Controlled Synthesis of Supramolecular Architectures of Homo- and

Nov 27, 2018 - Controlled Synthesis of Supramolecular Architectures of Homo- and Heterometallic Complexes by Programmable Self-Assembly...
0 downloads 0 Views 5MB Size
Communication Cite This: Cryst. Growth Des. XXXX, XXX, XXX−XXX

pubs.acs.org/crystal

Controlled Synthesis of Supramolecular Architectures of Homo- and Heterometallic Complexes by Programmable Self-Assembly Published as part of a Crystal Growth and Design virtual special issue on Crystalline Functional Materials in Honor of Professor Xin-Tao Wu Jin Tong, Li-Mei Jia, Ping Shang, and Shu-Yan Yu* Beijing Key Laboratory for Green Catalysis and Separation, Department of Chemistry and Chemical Industry, Beijing University of Technology, Beijing, 100124, P. R. China Crystal Growth & Design Downloaded from pubs.acs.org by UNIV OF WINNIPEG on 11/29/18. For personal use only.

S Supporting Information *

ABSTRACT: The reaction of ligand (H2L) named 3-(3,5-dimethyl1H-pyrazol-4-yl)pentane-2,4-dione and Cu(NO3)2 produces two types of supramolecular homometallic complexes well-controlled by stepwise deprotonation. With lower NaHCO3 concentrations as the base, the product [Cu(HL)2]4 (1) exhibits a two-dimensional (2D) rhombus with Cu/(HL) = 1:2. When NaHCO3 is further employed on complex 1, however, the product [Cu(HL)(L)]n (2) with Cu/(HL)/(L) = 1:1:1 displays a zigzag one-dimensional (1D) chain. The ditopic ligand is used also for the programmable self-assembly of a zero-dimensional (0D) mononuclear complex [Fe(HL)3] (3) and one-dimensional (1D) heterometallic multinuclear complex [AgFe(HL)3(NO3)]n (4). Complex 3 readily forms upon reaction of H2L with Fe(NO3)3 and NaHCO3 as the base. The reaction of 3 with the second metal salt yielded the complex 4 with a porous ladder structure. Complexes 1, 2, 3, and 4 were investigated by a combination of X-ray crystallography, powder X-ray diffraction, elemental analysis, thermogravimetric analysis, and infrared spectroscopy. Magnetic susceptibilities of these complexes 1 and 2 have been also measured by SQUID techniques.



INTRODUCTION The design and preparation of homo- and heterometallic supramolecular complexes have continuously received much attention in recent years1−13 because of wide-ranging applications such as gas adsorption,14,15 drug delivery,16−18 catalysis,19−24 magnetism,25,26 sensing,27,28 reactivity modulation,29−36 and host−guest inclusion.37,38 Although several supramolecular architectures have been reported, the homoand heterometallic systems incorporating multimetal centers constructed by the well-designed and hierarchical self-assembly are rare.39−42 The reasonable design and successful preparation of these systems by designing suitably multifunctional building blocks following stepwise synthesis and controllable selfassembly approach still display a great challenge. According to previous studies published by our group42−46 and others,47−55 there are a range of supramolecular structures with homometals and heterometals based on the compartmental ligands featuring a pyrazole bridging unit and chelating substituents in the 3-, 4-, or 5-positions of the functional group serve as valuable mono- or multinucleating scaffolds in coordination chemistry, where the metal could be unvalent, bivalent, and/or tervalent d-block cations typically representing coordination geometries from four to six in their each structures. Some specific pyrazolate bridged ditopic ligand systems such as HnLpz (n = 1 and 2; pz = pyrazole, Scheme 1) © XXXX American Chemical Society

are moieties containing O/N atoms that could be deprotonated and act as bridging pockets. Furthermore, they have found use in the programmable self-assembly of homo- and heterometallic complexes with various transition metal ions.56−68 On the basis of previous studies on metal−organic complexes via a programmable self-assembly approach represented by our group,60,63,64 here we report the syntheses and characterizations of four homo- and heterometallic supramolecular complexes [Cu(HL)2]4 (1), [Cu(HL)(L)]n (2), [Fe(HL)3] (3), and [AgFe(HL)3(NO3)]n (4), derived from the reaction of ligand(H2L) 3-(3,5-dimethyl-1H-pyrazol4-yl)pentane-2,4-dione and CuII, FeIII, and AgI ions in the mixed solvent by controlling the degree of deprotonation and programmable self-assembly (Scheme 2). Single X-ray diffraction, powder X-ray diffraction, and elemental analysis techniques were used to characterize the structures of these complexes. And their magnetic properties are evaluated by SQUID techniques. Received: September 18, 2018 Revised: November 20, 2018 Published: November 27, 2018 A

DOI: 10.1021/acs.cgd.8b01406 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Communication

Scheme 1. Pyrazole-Based Ditopic Ligands Used for Homo- and Heterometallic Complexes in Our Previous Work60,63,64

Scheme 2. Self-Assembly of the Homo- and Heterometallic Complexes 1−4 Based on Ligand (H2L)



by heating under an N2 gas flow in a range from 20 to 800 °C in 10 °C/min using a PerkinElmer Pyris1. Temperaturedependent magnetic susceptibilities were recorded by a Quantum-Design MPMS-XL-5 at 0.5 T from 1.8 to 300 K. Synthesis of Complex [Cu(HL)2]4 (1). H2L (0.2 mmol, 19.0 mg) in methanol (2.0 mL) and Cu(NO3)2·6H2O (0.1 mmol, 30.0 mg) in water (2.0 mL) were slowly mixed together, and the system became green. NaHCO3 (0.2 mmol, 17.3 mg in 0.80 mL H2O) aqueous solution was added to the above mixture after stirring overnight, and then the system was continued for 12 h, resulting in a clear reaction solution. The solution was left to evaporate slowly to give a light green

EXPERIMENTAL SECTION Materials and Instrumentation. All chemicals were of analytical grade and used directly without purified procedure. Ligand (H2L) 3-(3,5-dimethyl-1H-pyrazol-4-yl)pentane-2,4dione was prepared according to a literature.69 The IR spectra (KBr powder) were recorded ranging between 4000 and 400 cm−1 on an IR Affinity-1 instrument. Elemental analyses for C, H, and N were carried out on an EA 1108 (Carlo Erba Instruments) elemental analyzer. Powder Xray diffraction (PXRD) were carried out by BRUKER D8Focus Bragg−Brentano (Cu, λ = 1.54178 Å) at room temperature. Thermogravimetric analyses (TGA) were done B

DOI: 10.1021/acs.cgd.8b01406 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Communication

crystalline material. IR (solid state, cm−1): 3295 (m), 2922 (w), 1985 (w), 1566 (s), 1380 (s), 1009 (m), 923 (w), 676 (w), 621 (w). Anal. Calcd for C80H104Cu4N16O16: C, 53.44; H, 5.83; N, 12.47. Found: C, 53.53; H, 5.64; N, 12.38. Synthesis of Complex [Cu(HL)(L)]n (2). At room temperature, NaHCO3 (0.05 mmol, 4.20 mg in 0.80 mL H2O) aqueous solution was added to a solution of 1 (0.1 mmol, 180.0 mg) with water/methanol (5.0 mL, 1:1), and the system was continued for 1 day. Then the clear green solution was obtained by filtering and evaporated to prepare a deep green crystalline material. IR (solid state, cm−1): 3294 (m), 2922 (w), 1985 (w), 1574 (s), 1380 (s), 1001 (m), 916 (w), 668 (w), 606 (w). Anal. Calcd for C20H27CuN4O5: C, 51.49; H, 5.84; N, 12.02. Found: C, 51.22; H, 5.69; N, 12.21. Synthesis of Complex [Fe(HL)3] (3). Fe(NO3)3·9H2O (1.0 mmol, 404.0 mg) in 2.0 mL of H2O and H2L (3.0 mmol, 660.0 mg) in 2.0 mL of methanol were mixed, and then a solution of NaHCO3 (3.0 mmol, 260.0 mg in 0.8 mL H2O) aqueous solution was added. The red crystalline material was obtained by evaporation of the mixture after 24 h. IR (solid state, cm−1): 3208 (w), 2922 (w), 1985 (w), 1566 (s), 1357 (s), 1001 (m), 978 (w), 916 (m) 668 (m), 590 (w). Anal. Calcd for C30H39FeN6O6: C, 56.70; N, 13.22; H, 6.19. Found: C, 56.53; N, 13.38; H, 6.24. Synthesis of Complex [AgFe(HL)3(NO3)]n (4). It was obtained using a layering technique. In a typical synthesis, AgNO3 (0.06 mmol, 101.0 mg) solution was laid in 2.0 mL of CH3CN over 3 (0.02 mmol, 17.0 mg) solution in 2.0 mL of CH3OH. Slow interdiffusion of the system for 30 days led to crystallization of the red hexagonal-like crystalline material. IR (solid state, cm−1): 3208 (w), 2915 (w), 1559 (s), 1342 (s), 1009 (m), 923 (w), 916 (w), 676 (w), 583 (w). Anal. Calcd for C31H43AgFeN7O10: C, 44.46; H, 5.18; N, 11.71. Found: C, 44.52; H, 5.13; N, 11.58. X-ray Structure Determination and Structure Refinement. All single crystal X-ray diffraction data were recorded by a Rigaku SuperNova CCD X-ray diffractometer (Cu Kα, λ = 1.54184 Å) at 290 K (Tables 1 and 2, selected bond distances, angles, intermolecular interactions Tables S1−S5). Raw data collection and reduction were carried out with APEX2 software. All structures were solved by direct methods using SHELX-2015 software package and the Olex2 program.70 Non-hydrogen atoms of all complexes were refined with anisotropic thermal parameters. The H atoms were included in calculated positions. Crystallographic data of structures reported in this work have been deposited in the Cambridge Crystallographic Data Center with CCDC numbers 1831162− 1831165.

Table 1. Crystallographic Data for 1 and 2 empirical formula formula weight measurement temperature crystal system space group a [Å] b [Å] c [Å] α [deg] β [deg] γ [deg] V/Å3 Z ρcalcd (g/cm3) independent reflections (I > 2σ(I)) F(000) 2θ range for data collection limiting indices

goodness-of-fit on F2 R1,awR2b [I > 2σ(I)] R1,a wR2b (all data) largest diff. peak and hole (e/Å3) CCDC number a

1

2

C40H52Cu2N8O8 899.97 150.00(10) K monoclinic P21/c 8.95230(10) 32.1143(5) 16.3304(2) 90 102.6260(10) 90 4581.40(11) 4 1.305 9059 [0.0382]

C20H27CuN4O5 466.99 173.00(10) K orthorhombic Fdd2 27.3992(5) 24.4533(5) 14.1161(2) 90 90 90 9457.8(3) 16 1.312 4179 [0.0237]

1880.0 7.816−148.014 −10 ≤ h ≤ 9 −37 ≤ k ≤ 39 −16 ≤ l ≤ 20 1.033 0.0396, 0.1024 0.0463, 0.1076 0.33 and −0.38 1831164

3904.0 7.92−147.742 −32 ≤ h ≤ 23 −30 ≤ k ≤ 30 −17 ≤ l ≤ 17 1.048 0.0351, 0.0950 0.0359, 0.0962 0.53 and −0.32 1831165

R1= Σ(||F0| − |FC||)/Σ|F0|. bwR2 = [Σw(|F0|2 − |FC|2)2/Σw(F02)]1/2.

Table 2. Crystallographic Data for 3 and 4 empirical formula formula weight measurement temperature crystal system space group a [Å] b [Å] c [Å] α [deg] β [deg] γ [deg] V/Å3 Z ρcalcd (g/cm3) independent reflections (I > 2σ(I)) F(000) 2θ range for data collection limiting indices



RESULTS AND DISCUSSION Synthesis and General Characterization of Complexes 1−4. Complexes 1−4 were prepared by stepwise synthesis and programmable self-assembly. The ligand (H2L) displays the bis-bidentate double-anionic property and exhibits two kinds of coordination sites: N atoms of the pyrazolyl moiety and O atoms of the acetylacetonyl moiety. Pyrazolyl and acetylacetonyl moieties would be deprotonated and act as bridging linkers, which may bind multiple same or different kinds of metal centers controlled by the degree of deprotonation and programmable self-assembly. In 1−4, every ligand (H2L) may act as a monovalent (HL)− or divalent (L)2− anion by removing one or two protons. With

goodness-of-fit on F2 R1,a wR2b [I > 2σ(I)] R1,a wR2b (all data) largest diff. peak and hole (e/Å3) CCDC number a

C

3

4

C30H39FeN6O6 635.52 150.00(10) K monoclinic C2/c 16.6573(6) 16.2900(5) 31.5252(9) 90 103.419(3) 90 8320.7(5) 8 1.015 8243 [0.0456]

C31H43AgFeN7O10 837.44 490.00(2) K monoclinic P21/c 15.3183(8) 22.671(2) 14.0097(10) 90 106.924(7) 90 4654.5(6) 4 1.195 7876 [0.0645]

2680.0 7.696−147.972 −18 ≤ h ≤ 20 −19 ≤ k ≤ 19 −39 ≤ l ≤ 26 1.046 0.0667, 0.1872 0.0702, 0.1076 0.67 and −0.82 1831163

1724.0 7.182−148.226 −14 ≤ h ≤ 24 −26 ≤ k ≤ 25 −16 ≤ l ≤ 15 1.052 0.0748, 0.2097 0.1112, 0.2358 1.731 and −0.57 1831162

R1= Σ(||F0| − |FC||)/Σ|F0|. bwR2 = [Σw(|F0|2 − |FC|2)2/Σw(F02)]1/2.

DOI: 10.1021/acs.cgd.8b01406 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Communication

Figure 1. Structural fragments of 1: (a) asymmetric structure; (b) structure of tetragonal Cu4 crown-shaped metallic-macrocycle; (c) packing diagram of structure (dashed lines represent hydrogen bonds).

Figure 2. Packing patterns of the polycatenated 1 and 2.

in good yield in the crystalline state (Scheme 2). The single crystal X-ray diffraction further proved their constructions (Figures 1−3). Complex 1 displays a monoclinic system, space group P21/c, and there are four crystallographically independent one-half CuII and two (HL)− fragments in its asymmetric unit (Table 1, Figure 1a). As shown in Figures 1 and 2, it consists of four CuII ions and four (HL)− monodeprotonated ligands. Each metal ion is thus bound to two {O2} and one {N} compartments of trigonal bipyramid ligand strands, resulting in overall fivecoordination and an {NO4} donor environment that is a more or less strongly distorted trigonal bipyramid (Figure 1b). Four oxygen atoms of two acetylacetonyl units from two H2L are closely coplanar, and one coordinated N atom from pyrazole units completes the coordination geometry with somewhat longer Cu−N bonds as shown in Figure 1. The CuII centers are placed at the four corners of the rhombus with diagonal distances of 15.101 Å (Cu1···Cu1′) and 9.074 Å (Cu2···Cu2′), and the rhombus is composed of four bridging ligands (HL)− and four metal centers (Cu2+) resulting in metal−metal−metal angles of 61.994° (Cu2−Cu1−Cu2′) and 118.006° (Cu1− Cu2−Cu1′). Pairs of copper ions with the same coordination

the help of structural analysis, we can see that the deprotonation of H2L in complexes 1, 3, and 4 was observed only in the acetylacetonyl moiety, while the ligand H2L presents two states in 2: monodeprotonation of the acetylacetonyl group and double deprotonation of its pyrazolyl and acetylacetonyl groups. X-ray diffraction and elemental analysis supply the chemical formulas of four new complexes, and PXRD measurements confirm the phase purities of the crystalline materials 1−2. As depicted in Figure S1, the experimental PXRD patterns are in accordance with simulated ones derived from the single-crystal X-ray diffraction data. The TGA technique was used to investigate the thermal stability of complexes (Figure S2). The dehydration processes of complex 2 began from 99 °C, and 1 is stable up to 240 °C. With heating, both underwent slow, continuous weightlessness until absolute decomposition. Structural Analysis and Discussion. Reacting H2L with Cu(NO3)2·6H2O in a mixed solvent at room temperature gave a different reaction mixture with different equivalents of NaHCO3 as the base (see Experimental Section). After slow evaporation, two supramolecular homometallic complexes [Cu(HL)2]4 (1) and [Cu(HL)(L)]n (2) could be obtained D

DOI: 10.1021/acs.cgd.8b01406 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Communication

Figure 3. Structural fragments of 2: (a) asymmetric structure; (b) 1D zigzag chain.

one pyrazolyl group from the third ligand (Figure 3a). In this own-shaped structure, adjacent [CuIIO4] units are bridged by pyrazole backbones of (L)2− to form an infinite zigzag chain running (Figure 3b). The distance of Cu−N is 2.281 Å, and that of CuII−O is 1.945 Å; the distance of adjacent CuII···CuII is 8.849 Å in the zigzag chain. The frameworks mutually interpenetrate with the adjoining CuII···CuII distance of 7.397 Å between CuII atoms in the adjacent frameworks. The pyrazolyl and acetylacetonyl rings of the doubly deprotonated (L)2− ligand are twisted along the pivotal C−C bond with a dihedral angle of 81.201°, while the dihedral angle of the pyrazolyl and acetylacetonyl groups in the monodeprotonated (HL)− ligand is 86.664°. In addition, the adjacent chains are stacked by weak hydrogen-bonding interactions between molecules in the stacking lattice (Figure 3c). Complex 2 contains one molecule of uncoordinated H2O per asymmetric unit. The NH fragments from pyrazolyl rings and the OH fragments from crystallized water molecules play the role of donors, and the unprotonated pyrazolyl N atom and one of the CuIIcoordinated acetylacetonyl O atoms, as acceptors for traditional hydrogen bonds involving N−H···O, O−H···N, and O− H···O; they subtend 3D construction (Figure S6 and Table S5).

environment are placed at diagonal vertices of the rhombus, resulting in a significant difference in Cu−N and Cu−O bond lengths between four copper ions and ligands. The bond lengths of Cu1−N and Cu2−N are 2.275 and 2.286 Å. The bond lengths of Cu−O range between 1.921 and 1.946 Å for all four copper centers. The former long Cu−N interactions reflect the typical Jahn−Teller elongation in the d9 CuII ion.71 Selected bond lengths and angles for 1 are provided in Table S1. The parallelogram structure of complex 1 is further assembled through inter-parallelogram hydrogen-bond interactions of N−H···N and N−H···O from uncoordinated Npyrazolyl atoms to protonated Npyrazolyl atoms and from the protonated Npyrazolyl atoms to coordinated Oacetylacetonyl atoms, resulting in a 3D supramolecular framework (Figures 1c and S5; Table S5). The X-ray structure of 2 reveals a construction of zigzag chain from H2L and Cu(NO3)2 as shown in Figures 2 and 3. Complex 2 displays the orthorhombic system, space group Fdd2 (Table 1), showing one CuII atom, one monodeprotonated ligand (HL)−, one doubly deprotonated ligand (L)2−, and one guest molecule of water in the asymmetric unit. The distorted pentahedral CuII center coordinates to two chelated acetylacetonyl groups located in a plane from two ligands and E

DOI: 10.1021/acs.cgd.8b01406 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Communication

Figure 4. Structural fragments of 3: (a) asymmetric structure and (b) packing diagram (dashed lines represent hydrogen bonds).

Figure 5. Packing patterns of the polycatenated 3 and 4.

crystallographic center of inversion (Table 2). The Fe atom is bound to three {O2} compartments of three ligand strands, resulting in overall six-coordination and an {O6} donor environment (six O atoms from three O, O′ chelating H2L, Figure 4) in a slightly twisted octahedral geometry. The bond lengths d(Fe−O) for the Fe ion range from 1.979 to 2.002 Å indicating a high-spin (HS) state of the FeIII ion (Table S3) based on the reported work of the mononuclear analogue [FeIII(acac)3]72 and heteronuclear complex [FeIII(pyacac)3(Re(bpy)(CO)3)3](OTf)3.73,74 Similarly, there are two hydrogen bondings among the CH3 group of the ligand, oxygen atoms from acetylacetonyl moiety, and nitrogen atoms from the pyrazolyl moiety. The CH3 groups play the role of donors, the O and N atoms play the role of acceptors for traditional hydrogen bonds, and the weak hydrogen-bonding network has a three-dimensional character (Figures 4b and S7; Table S5). With respect to crystal engineering, we were interested in the heterometallic supramolecular complex constructed from 3 due to the compatibility with different site symmetries. So the designed heterometallic supramolecular complex 4 will be discussed in the following content. We prepared the new heterometallic complex 4 by the reaction of solutions of 3 and AgNO3 in mixed solvents of acetonitrile and methanol. The structure of complex 4 determined by X-ray crystallography is depicted in Figures 5 and 6. 4 displays the monoclinic system, space group P21/c (Table 2). In an

Comparing the structure of 2 with that of 1, it seems that the degree of deprotonation of ligand (H2L) in the form of the (HL)− or (L)2− anion plays an important role in the dimensionalities of the frameworks. Although all acetylacetonyl oxygen atoms are deprotonated, [Cu(HL)2]4 (1) are of 2D rhombus structure due to the pyrazolyl nitrogen atoms being protonated. In contrast, both the acetylacetonyl oxygen atoms and half of pyrazolyl nitrogen atoms are deprotonated, with the structure of [Cu(HL)(L)]n (2) being a 1D zigzag chain polymer. Complex [Fe(HL)3] (3) was synthesized from the reaction between Fe(NO3)3·9H2O and 3 equiv of H2L in mixed solvent with methanol and water using saturated NaHCO3 solution as the base. More stringent conditions are needed to deprotonate the oxygen donor of the ditopic ligand owing to the lack of soft coordination partners. Iron nitrate played a suitable role as the reactant, because of the oxophilic character of the iron cation which requires the basicity of the anion. The oxygen donors from acetylacetone group as a prior reaction site are coordinated by the FeIII center first, and then the pyrazole group provides nitrogen donors to construct multimetal− organic complexes. Purified samples for the designed complex 3 would be obtained by a one-step reaction. The crystal structure of mononuclear complex 3 was given by X-ray diffraction measurement (Figures 4 and 5). The diffraction experiment revealed that 3 crystallizes in the monoclinic system, space group C2/c with Fe3+ ions sited at the F

DOI: 10.1021/acs.cgd.8b01406 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Communication

present in the HS state (Table S4). And the average bond length of Ag−N is 2.233 Å. Also of note is that the complex 4 was isolated with the nitrate anion (NO3−) as the counteranion and contains a molecule of uncoordinated MeOH. The NH groups from the pyrazolyl unit and the OH groups from the crystallized CH3OH molecule acted as donors, and the oxygen of NO3− and of the crystallized CH3OH acted as acceptors for traditional hydrogen bonds, which has a 3D construction (Figures 6d and S8; Table S5). Magnetic Susceptibility Measurements. The thermal dependence of magnetic susceptibility was done on crystalline materials of 1 and 2 using the 0.1 T field from 1.8 to 300 K with a SQUID magnetometer; χMT vs T plots of homometallic supramolecular complexes 1 and 2 are displayed in Figure 7

Figure 7. χMT vs T plot for 1 (○) and 2 (□) in the range 2−298 K. Solid lines correspond to the best fits within the spin-only formalism (see text).

(χM: the molar magnetic susceptibility). According to the structures of 1 and 2, the supraexchange interactions of all CuII centers connected by the ligand should be neglected because of the longer length of the (HL)− and (L)2− anion ligands; therefore, magnetic signatures of both complexes are determined by single-ion properties of the CuII ions. The χMT value for both complexes at 2−298 K is 0.419 cm3 mol−1 K with g = 2.119, D = 0 for 1 and 0.415 cm3 mol−1K with g = 2.104, D = 0 for 2, respectively, nearly equal to the value of 0.375 cm3 mol−1 K of the spin-only CuII center (S = 1/2, g = 2).

Figure 6. Structural fragments of 4: (a) asymmetric structure; (b) 1D ladder-like metal−organic chain; (c) its topological representation; and (d) packing patterns of the polycatenated 4 (dashed lines represent hydrogen bonds).



CONCLUSIONS In this contribution we have presented four new 3-(3,5dimethyl-1H-pyrazol-4-yl)pentane-2,4-dione-bridged 0D, 1D, and 2D supramolecular frameworks composed of homo- and heterometallic centers: [Cu(HL)2]4 (1), [Cu(HL)(L)]n (2), [Fe(HL)3] (3), and [AgFe(HL)3(NO3)]n (4). The successfully constructed structures were found to depend on the coordination geometry of the metal center, the degree of deprotonation of the ligand, and the programmable selfassembly approach. On the basis of the derived principles for the construction of supramolecular structures, we are approaching the preparation of novel supramolecules with designed architecture and promising properties when suitable ligands or substrates are involved. The research data will embrave us to design and synthesize more supramolecular structures with a wide range of promising applications in many

asymmetric unit, there are three ligands, one FeIII atom, one Ag(I) atom, one uncoordinated MeOH molecule, and one nitrate as the counteranion, as shown in Figure 6a. The Fe atom’s coordination behavior is the same as that in the case of the atom in structure 3. All AgI ions are bridged by tridentate [Fe(HL)3] building units, and a 1D porous ladder-like structure is successfully constructed. The Ag atom is connected directly with three N atoms of different pyrazolyl fragments. Finally, 4 formed infinite 1D chain coordination networks (Figure 6b). In this structure, the diagonal metal ions are separated by 11.834 and 14.161 Å of Fe···Fe and 13.440 Å/ 11.017 Å of Ag···Ag within the ladder-like molecule, and the adjacent Fe···Ag distance is 8.817 Å/9.122 Å (Figure 6c). The Fe−O bond lengths of 4 (1.973 Å to 2.020 Å) are comparable to those found in 3, indicating that the FeIII centers are also G

DOI: 10.1021/acs.cgd.8b01406 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Communication

(9) Hill, J. A.; Thompson, A. L.; Goodwin, A. L. Dicyanometallates as Model Extended Frameworks. J. Am. Chem. Soc. 2016, 138, 5886− 5896. (10) Clemente-León, M.; Coronado, E.; Gómez-García, C.; SorianoPortillo, A. Increasing the Ordering Temperatures in Oxalate-Based 3D Chiral Magnets: the Series [Ir(ppy)2(bpy)][MIIMIII(ox)3]· 0.5H2O (MIIMIII = MnCr, FeCr, CoCr, NiCr, ZnCr, MnFe, FeFe); bpy = 2,2‘-Bipyridine; ppy = 2-Phenylpyridine; ox = Oxalate Dianion). Inorg. Chem. 2006, 45, 5653−5660. (11) Halper, S. R.; Do, L.; Stork, J. R.; Cohen, S. M. Topological Control in Heterometallic Metal-Organic Frameworks by Anion Templating and Metalloligand Design. J. Am. Chem. Soc. 2006, 128, 15255−15268. (12) Liu, G.; Zeller, M.; Su, K.; Pang, J.; Ju, Z.; Yuan, D.; Hong, M. Controlled Orthogonal Self-Assembly of Heterometal-Decorated Coordination Cages. Chem. - Eur. J. 2016, 22, 17345−17350. (13) Metherell, A. J.; Ward, M. D. Stepwise assembly of an adamantoid Ru4Ag6 cage by control of metal coordination geometry at specific sites. Chem. Commun. 2014, 50, 10979−10982. (14) Galan, A.; Ballester, P. Stabilization of reactive species by supramolecular encapsulation. Chem. Soc. Rev. 2016, 45, 1720−1727. (15) Pan, M.; Wu, K.; Zhang, J. H.; Su, C. Y. Chiral metal−organic cages/containers (MOCs): From structural and stereochemical design to applications. Coord. Chem. Rev. 2019, 378, 333. (16) Brown, C. J.; Toste, F. D.; Bergman, R. G.; Raymond, K. N. Supramolecular Catalysis in Metal−Ligand Cluster Hosts. Chem. Rev. 2015, 115, 3012−3035. (17) Therrien, B.; Süss-Fink, G.; Govindaswamy, P.; Renfrew, A. K.; Dyson, P. J. The “Complex-in-a-Complex” Cations [(acac)2M⊂Ru6(p-iPrC6H4Me)6(tpt)2(dhbq)3]6+: A Trojan Horse for Cancer Cells. Angew. Chem., Int. Ed. 2008, 47, 3773−3776. (18) Lewis, J. E. M.; Gavey, E. L.; Cameron, S. A.; Crowley, J. D. Stimuli-responsive Pd2L4 metallosupramolecular cages: towards targeted cisplatin drug delivery. Chem. Sci. 2012, 3, 778−784. (19) Samanta, S. K.; Moncelet, D.; Briken, V.; Isaacs, L. MetalOrganic Polyhedron Capped with Cucurbit[8]uril Delivers Doxorubicin to Cancer Cells. J. Am. Chem. Soc. 2016, 138, 14488−14496. (20) Kaphan, D. M.; Levin, M. D.; Bergman, R. G.; Raymond, K. N.; Toste, F. D. A supramolecular microenvironment strategy for transition metal catalysis. Science 2015, 350, 1235−1238. (21) Neelakandan, P. P.; Jiménez, A.; Thoburn, J. D.; Nitschke, J. R. An Autocatalytic System of Photooxidation-Driven Substitution Reactions on a FeII4L6 Cage Framework. Angew. Chem., Int. Ed. 2015, 54, 14378−14382. (22) Li, H.; Han, Y. F.; Lin, Y. J.; Guo, Z. W.; Jin, G. X. Stepwise Construction of Discrete Heterometallic Coordination Cages Based on Self-Sorting Strategy. J. Am. Chem. Soc. 2014, 136, 2982−2985. (23) Sanz, S.; O’Connor, H. M.; Marti-Centelles, V.; Comar, P.; Pitak, M. B.; Coles, S. J.; Lorusso, G.; Palacios, E.; Evangelisti, M.; Baldansuren, A.; Chilton, N. F.; Weihe, H.; McInnes, E. J. L.; Lusby, P. J.; Piligkos, S.; Brechin, E. K. [MIII2MII3]n+ trigonal bipyramidal cages based on diamagnetic and paramagnetic metalloligands. Chem. Sci. 2017, 8, 5526−5532. (24) Sanz, S.; O’Connor, H. M.; Pineda, E. M.; Pedersen, K. S.; Nichol, G. S.; Monsted, O.; Weihe, H.; Piligkos, S.; McInnes, E. J.; Lusby, P. J.; Brechin, E. K. [CrIII8MII6]12+ Coordination Cubes (MII=Cu, Co). Angew. Chem., Int. Ed. 2015, 54, 6761−6764. (25) Cooper, B. G.; Napoline, J. W.; Thomas, C. M. Catalytic Applications of Early/Late Heterobimetallic Complexes. Catal. Rev.: Sci. Eng. 2012, 54, 1−40. (26) Yan, X.; Wang, H.; Hauke, C. E.; Cook, T. R.; Wang, M.; Saha, M. L.; Zhou, Z.; Zhang, M.; Li, X.; Huang, F.; Stang, P. J. A Suite of Tetraphenylethylene-Based Discrete Organoplatinum(II) Metallacycles: Controllable Structure and Stoichiometry, AggregationInduced Emission, and Nitroaromatics Sensing. J. Am. Chem. Soc. 2015, 137, 15276−15286. (27) Brutschy, M.; Schneider, M. W.; Mastalerz, M.; Waldvogel, S. R. Porous Organic Cage Compounds as Highly Potent Affinity

areas, such as supramolecular catalysis, enzyme mimicking, cavity-directed chemical reaction, sensing, and so on.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.cgd.8b01406. PXRD and IR spectra, TGA picture, tables of selected bond lengths and angles (PDF) Accession Codes

CCDC 1831162−1831165 contain the supplementary crystallographic data for this paper. These data can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif, or by emailing [email protected], or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336033.



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. ORCID

Shu-Yan Yu: 0000-0002-9010-1424 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was financially supported by Beijing Natural Science Foundation (No. 2184096), Beijing Municipal Natural Science Fund Project of Beijing Education Committee Science and Technology Project (No. KZ201710005001), Beijing Municipal High Level Innovative Team Building Program, Seed capital project of Jinqiao project of Beijing science and Technology Association (no. JQ17054), National Natural Science Foundations of China (Nos. 21471011, 91622102), and Basic Research Foundation of Beijing University of Technology (No. 005000514118547).

■ ■

DEDICATION Dedicated to Xin-Tao Wu on the occasion of his 80th birthday. REFERENCES

(1) Lee, E.; Seo, S. J.; Lee, S. S.; Lindoy, L. F. Assembling latter dblock heterometal coordination polymers: Synthetic strategies and structural outcomes. Coord. Chem. Rev. 2017, 348, 121−170. (2) Cook, T. R.; Stang, P. J. Recent Developments in the Preparation and Chemistry of Metallacycles and Metallacages via Coordination. Chem. Rev. 2015, 115, 7001−7045. (3) Li, H.; Yao, Z. J.; Liu, D.; Jin, G. X. Multi-component coordination-driven self-assembly toward heterometallic macrocycles and cages. Coord. Chem. Rev. 2015, 293-294, 139−157. (4) Devic, T.; Serre, C. Article High valence 3p and transition metal based MOFs. Chem. Soc. Rev. 2014, 43, 6097−6115. (5) Smulders, M. M. J.; Riddell, I. A.; Browne, C.; Nitschke, J. R. Building on architectural principles for three-dimensional metallosupramolecular construction. Chem. Soc. Rev. 2013, 42, 1728−1754. (6) Przychodzeń, P.; Korzeniak, T.; Podgajny, R.; Sieklucka, B. Supramolecular coordination networks based on octacyanometalates: From structure to function. Coord. Chem. Rev. 2006, 250, 2234−2260. (7) Hidai, M. Chemical nitrogen fixation by molybdenum and tungsten complexes. Coord. Chem. Rev. 1999, 99 (185-186), 99−108. (8) Saalfrank, R. W.; Maid, H.; Scheurer, A. Supramolecular Coordination Chemistry: The Synergistic Effect of Serendipity and Rational Design. Angew. Chem., Int. Ed. 2008, 47, 8794−8824. H

DOI: 10.1021/acs.cgd.8b01406 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Communication

Materials for Sensing by Quartz Crystal Microbalances. Adv. Mater. 2012, 24, 6049−6052. (28) Kaloudi-Chantzea, A.; Karakostas, N.; Raptopoulou, C. P.; Psycharis, V.; Saridakis, E.; Griebel, J.; Hermann, R.; Pistolis, G. Coordination-Driven Self Assembly of a Brilliantly Fluorescent Rhomboid Cavitand Composed of Bodipy-Dye Subunits. J. Am. Chem. Soc. 2010, 132, 16327−16329. (29) Liu, Y.; Perez, L.; Mettry, M.; Easley, C. J.; Hooley, R. J.; Zhong. Self-Aggregating Deep Cavitand Acts as a Fluorescence Displacement Sensor for Lysine Methylation. J. Am. Chem. Soc. 2016, 138, 10746−10749. (30) Liu, S.; Gan, H.; Hermann, A. T.; Rick, S. W.; Gibb, B. C.; Liu, S.; Gan, H.; Hermann, A. T.; Rick, S. W.; Gibb, B. C. Nat. Chem. 2010, 2, 847−852. (31) Mondal, B.; Acharyya, K.; Howlader, P.; Mukherjee, P. S. Molecular Cage Impregnated Palladium Nanoparticles: Efficient, Additive-Free Heterogeneous Catalysts for Cyanation of Aryl Halides. J. Am. Chem. Soc. 2016, 138, 1709−1716. (32) Horiuchi, S.; Murase, T.; Fujita, M. A Remarkable Organometallic Transformation on a Cage-Incarcerated Dinuclear Ruthenium Complex. Angew. Chem., Int. Ed. 2012, 51, 12029−12031. (33) Breiner, B.; Clegg, J. K.; Nitschke, J. R. Reactivity modulation in container molecules. Chem. Sci. 2011, 2, 51−56. (34) Mal, P.; Breiner, B.; Rissanen, K.; Nitschke, J. R. White Phosphorus Is Air-Stable Within a Self-Assembled Tetrahedral Capsule. Science 2009, 324, 1697−1699. (35) Yamashina, M.; Sei, Y.; Akita, M.; Yoshizawa, M. Safe Storage of Radical Initiators within a Polyaromatic Nanocapsule. Nat. Commun. 2014, 5, 4662−4668. (36) Brenner, W.; Ronson, T. K.; Nitschke, J. R. Separation and Selective Formation of Fullerene Adducts within an MII8L6 Cage. J. Am. Chem. Soc. 2017, 139, 75−78. (37) Holloway, L. R.; McGarraugh, H. H.; Young, M. C.; Sontising, W.; Beran, G. J.; Hooley, R. J. Structural switching in self-assembled metal-ligand helicate complexes via ligand-centered reactions. Chem. Sci. 2016, 7, 4423−4427. (38) Custelcean, R. Anion encapsulation and dynamics in selfassembled coordination cages. Chem. Soc. Rev. 2014, 43, 1813−1824. (39) Ward, M. D.; Raithby, P. R. Functional behaviour from controlled self-assembly: challenges and prospects. Chem. Soc. Rev. 2013, 42, 1619−1636. (40) Wang, S.; Zuo, J. L.; Zhou, H. C.; Choi, H. J.; Ke, Y.; Long, J. R.; You, X. Z. [(Tp)8(H2O)6CuII6FeIII8(CN)24]4+: A CyanideBridged Face-Centered-Cubic Cluster with Single-Molecule-Magnet Behavior. Angew. Chem., Int. Ed. 2004, 43, 5940−5943. (41) Sanz, S.; O’Connor, H. M.; Pineda, E. M.; Pedersen, S.; Nichol, G. S.; Mønsted, O.; Weihe, H.; Piligkos, S. E.; McInnes, J. L.; Lusby, P. J.; Brechin, E. K. [CrIII8MII6]12+ Coordination Cubes (MII = Cu, Co). Angew. Chem., Int. Ed. 2015, 54, 6761−6764. (42) Deria, P.; Mondloch, J. E.; Karagiaridi, O.; Bury, W.; Hupp, J. T.; Farha, O. K. Beyond post-synthesis modification: evolution of metal−organic frameworks via building block replacement. Chem. Soc. Rev. 2014, 43, 5896−5912. (43) Yu, S. Y.; Li, S. H.; Huang, H. P.; Zhang, Z. X.; Jiao, Q.; Shen, H.; Hu, X. X.; Huang, H. Molecular Self-Assembly with Modularization and Directionality: Vector-Manipulation at Metal Centers. Curr. Org. Chem. 2005, 9, 555−563. (44) Yu, S. Y.; Huang, H. P.; Li, S. H.; Jiao, Q.; Li, Y. Z.; Wu, B.; Sei, Y.; Yamaguchi, K.; Pan, Y. J.; Ma, H. W. Solution Self-Assembly, Spontaneous Deprotonation, and Crystal Structures of BipyrazolateBridged Metallomacrocycles with Dimetal Centers. Inorg. Chem. 2005, 44, 9471−9488. (45) Jiang, X. F.; Deng, W.; Jin, R.; Qin, L.; Yu, S. Y. Self-Assembly of Tri-Pyrazolate Linked Cages with Di-Palladium Coordination Motifs. Dalton Trans 2014, 43, 16015−16024. (46) Cui, Y.; Chen, Z. M.; Jiang, X. F.; Tong, J.; Yu, S. Y. Selfassembly and anion sensing of metal-organic [M6L2] cages from fluorescent triphenylamine tri-pyrazoles with dipalladium(II,II) corners. Dalton Trans 2017, 46, 5801−5805.

(47) Ward, M. D.; McCleverty, J. A.; Jeffery, J. C. Coordination and supramolecular chemistry of multinucleating ligands containing two or more pyrazolyl-pyridine ‘arms’. Coord. Chem. Rev. 2001, 222, 251− 272. (48) Trofimenko, S. The Coordination Chemistry of PyrazoleDerived Ligands. In Progress in Inorgic Chemistry; Lippard, S. J., Ed.; John Wiley & Sons, Inc.: 1986; Vol. 34, pp 115−206. (49) Trofimenko, S. The Coordination Chemistry of Pyrazolederived Ligands. Chem. Rev. 1972, 72, 497−509. (50) Pettinari, C.; Santini, C. In Comprehensive Coordination Chemistry II, From Biology to Nanotechnology; McCleverty, J. A., Meyer, T. J., Eds.-In-Chief; Elsevier Ltd.: 2004; Vol. 1, pp 159−210. (51) Klingele, J.; Dechert, S.; Meyer, F. Polynuclear transition metal complexes of metal···metal-bridging compartmental pyrazolate ligands. Coord. Chem. Rev. 2009, 253, 2698−2741. (52) Trofimenko, S. The Coordination Chemistry of Poly(pyrazolyl) borate Ligands; Imperial Colledge Press: 1999, London. (53) Leibeling, G.; Demeshko, S.; Dechert, S.; Meyer, F. Hysteretic Magnetic Bistability Based on a Molecular Azide Switch. Angew. Chem., Int. Ed. 2005, 44, 7111−7114. (54) Duan, P. C.; Manz, D. H.; Dechert, S.; Demeshko, S.; Meyer, F. Reductive O2 Binding at a Dihydride Complex Leading to Redox Interconvertible μ-1,2-Peroxo and μ-1,2-Superoxo Dinickel(II) Intermediates. J. Am. Chem. Soc. 2018, 140, 4929−4939. (55) Li, L.; Zhang, Y.; Avdeev, M.; Lindoy, L. F.; Harman, D. G.; Zheng, R. K.; Cheng, Z. X.; Aldrich-Wright, J. R.; Li, F. Self-assembly of a unique 3d/4f heterometallic square prismatic box-like coordination cage. Dalton Trans 2016, 45, 9407−9411. (56) Blusch, L. K.; Craigo, K. E.; Martin-Diaconescu, V.; McQuarters, A. B.; Bill, E.; Dechert, S.; DeBeer, S.; Lehnert, N.; Meyer, F. Hidden Non-Innocence in an Expanded Porphyrin: Electronic Structure of the Siamese-Twin Porphyrin’s Dicopper Complex in Different Oxidation States. J. Am. Chem. Soc. 2013, 135, 13892−13899. (57) Mulyana, Y.; Kepert, C. J.; Lindoy, L. F.; Parkin, A.; Turner, P. New cobalt(II) and zinc(II) coordination frameworks incorporating a pyridyl−pyrazole ditopic ligand. Dalton Trans 2005, 9, 1598−1601. (58) Yang, G.; Raptis, R. G.; Safar, P. Cadmium(II) Complexes of 4(4-Pyridyl)Pyrazole: A Case of Two Conformational Supramolecular Isomers Polythreaded in the Same Crystal and a Rare Example of 5Connected nov Net. Cryst. Growth Des. 2008, 8, 981−985. (59) Lu, J. Z.; Turner, D. R.; Harding, L. P.; Batten, S. R. Heteroligand Molecular “Stirrups” Using Conformationally Flexible Ditopic Pyridyl−Pyrazolyl Ligands. Inorg. Chem. 2009, 48, 7525− 7527. (60) Tong, J.; Yu, S. Y.; Li, H. Programmable self-assembly of homoor hetero-metallomacrocycles using 4-(1H-pyrazolyl-4-yl)pyridine. Chem. Commun. 2012, 48, 5343−5345. (61) Li, F.; Lin, R. B.; Wei, Y. S.; Liao, P. Q.; Bai, J.; Xue, W.; Zhang, W.; Zhang, J. P.; Chen, X. M. Metal-ion controlled solid-state reactivity and photoluminescence in two isomorphous coordination polymers. Inorg. Chem. Front. 2014, 1, 172−177. (62) Guo, Q. Q.; Englert, U. An Acetylacetonate or a Pyrazole? Both! 3-(3,5-Dimethyl-pyrazol-4-yl)pentane-2,4-dione as a Ditopic Ligand. Cryst. Growth Des. 2016, 16, 5127−5135. (63) Chen, Z. M.; Cui, Y.; Jiang, X. F.; Tong, J.; Yu, S. Y. Programmable self-assembly of water-soluble organo-heterometallic cages [M12M’4L12] using 3-(3,5-dimethyl-1H-pyrazol-4-yl)pentane2,4-dione (H2L). Chem. Commun. 2017, 53, 4238−4241. (64) Sun, W. Q.; Tong, J.; Lu, H.; Ma, T. T.; Ma, H. W.; Yu, S. Y. Programmable Self-Assembly of Heterometallic Palladium(II)Copper(II) 1D Grid-Chain using Dinuclear Palladium(II) Corners with Pyrazole-Carboxylic Acid Ligands. Chem. - Asian J. 2018, 13, 1108−1113. (65) Wang, C.; Liu, C.; Tian, H. R.; Li, L. J.; Sun, Z. M. Supramolecular Structures of Titanium(IV)-Substituted Wells-Dawson Polyoxotungstates. Chem. - Eur. J. 2018, 24, 2952−2961. (66) Ye, Z. M.; He, C. T.; Xu, Y. T.; Krishna, R.; Xie, Y.; Zhou, D. D.; Zhou, H. Lo.; Zhang, J. P.; Chen, X. M. A New Isomeric Porous I

DOI: 10.1021/acs.cgd.8b01406 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Communication

Coordination Framework Showing Single-Crystal to Single-Crystal Structural Transformation and Preferential Adsorption of 1,3Butadiene from C4 Hydrocarbons. Cryst. Growth Des. 2017, 17, 2166−2171. (67) He, C. T.; Liao, P. Q.; Zhou, D. D.; Wang, B. Y.; Zhang, W. X.; Zhang, J. P.; Chen, X. M. Visualizing the distinctly different crystal-tocrystal structural dynamism and sorption behavior of interpenetrationdirection isomeric coordination networks. Chem. Sci. 2014, 5, 4755− 4762. (68) Padial, N. M.; Quartapelle, P. E.; Montoro, C.; Lopez, E.; Oltra, J. E.; Colombo, V.; Maspero, A.; Masciocchi, N.; Galli, S.; Senkovska, I. Highly Hydrophobic Isoreticular Porous Metal−Organic Frameworks for the Capture of Harmful Volatile Organic Compounds. Angew. Chem., Int. Ed. 2013, 52, 8290−8294. (69) Charles, R. G. Working with Hazardous Chemicals. Org. Synth. 1959, 39, 61−63. (70) Dolomanov, O. V.; Bourhis, L. J.; Gildea, R. J.; Howard, J. A. K.; Puschmann, H. OLEX2: a complete structure solution, refinement and analysis program. J. Appl. Crystallogr. 2009, 42, 339−341. (71) Van der Vlugt, J. I.; Demeshko, S.; Dechert, S.; Meyer, F. Tetranuclear CoII, MnII, and CuII Complexes of a Novel Binucleating Pyrazolate Ligand Preorganized for the Self-Assembly of Compact [2 × 2]-Grid Structures. Inorg. Chem. 2008, 47, 1576−1585. (72) Pernicone, N. C.; Geri, J. B.; York, J. T. Using a Combination of Experimental and Computational Methods To Explore the Impact of Metal Identity and Ligand Field Strength on the Electronic Structure of Metal Ions. J. Chem. Educ. 2011, 88, 1323−1327. (73) Knight, T. E.; Guo, D.; Claude, J. P.; McCusker, J. K. Energy Transfer Dynamics in ReI−Based Polynuclear Assemblies: A Quantitative Application of Förster Theory. Inorg. Chem. 2008, 47, 7249−7261. (74) Vreshch, V. D.; Lysenko, A. B.; Chernega, A. N.; Howard, J. A. K.; Krautscheid, H.; Sieler, J.; Domasevitch, K. V. Extended coordination frameworks incorporating heterobimetallic squares. Dalton Trans 2004, 2899−2903.

J

DOI: 10.1021/acs.cgd.8b01406 Cryst. Growth Des. XXXX, XXX, XXX−XXX