this case the observed 6.6% weight loss is equivalent to eight water molecules). The final step corresponds to the decomposition of the organic content.
Variable-temperature powder X-ray diffraction studies on [Ce4(H3btp)(H4btp)(H5btp)(H2O)8]∙3H2O (2Ce) further corroborate the aforementioned assumptions concerning the release of water molecules and phase modifications (Figure S5 – top – see the Supporting Information): from ambient temperature to ca. 300 °C the structure changes, leading to the formation of a new crystalline phase as a consequence of the release of crystallization and coordination water molecules. The crystalline structure starts to collapse above this temperature becoming almost fully amorphous at 550 °C.
While these compounds share many structural similarities, the MOF compound 2 shows superior framework stability when compared to the previously reported [Ln4(H6btp)2(H4btp)2(H8btp)(H2O)16]∙12H2O [28] and the herein reported [Ln7(H5btp)5(H6btp)3(H2O)12]∙23.5H2O∙MeOH (1) materials that became both amorphous at rather low temperatures. We attribute this feature to the increased connectivity between metal ions and organic linkers: in 2 the structure contains less solvent molecules, both coordinated and of crystallization, and a higher connectivity of the phosphonate groups; thus, the liberation of the solvent molecules, including those which may be coordinated to the metal centers and pointing towards the small channels of the network, do not promote such an earlier onset decomposition of the material, with crystalline material being observed at temperature as high as 450 °C (see inset in Figure S5 in the Supporting Information).

2.2.2. FT-IR Spectroscopic Studies

The vibrational FT-IR spectroscopy studies support the structural features unveiled by the X-ray diffraction studies. Figures S5 and S6 (in the Supporting Information) show the FT-IR spectra of the materials in the 3650–350 cm−1 spectral region, including assignments for each of the main observed bands. Because 1 and 2 have the same functional groups, the spectra contain the same vibrational modes. A broad band is centered in the ca. 3600–2550 cm−1 spectral range, being attributed to both ν(O–H) stretching vibrational modes from coordination and crystallization water molecules and to the ν(PO–H) stretching vibration of the phosphonate groups. In the central region of the spectrum, between ca. 1745 and 1560 cm−1, it is possible to observe the typical ν(C=C) stretching vibrational modes arising from the aromatic rings and the deformation stretching of the water molecules. We further observe a sharp vibrational mode at ca. 695 cm−1, corresponding to the ν(P–C) stretching vibration. Also, in this region, the stretching vibrational modes of ν(P=O) were noticed from ca. 1300–1060 cm−1, and those of ν(P–O) from ca. 1070 to 830 cm−1.

2.3. Photoluminescence Studies

The excitation spectra of 1Eu were recorded at ambient temperature (ca. 300 K) and 12 K monitoring the strongest Eu3+ 5D07F2 emission transition (Figure 7). Both spectra are dominated by a broad UV band (240–320 nm) attributed to π–π* transitions of the organic ligands, which are very similar to the one previous reported for the Ln3+ MOFs obtained with the H8btp organic linker [28,29,30,31]. The additional sharp lines in the spectra of 1Eu are ascribed to the intra-4f6 7F0,15D1-4, 5L6, 5G2-6 and 5H3-7 transitions of Eu3+. The strong improvement of the 240–320 nm broad band, when the temperature decreases, demonstrates a non-effective energy transfer from the ligand to the Eu3+ ions (antenna effect) at the ambient temperature.
The emission spectra of 1Eu recorded at ambient temperature and 12 K, excited at 393 nm are presented in Figure 8. The sharp Eu3+ emission lines are assigned to the 5D07F0-4 transitions. The 5D17F0-3 transitions are only properly noticed, even at 12 K, after magnification (Figure S15 in the Supporting Information), demonstrating an efficient relaxation between the 5D1 and 5D0 excited states promoted by the vibrations of the organic linkers and the coordinated water molecules. The dominance of the 5D07F2 over the 5D07F1 transition is typical of Eu3+ environments without inversion center. Although, the structure of 1Eu presents four independent Eu3+ crystal sites, one of which in an inversion center, the corresponding emission spectra do not allow discriminating the corresponding individual photoluminescence signatures. Nevertheless, changing the excitation wavelength from 393 to 310 nm at 12 K (inset of Figure 8) results on a substantial change on the emission profile, particularly on the 5D07F1 transition region. The emission spectrum obtained with the 310 nm excitation, in addition to the three lines of the 393 nm spectrum shows four additional Stark components, which is in good agreement with the presence of multiple Eu3+ local sites. In accordance, the 5D0 decay curve of 1Eu under direct excitation at 393 nm (5L6 excited level), while monitoring the strongest 7F2 Stark component at ambient temperature (Figure S16 in the Supporting Information), cannot be properly described by a single-exponential function. The data is only well fitted by a second order exponential function yielding lifetimes of 0.39±0.01 and 1.16±0.10ms, with an average lifetime of 0.52 ms (details in the Supporting Information).

3. Materials and Methods

3.1. Synthesis of [Ln7(H5btp)4(H5.5btp)2(H6btp)2(H2O)12]∙23.5H2O∙MeOH (1)

Reactive mixtures composed of the respective lanthanide(III) chloride hydrates [LnCl3∙6H2O, where Ln3+ = Eu3+ (1Eu) and Gd3+ (1Gd)] and 0.025 g of [1,1′-biphenyl]-3,3′,5,5′-tetrayltetrakis(phosphonic acid) (H8btp), with an overall molar ratio of approximately 1:4 (H8btp: Ln3+), were individually prepared at ambient temperature in a mixture of distilled water, HCl (6 M) and methanol (2 mL each solvent), and placed inside a 10 mL IntelliVent microwave reactor. The reaction took place inside a CEM Focused Microwave Synthesis System Discover S-Class equipment, under constant magnetic stirring (controlled by the microwave equipment), at 100 °C for 60 min (using an irradiation power of 50 W). A constant flow of air (ca. 20–30 psi of pressure) ensured a close control of the temperature inside the reactor. After irradiation, the reactors with the homogeneous solutions were uncapped and left motionless until ca. 3/4 of the reaction mixture volume had evaporated. This methodology successfully led to the preparation of [Ln7(H5btp)4(H5.5btp)2(H6btp)2(H2O)12]∙23.5H2O∙MeOH (1Eu and 1Gd) materials as white crystalline plates. The resulting compounds were recovered by vacuum filtration, washed with copious amounts of distilled water and then air-dried at ambient temperature.
For [Gd7(H5btp)4(H5.5btp)2(H6btp)2(H2O)12]∙23.5H2O∙MeOH
Elemental CH composition (%): Calcd: C 21.0; H 3.01. Found: C 19.9; H 3.02.
Thermogravimetric analysis (TGA) data (weight losses in %) and derivative thermogravimetric peaks (DGT; in italics inside parentheses): 25–100 °C (89 °C) −8.2%; 100–200 °C (179 °C) −4.3%; 200–440 °C (410 °C) −7.4%; 440–800 °C (503 °C) −20.1%
Selected FT-IR data (in cm−1; from KBr pellets): ν(H2O + POH) = 3600–2800br; ν(C=C) + δ(H2O) = 1745–1560m; ν(P=O) = 1300–1070vs; ν(P–O) = 1070–830vs; ν(P–C) = 695m.

3.2. Synthesis of [Ln4(H3btp)(H4btp)(H5btp)(H2O)8]∙3H2O (2)

Reactive mixtures composed of the respective lanthanide(III) chloride hydrates [LnCl3∙6H2O, where Ln3+ = Ce3+ (2Ce), Pr3+ (2Pr), and Nd3+ (2Nd)] and 0.025 g of [1,1′-biphenyl]-3,3′,5,5′-tetrayltetrakis(phosphonic acid) (H8btp), with an overall molar ratio of approximately 1:4 (H8btp: Ln3+), were individually prepared in a mixture of distilled water, HCl (6 M) and methanol (2 mL each solvent). Mixtures were kept under constant magnetic stirring in open air and ambient temperature for approximately 15 min. The resulting homogeneous suspensions were transferred to Teflon-lined Parr Instrument reaction vessels and placed inside an MMM Venticell oven. The heating program included:
(i)
heating during 48 h up to 140 °C;
(ii)
24 h uphold of the temperature (140 °C);
(iii)
cooling down during 48 h until ambient temperature.
The resulting materials were isolated as white microcrystalline powders, recovered by vacuum filtration, washed with abundant amounts of distilled water and dried at ambient temperature. Single-crystals of this compound were obtained from different experimental conditions (denoted as 2La—see the Supporting Information for further details).
Elemental CH composition (%):
Calcd for 2Ce: C 19.9; H 2.42. Found: C 21.0; H 2.47.
Calcd for 2Pr: C 19.9; H 2.41. Found: C 20.7; H 2.37.
Calcd for 2Nd: C 19.8; H 2.40. Found: C 20.7; H 2.41.
Thermogravimetric analysis (TGA) data (weight losses in %) and derivative thermogravimetric peaks (DTG, in italics inside the parentheses):
2Ce: 25–200 °C −18.05% (110 °C); 200–400 °C −5.98% (271 °C). Total loss: 24.0%.
2Pr: 25–122 °C –10.13% (109 °C); 122–200 °C –5.60% (135 °C); 200–294 °C –6.05% (270 °C); 294–400 °C –6.51% (324 °C). Total loss: 28.3%.
2Nd: 25–133 °C −12.78% (108 °C); 133–205 °C –5.92% (144 °C); 205–367 °C −6.58% (281 °C). Total loss: 25.3%.
Selected FT-IR data (in cm−1; from KBr pellets):
2Ce: ν(H2O + POH) = 3600–2550br; ν(C=C) + δ(H2O) = 1700–1570m; ν(P=O) = 1270–1060vs; ν(P–O) = 945–840vs; ν(P–C) = 696m.
2Pr: ν(H2O + POH) = 3610–2570br; ν(C=C) + δ(H2O) = 1715–1540m; ν(P=O) = 1245–1070vs; ν(P–O) = 945–850vs; ν(P–C) = 696m.
2Nd: ν(H2O + POH) = 3625–2565br; ν(C=C) + δ(H2O) = 1695–1560m; ν(P=O) = 1245–1070vs; ν(P–O) = 940–850vs; ν(P–C) = 696m.

3.3. Single-Crystal X-ray Diffraction Studies

Single crystals of compounds [Eu7(H5btp)4(H5.5btp)2(H6btp)2(H2O)12]∙23.5H2O∙MeOH (1Eu) and [La4(H3btp)(H4btp)(H5btp)(H2O)8]∙3H2O (2La) were manually harvested from the crystallization vials and immersed in highly viscous FOMBLIN Y perfluoropolyether vacuum oil (LVAC 140/13, Sigma-Aldrich) to avoid degradation caused by the evaporation of the solvent [40]. Crystals were mounted on either Hampton Research CryoLoops or MiTeGen MicroLoops, typically with the help of a Stemi 2000 stereomicroscope equipped with Carl Zeiss lenses.
X-ray diffraction data for 1Eu and 2La were collected at 150(2)K on a Bruker D8 QUEST equipped with Mo Kα sealed tube (λ = 0.71073 Å), a multilayer TRIUMPH X-ray mirror, a PHOTON 100 CMOS detector, and an Oxford Instruments Cryostrem 700+ Series low temperature device. Diffraction images were processed using the software package SAINT+ [41], and data were corrected for absorption by the multiscan semi-empirical method implemented in SADABS 2016/2 [42]. Structures were solved using the algorithm implemented in SHELXT-2014/5 [43], which allowed the immediate location of almost all of the heaviest atoms composing the molecular unit of the two compounds. The remaining missing and misplaced non-hydrogen atoms were located from difference Fourier maps calculated from successive full-matrix least-squares refinement cycles on F2 using the latest SHELXL from the 2018/3 release [44]. All structural refinements were performed using the graphical interface ShelXle [45].
For 1Eu, the hydrogen atoms bound to carbon were placed at their idealized positions using the HFIX 43 in SHELXL-2014, which were included in subsequent refinement cycles with isotropic displacement parameters (Uiso) fixed at 1.2×Ueq of the parent carbon atoms. Hydrogen atoms associated with the terminal –POH groups were placed at their idealized positions using the HFIX 83 instruction and were refined assuming isotropic thermal displacements parameters (Uiso) fixed at 1.5×Ueq of the parent oxygen atoms. To find which groups should be protonated it was necessary to simultaneously take into account the P–O bond lengths and the possibility of forming hydrogen bonds with neighboring moieties. Two of these –POH groups shared the hydrogen atom with the neighboring phosphonic group. O23 and O33 shared the hydrogen atom with O30 and O38, respectively. These hydrogen atoms were refined with a 50% occupancy rate each.
The crystallization methanol molecule was refined with 50% of occupancy rate. The hydrogen atoms from the methyl and hydroxyl groups were placed at their idealized positions using the HFIX 137 and HFIX 83, respectively, and were included in subsequent refinement cycles with isotropic displacement parameters (Uiso) fixed at 1.5×Ueq of the parent non-hydrogen atoms.
Due to the considerable disorder of the crystallization water molecules, their hydrogen atoms could not be located from difference Fourier maps and no attempts were made to place them in calculated positions. These atoms were, however, included in the empirical formula of the compound (Table 1). The disorder of these water molecules was refined as follow: 30% occupancy rate for O16W and O19W; 40% for O15W, O21W, O22W and O23W; 50% for O10W, O20W, O25W and O26W; 60% for O14W and O17W; 70% for O13W; 80% for O24W; and 85% for O12W. These occupancy factors were fixed in the final structural model but were obtained from a previous unrestrained refinement for each site. All oxygen atoms belonging to the water molecules of crystallization were refined using a common isotropic displacement parameter.
For 2La the hydrogen atoms bound to carbon were placed at their idealized positions using the HFIX 43 in SHELXL-2014, which were included in subsequent refinement cycles with isotropic displacement parameters (Uiso) fixed at 1.2×Ueq of the parent carbon atoms. Hydrogen atoms associated with the terminal –POH groups were placed at their idealized positions using the HFIX 147 instruction and were refined assuming isotropic thermal displacements parameters (Uiso) fixed at 1.5×Ueq of the parent atoms. Two crystallization water molecules, namely O11W and O12W, were disordered and refined with an occupancy rate of 20% and 80% respectively. The hydrogen atoms of all water molecules could not be located from difference Fourier maps and no attempts were made to place them in calculated positions. These atoms were, however, included in the empirical formula of the compound.
The last difference Fourier map synthesis showed: for 1Eu the highest peak (2.005 eÅ−3) located at 0.65 Å from Eu1 and the deepest hole (−2.108 eÅ−3) located at 0.13 Å from O24W, while compound 2La showed the highest peak (2.468 eÅ−3) and the deepest hole (−1.888 eÅ−3) located at 0.56 Å and 0.90 Å from La4, respectively. Structural drawings have been created using the software package Crystal Impact Diamond [46].
Information concerning crystallographic data collection and structure refinement details is summarized in Table 1. Tables S1–S4 (in the Supporting Information) gather the most significant distances and angles of the lanthanide coordination environments and hydrogen bonding geometric details. Crystallographic data (including structure factors) for the crystal structures of 1Eu and 2La have been deposited with the Cambridge Crystallographic Data Centre as supplementary publication data No. 1,964,385 and 1964384, respectively. Copies of the data can be obtained free of charge on application to CCDC, 12 Union Road, Cambridge CB2 2EZ, U.K. FAX: (+44) 1223 336033. E-mail: [email protected].

4. Conclusions

The rigid tetrapodal organic linker [1,1′-biphenyl]-3,3′,5,5′-tetrayltetrakis(phosphonic acid) (H8btp) was used for the preparation of two new isotypical families of lanthanide–organic frameworks: Ln7(H5btp)4(H5.5btp)2(H6btp)2(H2O)12]∙23.5H2O∙MeOH [where Ln3+ = Eu3+ (1Eu) and Gd3+ (1Gd)], and [Ln4(H3btp)(H4btp)(H5btp)(H2O)8]∙3H2O [where Ln3+ = Ce3+ (2Ce), Pr3+ (2Pr), and Nd3+ (2Nd)]. This work complements our past research efforts using this ligand, particularly that which described in detail the photoluminescent family [Ln4(H6btp)2(H4btp)2(H8btp)(H2O)16]∙12H2O [where Ln3+ = La3+, (La0.9Eu0.1)3+ and (La0.9Tb0.1)3+] [28].
In our previous report we employed microwave-assisted synthesis at 100 °C (60 min at 50 W) followed by slow evaporation over a period of 7 days. Family 1 herein reported followed an identical synthetic procedure, while trying to explore the structural effect when using smaller radii for the lanthanide cations: we note that in our past work photoluminescent materials were only obtained by doping an inert La3+-based matrix. Two-dimensional compact layers were again obtained, in which the ligand-to-metal ratio is greater than unity. This is an important structural aspect because it means that, despite both synthesis contain in the reactive mixtures a great excess of metal centers, the microwave heating procedure followed by slow evaporation seems to favor the inclusion of a large number of solvent molecules, increasing concomitantly the number of supramolecular contacts in the structures. To note that the same experimental conditions using the hydrothermal method do not permit the preparation of this material. As described for [Ln4(H6btp)2(H4btp)2(H8btp)(H2O)16]∙12H2O, compound 1 is also formed by compact layers in which organic cylinders are interconnected by the metallic centers. The presence of optically active Eu3+ lanthanide centers yielded photoluminescent materials. As for our previous reported material, the presence of several crystallization and coordinated water molecules in the structure accounts for the lower emission of the Eu3+-based material.
Family 2 of compounds is formed by a three-dimensional network in which, comparatively, the ligand-to-metal ratio is below the unity. We achieved this by increasing the temperature to 140 °C and change the synthetic method to the traditional hydro(solvo)thermal one. This strategy, allied to the usage of larger lanthanides, seemed to favor metallic connectivity in detriment to the various supramolecular contacts involving solvent molecules. This new 3D family of compounds is significantly more compact, but it is still formed by organic cylinders interconnected by the metallic centers, as for family 1 and our previously reported [Ln4(H6btp)2(H4btp)2(H8btp)(H2O)16]∙12H2O compounds.
These two families of compound clearly show the importance of a careful selection of the synthetic experimental method. Despite both syntheses start with the same reactive mixtures (with a great excess of metal), the microwave heating procedure followed by slow evaporation seems to favor the inclusion of a large number of solvent molecules. This leads to an increasing of the number of supramolecular contacts in the structures, originating a 2D layered material. On the other hand, hydrothermal synthesis at higher temperature seems to favor metal coordination of a higher number of phosphonic groups, leading to a more compact 3D network.
Future work related to this system is now being currently focused at the selective control of the dimensionality of the networks so to explore various properties, particularly those related with conductivity: the presence of solvent molecules and many protonated phosphonate groups is expected to significantly boost proton conductivity.

Supplementary Materials

Additional crystallographic data: crystallographic tables, powder x-ray diffraction studies, topological studies. Additional characterization data: thermogravimetry, FT-IR, electron microscopy. Additional details on the synthesis and characterization techniques.

Author Contributions

Laboratory research: A.D.G.F. Ideas and concepts: F.A.A.P. and J.P.C.T. Crystal data acquisition and discussion: R.F.M., J.S.B. and F.A.A.P. Photoluminescence studies: D.A. Writing the original draft preparation: A.D.G.F., R.F.M., J.S.B. and D.A. Writing review and editing: J.P.C.T. and F.A.A.P. Supervision of students: J.P.C.T. and F.A.A.P. All authors have read and agreed to the published version of the manuscript.

Funding

This work was developed within the scope of the project CICECO-Aveiro Institute of Materials, UIDB/50011/2020 & UIDP/50011/2020, financed by national funds through the Portuguese Foundation for Science and Technology/MCTES. Thanks are also due to the University of Aveiro and FCT/MCT for the financial support for the QOPNA research Unit (FCT UID/QUI/00062/2019) and CQE (FCT UID/QUI/0100/2019) research units through national funds and, where applicable, co-financed by the FEDER, within the PT2020 Partnership Agreement.

Acknowledgments

FCT is gratefully acknowledged for the Ph.D. grant No. SFRH/BD/84495/2012 (to ADF) and PD/BD/135104/2017 (to JB). RFM also gratefully acknowledge FCT for the Junior Research Position (CEECIND/00553/2017).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Silva, P.; Vilela, S.M.F.; Tomé, J.P.C.; Paz, F.A.A. Multifunctional metal-organic frameworks: From academia to industrial applications. Chem. Soc. Rev. 2015, 44, 6774–6803. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Li, B.; Wen, H.M.; Cui, Y.J.; Zhou, W.; Qian, G.D.; Chen, B.L. Emerging multifunctional metal-organic framework materials. Adv. Mater. 2016, 28, 8819–8860. [Google Scholar] [CrossRef] [PubMed]
  3. Sanz-Perez, E.S.; Murdock, C.R.; Didas, S.A.; Jones, C.W. Direct capture of co2 from ambient air. Chem. Rev. 2016, 116, 11840–11876. [Google Scholar] [CrossRef] [PubMed]
  4. Suh, M.P.; Park, H.J.; Prasad, T.K.; Lim, D.W. Hydrogen storage in metal-organic frameworks. Chem. Rev. 2012, 112, 782–835. [Google Scholar] [CrossRef] [PubMed]
  5. Joharian, M.; Morsali, A.; Tehrani, A.A.; Carlucci, L.; Proserpio, D.M. Water-stable fluorinated metal-organic frameworks (f-mofs) with hydrophobic properties as efficient and highly active heterogeneous catalysts in aqueous solution. Green Chem. 2018, 20, 5336–5345. [Google Scholar] [CrossRef]
  6. Huang, Y.B.; Liang, J.; Wang, X.S.; Cao, R. Multifunctional metal-organic framework catalysts: Synergistic catalysis and tandem reactions. Chem. Soc. Rev. 2017, 46, 126–157. [Google Scholar] [CrossRef]
  7. Tehrani, A.A.; Esrafili, L.; Abedi, S.; Morsali, A.; Carlucci, L.; Proserpio, D.M.; Wang, J.; Junk, P.C.; Liu, T.F. Urea metal-organic frameworks for nitro-substituted compounds sensing. Inorg. Chem. 2017, 56, 1446–1454. [Google Scholar] [CrossRef]
  8. Smith, J.A.; Singh-Wilmot, M.A.; Carter, K.P.; Cahill, C.L.; Ridenour, J.A. Lanthanide-2,3,5,6-tetrabromoterephthalic acid metal-organic frameworks: Evolution of halogen center dot center dot center dot halogen interactions across the lanthanide series and their potential as selective bifunctional sensors for the detection of fe3+, cu2+, and nitroaromatics. Cryst. Growth Des. 2019, 19, 305–319. [Google Scholar]
  9. Lustig, W.P.; Mukherjee, S.; Rudd, N.D.; Desai, A.V.; Li, J.; Ghosh, S.K. Metal-organic frameworks: Functional luminescent and photonic materials for sensing applications. Chem. Soc. Rev. 2017, 46, 3242–3285. [Google Scholar] [CrossRef]
  10. Pereira, C.F.; Figueira, F.; Mendes, R.F.; Rocha, J.; Hupp, J.T.; Farha, O.K.; Simões, M.M.Q.; Tomé, J.P.C.; Paz, F.A.A. Bifunctional porphyrin-based nano-metal-organic frameworks: Catalytic and chemosensing studies. Inorg. Chem. 2018, 57, 3855–3864. [Google Scholar] [CrossRef]
  11. Venkatramaiah, N.; Pereira, C.F.; Mendes, R.F.; Paz, F.A.A.; Tomé, J.P.C. Phosphonate appended porphyrins as versatile chemosensors for selective detection of trinitrotoluene. Anal. Chem. 2015, 87, 4515–4522. [Google Scholar] [CrossRef] [PubMed]
  12. Chen, W.; Wu, C.S. Synthesis, functionalization, and applications of metal-organic frameworks in biomedicine. Dalton Trans. 2018, 47, 2114–2133. [Google Scholar] [CrossRef] [PubMed]
  13. Horcajada, P.; Gref, R.; Baati, T.; Allan, P.K.; Maurin, G.; Couvreur, P.; Férey, G.; Morris, R.E.; Serre, C. Metal-organic frameworks in biomedicine. Chem. Rev. 2012, 112, 1232–1268. [Google Scholar] [CrossRef] [PubMed]
  14. Dong, J.; Zhao, D.; Lu, Y.; Sun, W.Y. Photoluminescent metal-organic frameworks and their application for sensing biomolecules. J. Mater. Chem. A 2019, 7, 22744–22767. [Google Scholar] [CrossRef]
  15. He, J.; Xu, J.L.; Yin, J.C.; Li, N.; Bu, X.H. Recent advances in luminescent metal-organic frameworks for chemical sensors. Sci. China-Mater. 2019, 62, 1655–1678. [Google Scholar] [CrossRef] [Green Version]
  16. Kuyuldar, S.; Genna, D.T.; Burda, C. On the potential for nanoscale metal-organic frameworks for energy applications. J. Mater. Chem. A 2019, 7, 21545–21576. [Google Scholar] [CrossRef]
  17. Zhang, Z.; Sang, W.; Xie, L.S.; Dai, Y.L. Metal-organic frameworks for multimodal bioimaging and synergistic cancer chemotherapy. Coord. Chem. Rev. 2019, 399, 213022. [Google Scholar] [CrossRef]
  18. Bobbitt, N.S.; Mendonça, M.L.; Howarth, A.J.; Islamoglu, T.; Hupp, J.T.; Farha, O.K.; Snurr, R.Q. Metal-organic frameworks for the removal of toxic industrial chemicals and chemical warfare agents. Chem. Soc. Rev. 2017, 46, 3357–3385. [Google Scholar] [CrossRef]
  19. Wang, H.L.; Zhu, Q.L.; Zou, R.Q.; Xu, Q. Metal-organic frameworks for energy applications. Chem 2017, 2, 52–80. [Google Scholar] [CrossRef] [Green Version]
  20. Colodrero, R.M.P.; Salcedo, I.R.; Bazaga-Garcia, M.; Barouda, E.; Papadaki, M.; Papathanasiou, K.E.; Hernandez-Alonso, D.; Rius, J.; Aranda, M.A.G.; Losilla, E.R.; et al. High-throughput synthesis of pillared-layered magnesium tetraphosphonate coordination polymers: Framework interconversions and proton conductivity studies. Inorganics 2018, 6, 96. [Google Scholar] [CrossRef] [Green Version]
  21. Salcedo, I.R.; Colodrero, R.M.P.; Bazaga-Garcia, M.; Vasileiou, A.; Papadaki, M.; Olivera-Pastor, P.; Infantes-Molina, A.; Losilla, E.R.; Mezei, G.; Cabeza, A.; et al. From light to heavy alkali metal tetraphosphonates (m = li, na, k, rb, cs): Cation size-induced structural diversity and water-facilitated proton conductivity. Crystengcomm 2018, 20, 7648–7658. [Google Scholar] [CrossRef]
  22. Ji, B.M.; Deng, D.S.; Ma, J.Y.; Sun, C.W.; Zhao, B. Two- and three-dimensional lanthanide-based coordination polymers assembled by the synergistic effect of various lanthanide radii and flexibility of a new binicotinate-containing ligand: In situ synthesis, structures, and properties. RSC Adv. 2015, 5, 2239–2248. [Google Scholar] [CrossRef]
  23. Xiang, S.C.; Hu, S.M.; Sheng, T.L.; Chen, J.S.; Wu, X.T. Structural diversity of infinite 3d-4f heterometallic cluster compounds driven by various lanthanide radii. Chem.-Eur. J. 2009, 15, 12496–12502. [Google Scholar] [CrossRef] [PubMed]
  24. Yan, X.H.; Li, Y.F.; Wang, Q.; Huang, X.G.; Zhang, Y.; Gao, C.J.; Liu, W.S.; Tang, Y.; Zhang, H.R.; Shao, Y.L. Two- to one-dimensional: Radii-dependent self-assembly crystal structures and luminescent properties of lanthanide coordination polymers with an amide type semirigid bridging ligand. Cryst. Growth Des. 2011, 11, 4205–4212. [Google Scholar] [CrossRef]
  25. Wang, H.S.; Li, J.; Li, J.Y.; Wang, K.; Ding, Y.; Xia, X.H. Lanthanide-based metal-organic framework nanosheets with unique fluorescence quenching properties for two-color intracellular adenosine imaging in living cells. NPG Asia Mater. 2017, 9, e354. [Google Scholar] [CrossRef] [Green Version]
  26. Zareba, J.K. Tetraphenylmethane and tetraphenylsilane as building units of coordination polymers and supramolecular networks—A focus on tetraphosphonates. Inorg. Chem. Commun. 2017, 86, 172–186. [Google Scholar] [CrossRef]
  27. Zheng, T.; Yang, Z.X.; Gui, D.X.; Liu, Z.Y.; Wang, X.X.; Dai, X.; Liu, S.T.; Zhang, L.J.; Gao, Y.; Chen, L.H.; et al. Overcoming the crystallization and designability issues in the ultrastable zirconium phosphonate framework system. Nat. Commun. 2017, 8, 15369. [Google Scholar] [CrossRef]
  28. Firmino, A.D.G.; Mendes, R.F.; Ananias, D.; Vilela, S.M.F.; Carlos, L.D.; Tomé, J.P.C.; Rocha, J.; Paz, F.A.A. Microwave synthesis of a photoluminescent metal-organic framework based on a rigid tetraphosphonate linker. Inorg. Chim. Acta 2017, 455, 584–594. [Google Scholar] [CrossRef]
  29. Mendes, R.F.; Ananias, D.; Carlos, L.D.; Rocha, J.; Paz, F.A.A. Photoluminescent lanthanide–organic framework based on a tetraphosphonic acid linker. Cryst. Growth Des. 2017, 17, 5191–5199. [Google Scholar] [CrossRef]
  30. Vilela, S.M.F.; Firmino, A.D.G.; Mendes, R.F.; Fernandes, J.A.; Ananias, D.; Valente, A.A.; Ott, H.; Carlos, L.D.; Rocha, J.; Tomé, J.P.C.; et al. Lanthanide-polyphosphonate coordination polymers combining catalytic and photoluminescence properties. Chem. Commun. 2013, 49, 6400–6402. [Google Scholar] [CrossRef]
  31. Ananias, D.; Firmino, A.D.G.; Mendes, R.F.; Paz, F.A.A.; Nolasco, M.; Carlos, L.D.; Rocha, J. Excimer formation in a terbium metal-organic framework assists luminescence thermometry. Chem. Mat. 2017, 29, 9547–9554. [Google Scholar] [CrossRef]
  32. Mendes, R.F.; Antunes, M.M.; Silva, P.; Barbosa, P.; Figueiredo, F.; Linden, A.; Rocha, J.; Valente, A.A.; Paz, F.A.A. A lamellar coordination polymer with remarkable catalytic activity. Chem.-A Eur. J. 2016, 22, 13136–13146. [Google Scholar] [CrossRef]
  33. Mendes, R.F.; Silva, P.; Antunes, M.M.; Valente, A.A.; Paz, F.A.A. Sustainable synthesis of a catalytic active one-dimensional lanthanide–organic coordination polymer. Chem. Commun. 2015, 51, 10807–10810. [Google Scholar] [CrossRef] [PubMed]
  34. Firmino, A.D.G.; Mendes, R.F.; Antunes, M.M.; Barbosa, P.C.; Vilela, S.M.F.; Valente, A.A.; Figueiredo, F.M.L.; Tomé, J.P.C.; Paz, F.A.A. Robust multifunctional yttrium-based metal organic frameworks with breathing effect. Inorg. Chem. 2017, 56, 1193–1208. [Google Scholar] [CrossRef] [PubMed]
  35. Alexandrov, E.V.; Blatov, V.A.; Kochetkov, A.V.; Proserpio, D.M. Underlying nets in three-periodic coordination polymers: Topology, taxonomy and prediction from a computer-aided analysis of the cambridge structural database. Crystengcomm 2011, 13, 3947–3958. [Google Scholar] [CrossRef]
  36. Blatov, V.A.; Shevchenko, A.P.; Proserpio, D.M. Applied topological analysis of crystal structures with the program package topospro. Cryst. Growth Des. 2014, 14, 3576–3586. [Google Scholar] [CrossRef]
  37. In RCSR (Reticular Chemistry Structure Resource). Available online: http://rcsr.anu.edu.au/ (accessed on 15 January 2020).
  38. In EPINET. Available online: http://epinet.anu.edu.au (accessed on 15 January 2020).
  39. Bonneau, C.; O’Keeffe, M.; Proserpio, D.M.; Blatov, V.A.; Batten, S.R.; Bourne, S.A.; Lah, M.S.; Eon, J.G.; Hyde, S.T.; Wiggin, S.B.; et al. Deconstruction of crystalline networks into underlying nets: Relevance for terminology guidelines and crystallographic databases. Cryst. Growth Des. 2018, 18, 3411–3418. [Google Scholar] [CrossRef]
  40. Kottke, T.; Stalke, D.J. Crystal handling at low temperatures. Appl. Crystallogr. 1993, 26, 615–619. [Google Scholar] [CrossRef] [Green Version]
  41. SAINT+. Data Integration Engine v. 8.37a©; Bruker AXS: Madison, WI, USA, 1997–2015. [Google Scholar]
  42. Krause, L.; Herbst-Irmer, R.; Sheldrick, G.M.; Stalke, D. Comparison of silver and molybdenum microfocus x-ray sources for single-crystal structure determination. J. Appl. Crystallogr. 2015, 48, 3–10. [Google Scholar] [CrossRef] [Green Version]
  43. Sheldrick, G.M. Shelxt - integrated space-group and crystal-structure determination. Acta Cryst. A 2015, 71, 3–8. [Google Scholar] [CrossRef] [Green Version]
  44. Sheldrick, G.M. Crystal structure refinement with shelxl. Acta Crystallogr. Sect. C-Struct. Chem. 2015, 71, 3–8. [Google Scholar] [CrossRef] [PubMed]
  45. Hübschle, C.B.; Sheldrick, G.M.; Dittric, B.J. ShelXle: A Qt graphical user interface for SHELXL. Appl. Crystallogr. 2011, 44, 1281–1284. [Google Scholar]
  46. Brandenburg, K. DIAMOND, Version 3.2f; Crystal Impact GbR: Bonn, Germany, 1997–2010. [Google Scholar]
Sample Availability: Limited amount of samples of the compounds may be available from the authors upon request.
Scheme 1. Schematic representation of the organic linker [1,1′-biphenyl]-3,3′,5,5′-tetrayltetrakis(phosphonic acid) (H8btp) used in this work.
Scheme 1. Schematic representation of the organic linker [1,1′-biphenyl]-3,3′,5,5′-tetrayltetrakis(phosphonic acid) (H8btp) used in this work.
Molecules 25 01835 sch001
Figure 1. Polyhedral representation of the four distorted Eu3+ coordination environments present in the crystal structure of [Eu7(H5btp)4(H5.5btp)2(H6btp)2(H2O)12]∙23.5H2O∙MeOH (1Eu). Table S1 (in the Supporting Information) gather all the geometrical details on the represented bond lengths, angles. Symmetry transformations used to generate equivalent atoms: (i) −x, 1−y, 1−z; (ii) −x, 1−y, -z; (iii) x, 3/2−y, −1/2 + z.
Figure 1. Polyhedral representation of the four distorted Eu3+ coordination environments present in the crystal structure of [Eu7(H5btp)4(H5.5btp)2(H6btp)2(H2O)12]∙23.5H2O∙MeOH (1Eu). Table S1 (in the Supporting Information) gather all the geometrical details on the represented bond lengths, angles. Symmetry transformations used to generate equivalent atoms: (i) −x, 1−y, 1−z; (ii) −x, 1−y, -z; (iii) x, 3/2−y, −1/2 + z.
Molecules 25 01835 g001
Figure 2. Crystal packing of [Eu7(H5btp)4(H5.5btp)2(H6btp)2(H2O)12]∙23.5H2O∙MeOH (1Eu) viewed in perspective along the (a) [100] and (b) [010] directions of the unit cell. The representations emphasize simultaneously the presence of several dimmers formed by the metal centers and the phosphonic acid groups that connects the “1D” organic cylinders of 1Eu.
Figure 2. Crystal packing of [Eu7(H5btp)4(H5.5btp)2(H6btp)2(H2O)12]∙23.5H2O∙MeOH (1Eu) viewed in perspective along the (a) [100] and (b) [010] directions of the unit cell. The representations emphasize simultaneously the presence of several dimmers formed by the metal centers and the phosphonic acid groups that connects the “1D” organic cylinders of 1Eu.
Molecules 25 01835 g002
Figure 3. Schematic representation of the organic cylinder present in the crystal structure of [Eu7(H5btp)4(H5.5btp)2(H6btp)2(H2O)12]∙23.5H2O∙MeOH (1Eu) viewed along the [010] and [100] directions of the unit cell. The schemes emphasize the close packing of the organic linkers, with inter-planar distances ranging between 3.5805(4) and 3.8321(5) Å. On the right there is a representation of the same organic skeleton in the cylinder emphasizing the close packing and rotation along the a-axis of the four composing aromatic cores.
Figure 3. Schematic representation of the organic cylinder present in the crystal structure of [Eu7(H5btp)4(H5.5btp)2(H6btp)2(H2O)12]∙23.5H2O∙MeOH (1Eu) viewed along the [010] and [100] directions of the unit cell. The schemes emphasize the close packing of the organic linkers, with inter-planar distances ranging between 3.5805(4) and 3.8321(5) Å. On the right there is a representation of the same organic skeleton in the cylinder emphasizing the close packing and rotation along the a-axis of the four composing aromatic cores.
Molecules 25 01835 g003
Figure 4. Polyhedral representation of the distorted coordination environments present in the crystal structure of [La4(H5btp)(H4btp)(H3btp)(H2O)8]∙3H2O (2La). See Table S1 for geometrical details on the represented bond lengths and angles. Symmetry transformation used to generate equivalent atoms: (i) −x + 1, y − 1/2, −z + 3/2; (ii) x + 1, y, z; (iii) −x + 1, −y, −z + 1; (iv) −x + 1, −y + 1, −z + 1; (v) −x, −y + 1, −z + 1; (vi) x + 1, −y + 1/2, z + 1/2; (vii) x − 1, y, z; (viii) x − 1, −y + 1/2, z + 1/2; (ix) x, −y + 1/2, z + 1/2.
Figure 4. Polyhedral representation of the distorted coordination environments present in the crystal structure of [La4(H5btp)(H4btp)(H3btp)(H2O)8]∙3H2O (2La). See Table S1 for geometrical details on the represented bond lengths and angles. Symmetry transformation used to generate equivalent atoms: (i) −x + 1, y − 1/2, −z + 3/2; (ii) x + 1, y, z; (iii) −x + 1, −y, −z + 1; (iv) −x + 1, −y + 1, −z + 1; (v) −x, −y + 1, −z + 1; (vi) x + 1, −y + 1/2, z + 1/2; (vii) x − 1, y, z; (viii) x − 1, −y + 1/2, z + 1/2; (ix) x, −y + 1/2, z + 1/2.
Molecules 25 01835 g004
Figure 5. Crystal packing of [La4(H5btp)(H4btp)(H3btp)(H2O)8]∙3H2O (2La) viewed in perspective along the (a) [001] and (b) [010] directions of the unit cell. The representations also emphasize, simultaneously, not only the presence of a metallic dimer formed by two metal centers and the phosphonic acid groups, responsible for the connection between the “1D” cylinders along the [100] direction, but also the presence of a metallic tetramer which connects the same cylinders along the [010] direction.
Figure 5. Crystal packing of [La4(H5btp)(H4btp)(H3btp)(H2O)8]∙3H2O (2La) viewed in perspective along the (a) [001] and (b) [010] directions of the unit cell. The representations also emphasize, simultaneously, not only the presence of a metallic dimer formed by two metal centers and the phosphonic acid groups, responsible for the connection between the “1D” cylinders along the [100] direction, but also the presence of a metallic tetramer which connects the same cylinders along the [010] direction.
Molecules 25 01835 g005
Figure 6. Schematic representation of the organic cylinder present in the crystal structure of [La4(H5btp)(H4btp)(H3btp)(H2O)8]∙3H2O (2La). The schemes emphasize the close packing of the organic linkers, with inter-planar distances ranging between 3.5545(5) and 3.9692(5) Å. On the right there is a representation of the same organic skeleton in the cylinder emphasizing the close packing and rotation along the a-axis of the three aromatic cores.
Figure 6. Schematic representation of the organic cylinder present in the crystal structure of [La4(H5btp)(H4btp)(H3btp)(H2O)8]∙3H2O (2La). The schemes emphasize the close packing of the organic linkers, with inter-planar distances ranging between 3.5545(5) and 3.9692(5) Å. On the right there is a representation of the same organic skeleton in the cylinder emphasizing the close packing and rotation along the a-axis of the three aromatic cores.
Molecules 25 01835 g006
Figure 7. Excitation spectra of [Eu7(H5btp)4(H5.5btp)2(H6btp)2(H2O)12]∙23.5H2O∙MeOH (1Eu), recorded at ambient temperature (black line) and at 12 K (red line), while monitoring the Eu3+ emission at 610.5 nm.
Figure 7. Excitation spectra of [Eu7(H5btp)4(H5.5btp)2(H6btp)2(H2O)12]∙23.5H2O∙MeOH (1Eu), recorded at ambient temperature (black line) and at 12 K (red line), while monitoring the Eu3+ emission at 610.5 nm.
Molecules 25 01835 g007
Figure 8. Emission spectra of [Eu7(H5btp)4(H5.5btp)2(H6btp)2(H2O)12]∙23.5H2O∙MeOH (1Eu), recorded at ambient temperature (black line) and at 12 K (red line), with the excitation fixed at 393 nm. The inset shows the Eu3+5D17F0,1,2 transitions region recorded at 12 K using excitation wavelengths of 393 (red line) and 310 nm (magenta line).
Figure 8. Emission spectra of [Eu7(H5btp)4(H5.5btp)2(H6btp)2(H2O)12]∙23.5H2O∙MeOH (1Eu), recorded at ambient temperature (black line) and at 12 K (red line), with the excitation fixed at 393 nm. The inset shows the Eu3+5D17F0,1,2 transitions region recorded at 12 K using excitation wavelengths of 393 (red line) and 310 nm (magenta line).
Molecules 25 01835 g008
Table 1. Crystal data collection and structure refinement details for [Eu7(H5btp)4(H5.5btp)2(H6btp)2(H2O)12]∙23.5H2O∙MeOH (1Eu) and [La4(H5btp)(H4btp)(H3btp)(H2O)8]∙3H2O (2La).
Table 1. Crystal data collection and structure refinement details for [Eu7(H5btp)4(H5.5btp)2(H6btp)2(H2O)12]∙23.5H2O∙MeOH (1Eu) and [La4(H5btp)(H4btp)(H3btp)(H2O)8]∙3H2O (2La).
1Eu2La
FormulaC97H166Eu7O132.5P32C36H52La4O47P12
Formula weight5507.052164.05
Temperature/K150(2)150(2)
Crystal systemMonoclinicMonoclinic
Space groupP21/cP21/c
a13.6342(19)12.5332(16)
b34.172(5)24.147(3)
c19.049(3)20.852(3)
b93.213(5)100.037(3)
Volume/Å38861(2)6213.9(14)
Z24
ρcalc/g cm−32.0642.313
m(Mo Ka)/mm−12.8543.122
Crystal typeColorless plateColorless plate
Crystal size/mm0.20 × 0.10 × 0.020.17 × 0.06 × 0.02
θ range (°)3.52–25.353.518–29.13
Index ranges−16 ≤ h ≤ 16
−41 ≤ k ≤ 41
−22 ≤ l ≤ 22
−17 ≤ h ≤ 17
−33 ≤ k ≤ 33
−26 ≤ l ≤ 28
Collected Reflections243,730167,012
Independent Reflections16,181 (Rint = 0.1633)16,687 (Rint = 0.0779)
Completeness99.7% (θ = 25.24)99.7% (θ = 25.24)
Final R indices [I > 2σ(I)]R1 = 0.0639
wR2 = 0.1327
R1 = 0.0549
wR2 = 0.1101
Final R indices (all data)R1 = 0.1088
wR2 = 0.1507
R1 = 0.0709
wR2 = 0.1153
Largest diff. peak and hole/eÅ−32.005 and −2.1082.468 and −1.888
R 1 = F 0 F c / F 0 ;
w R 2 = [ w ( F 0 2 F c 2 ) 2 ] / [ w ( F 0 2 ) 2 ]
w = 1 / [ σ 2 ( F 0 2 ) + ( m P ) 2 + n P ] where P = ( F 0 2 + 2 F c 2 ) / 3

Share and Cite

MDPI and ACS Style

Firmino, A.D.G.; Mendes, R.F.; Ananias, D.; Barbosa, J.S.; Tomé, J.P.C.; Almeida Paz, F.A. Coordination Polymers Based on a Biphenyl Tetraphosphonate Linker: Synthesis Control and Photoluminescence. Molecules 2020, 25, 1835. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules25081835

AMA Style

Firmino ADG, Mendes RF, Ananias D, Barbosa JS, Tomé JPC, Almeida Paz FA. Coordination Polymers Based on a Biphenyl Tetraphosphonate Linker: Synthesis Control and Photoluminescence. Molecules. 2020; 25(8):1835. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules25081835

Chicago/Turabian Style

Firmino, Ana D. G., Ricardo F. Mendes, Duarte Ananias, Jéssica S. Barbosa, João P. C. Tomé, and Filipe A. Almeida Paz. 2020. "Coordination Polymers Based on a Biphenyl Tetraphosphonate Linker: Synthesis Control and Photoluminescence" Molecules 25, no. 8: 1835. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules25081835

Article Metrics

Back to TopTop