based metal-organic frameworks - ACS Publications

multifunctional molecular materials.6–12 Besides the rational design of ..... found in Table 1 and in the deposited files at Cambridge Crystallograp...
0 downloads 0 Views 2MB Size
Subscriber access provided by Chalmers Library

Article

Influence of the metal ion in the topology and interpenetration of pyridylvinyl(benzoate) based metal-organic frameworks Yuri Dezotti, Marcos Antônio Ribeiro, Kleber R. Pirota, and Wdeson Pereira Barros Cryst. Growth Des., Just Accepted Manuscript • DOI: 10.1021/acs.cgd.9b00552 • Publication Date (Web): 21 Aug 2019 Downloaded from pubs.acs.org on August 25, 2019

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 46 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

Influence of the metal ion in the topology and interpenetration of pyridylvinyl(benzoate) based metal-organic frameworks Yuri Dezotti,a Marcos Antônio Ribeiro,b Kleber R. Pirota,c Wdeson Pereira Barrosa* aInstituto

de Química, Universidade Estadual de Campinas, Departamento de Química

Inorgânica, Campinas, SP 13083-970, Brazil bDepartamento

de Química, Universidade Federal do Espírito Santo, Vitória, ES 29075-

910, Brazil. aInstituto

de Física Gleb Wataghin, Universidade Estadual de Campinas, Departamento

de Física da Matéria Condensada, Campinas, SP 13083-970, Brazil KEYWORDS: metal-organic frameworks, pyridylvinyl(benzoate); magnetic properties.

ABSTRACT

A family of four M(II)-MOFs of general formula {[Mx(pvb)2x]·y(dmf)}n (M=Cu, 1; M=Co, 2; M=Ni, 3; M=Mn, 4), based on the bis{4-[2-(4-pyridyl)ethenyl]} benzoic acid (Hpvb)

ACS Paragon Plus Environment

1

Crystal Growth & Design 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 46

ligand were obtained. 1 exhibits a fivefold interpenetrated lvt framework, 2 and 3 a sevenfold interpenetrating dia framework, and 4 a twofold interpenetrated dmc framework. Magnetic properties of 1-4 have been investigated. 1 was analyzed by a Curie-Weiss model, while 2 and 3 where analyzed by a ZFS model due to the very long metal-metal distances, which results in very weak antiferromagnetic interactions. The coupling pathway in 4 is done by carboxylate bridges instead of the pvb pathway, affording short metal-metal separation that was analyzed by a isotropic Heisenberg Spin Hamiltonian for a linear trinuclear Mn(II) cluster. The different metal coordination modes and geometries, along with template effects induced by the solvent, play an important role in the formation of distinctive structural topologies.

INTRODUCTION Metal-organic frameworks (MOFs) have been widely investigated due to their complex structural arrangements and topologies, which can directly influence their optical, magnetic or electronic physical properties.1–5 Along with their wide range of chemical properties like storage, separation or transport of gases, catalysis, controlled release of drug molecules, this flexibility can lead to the design of very interesting multifunctional molecular materials.6–12 Besides the rational design of distinctive MOFs involves the control of metal/ligand geometries and ligand shapes, by either a modular,13,14 molecular building block15–17 or secondary building unit approach,18–20 the

ACS Paragon Plus Environment

2

Page 3 of 46 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

difficulty to establish the exact chemical and physical conditions may introduce some limitation on the prediction of MOFs with different dimensionalities, topologies and pore sizes and shapes.21,22 Despite this lack of predictability, supramolecular chemistry offers a great possibility for the development of new materials where intermolecular interactions between molecular building blocks can be explored, making feasible the formation of versatile systems due to a great variety of possible structural arrangements.23,24 Due to some sort of unpredictable structure variations, the study of parameters that controls the formation of different secondary building units and interpenetration is of great value for the understanding of topological variations from the same primary building units. In this work, we show how changes on transition metal ion affect the 3D structure by the influence of the secondary building unit on the definition of the net and also on the degree of net interpenetration. Four 3D coordination polymers were synthesized through the reaction between M(II) (M = Cu, Ni,

Co, and Mn) and Bis{4-[2-(4-

pyridyl)ethenyl]} benzoic acid (Hpvb), in N,N-dimethylformamide (dmf) using solvothermal procedure, giving rise to a product of general formula {[Mx(pvb)2x]· y(dmf)}n. Depending on the metal ion, MOFs with distinct topologies was obtained: a Cu(II) complex presenting a fivefold interpenetrated lvt framework, Co(II) and Ni(II) complexes exhibiting a sevenfold interpenetrating dia framework, and a Mn(II) complex with a twofold interpenetrated dmc framework.

ACS Paragon Plus Environment

3

Crystal Growth & Design 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 46

EXPERIMENTAL SECTION Materials and Physical Measurements. All reagents and solvents employed were purchased from commercial sources and used as received without further purification. Elemental analyses were obtained using a Perkin Elmer model CHN2400 elemental analyzer. Infrared spectra were recorded on an Agilent Cary Model 630 spectrometer in the wavenumber range of 4000–400 cm−1 with an average of 64 scans and 4 cm−1 spectral resolution using an attenuated total reflection (ATR) apparatus. 1H and

13C

NMR

spectra were recorded on a Bruker model Avance III instrument (500 MHz) in dmso-d6. Single Crystal X-ray diffraction data (SCXRD) were obtained on Brazilian Synchrotron Light Laboratory (LNLS), in MX2 Beamline at 100 K. Powder X-ray diffraction patterns (PXRD) were acquired using a Shimadzu XDR Model 7000 diffractometer using Cu K radiation, 2o /min scan rate, 298 K, and  = 1.54184 nm. Temperature dependent magnetic susceptibility was obtained on polycrystalline samples using a Quantum Design MPMS XL-7 SQUID magnetometer under an applied dc magnetic field of 0.1 T over a temperature range of 2.0−300 K. Diamagnetic corrections were applied to the observed paramagnetic susceptibilities using Pascal’s constants.25,26 Thermogravimetric analysis (TGA) was carried out on a Thermal Analysis Instrument, model TGA 2050, in a dynamic argon atmosphere (100 mL min-1), in an alumina crucible over a temperature

ACS Paragon Plus Environment

4

Page 5 of 46 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

range of 25-800 °C with a heating rate of 10 °C min-1. Bis{4-[2-(4-pyridyl)ethenyl]} benzoic acid, Hpvb, was prepared according to a previously reported procedure.27 A mixture of 4-methylpyridine (9.8 mL, 100.0 mmol) and 4-formylbenzoic acid (15.2 g, 100.0 mmol) was dissolved in 80 mL of acetic anhydride. Under vigorous stirring, the system was refluxed for 8 hours. After this period, the system was cooled to room temperature, and then vacuum filtered and washed with water, cold ethanol, and cold diethyl ether. The resulting yellow solid was dried under vacuum for 3 hours. Yield: ca. 65%. Anal. Calcd. for C14H11NO2 (%): C, 74.65; H, 4.92; N, 6.22. Found: C, 74.49; H, 4.78; N, 6.20. IR/cm-1: 1665 [C=O], 3443 [O-H], 970 [C-H (vinyl)]. 1H NMR (500 MHz, dmso-d6)  (ppm), 7.98, 7.96, 7.78 and 7.76, aromatic rings, 7.38 and 7.41 hydrogen of vinyl group, 8.5 acid hydroxyl.

13C

NMR (dmso-d6, 125 MHz)  (ppm) 166.9, 150.1,

143.8, 140.3, 131.9, 130.5, 129.6, 128.4, 127.0 and 121.0.

Synthesis of {[Cu(pvb)2]·2(dmf)}n (1·dmf·2H2O). A solution of CuCl2·2H2O (0.6819 g, 4 mmol) in 10 mL of dmf was added dropwise in a suspension of Hpvb (1.7984 g, 8 mmol) in 60 mL of dmf at 90 oC. The resulting light green suspension was kept under continuous stirring at 90 oC for 45 minutes. After this period, the suspension was filtered and the solid was placed for drying for 5 days in the open air. A suspension of the resulting light green solid (0.3099 g, 0.5 mmol) in 12 mL of dmf was placed in a teflon-lined stainless steel autoclave and heated (solvothermal reactor) at 140 °C for 3

ACS Paragon Plus Environment

5

Crystal Growth & Design 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 46

days. Subsequent cooling to room temperature at a rate of 20 oC per hour afforded dark purple plate-like crystals of 1. Yield: ca. 38%; Anal. Calcd. for C37H45CuN5O9 (1·dmf·2H2O) (%): C, 57.91; H, 5.91; N, 9.13. Found: C, 57.06; H, 6.05; N, 9.21. IR/cm-1: 1599 [C=O (w)], 956 [C-H (vinyl) (m)], 1679 [C=C (vinyl) (s)]. Synthesis of {[Co(pvb)2]·(dmf)}n (2). A solution of CoCl2·6H2O (0.1455 g, 0.5 mmol) in 6 mL of dmf was added dropwise in a suspension of Hpvb (0.2248 g, 1 mmol) in 6 mL of dmf at room temperature. The resulting pink suspension was kept under continuous stirring for 10 minutes. After this period, the suspension was placed in a teflon-lined stainless steel autoclave and heated (solvothermal reactor) at 140 °C for 3 days. Subsequent cooling to room temperature at a rate of 20 oC per hour afforded dark red plate-like crystals of 2. Yield: ca. 86%; Anal. Calcd. for C31H27CoN3O5 (2) (%): C, 64.14; H, 4.69; N, 7.24. Found: C, 63.34; H, 4.79; N, 6.94. IR/cm-1: 1582 [C=O (m)], 959 [C-H (vinyl) (m)], 1654 [C=C (vinyl) (w)]. Synthesis of {[Ni(pvb)2]·(dmf)}n (3·0.5H2O). The procedure was the same as that for 2, using NiCl2·4H2O (0.1188 g, 0.5 mmol). After cooling under the same conditions, green plate-like crystals were formed. Yield: ca. 64%; Anal. Calcd. for C31H28NiN3O5.5 (3) (%): C, 63.19; H, 4.79; N, 7.13. Found: C, 61.83; H, 4.64; N, 6.99. IR/cm-1: 1606 [C=O (s)], 952 [C-H (vinyl) (w)], 1654 [C=C (vinyl) (m)]. Synthesis of {[Mn3(pvb)6]·2(dmf)}n (4·H2O). The procedure was the same as that for 2, using Mn(H3CCOO)2·4H2O (0.1225 g, 0,5 mmol). After cooling under the same

ACS Paragon Plus Environment

6

Page 7 of 46 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

conditions, orange block-like crystals were formed. Yield: ca. 14%; Anal. Calcd. for C90H76Mn3N8O15 (4) (%): C, 64.56; H, 4.57; N, 6.69. Found: C, 62.09; H, 4.64; N, 6.51. IR/cm-1: 1593 [C=O (s)], 950 [C-H (vinyl) (m)], 1664, 1654 [C=C (vinyl) (w)].

X Ray Data collection and structural refinement. Single Crystal XRD data for 1-4 were collected on MX2 beamline at Brazilian Synchrotron Light Laboratory.28 MX2 is a wiggler beamline dedicated to Macromolecular Crystallography at the UVX synchrotron source at the Brazilian Synchrotron Light Source. It operates on a 2.0 T hybrid 30-pole wiggler and its optical layout includes collimating mirror, Si(111) double-crystal monochromator and toroidal bendable mirror. The MX2 beamline provides wide tunability between 5 and 15 keV with maximum flux at 8.5 keV. The beamline is equipped with PILATUS2M detector from Dectris and a mini-Kappa goniometer from Arinax. The frames were collected using phi-scans (1°/frame) and the crystals were kept at 100 ± 2 K during all experiment. Reflection indexing, unit-cell parameters refinement, integration and corrections were performed by CCP4,29 XIA2 0.5.653-g9f819c0c-dials-1.11,30 DIALS 1.11.2-g01fb9e997release.31 The data merging and scale were performed using Aimless32 and Pointless.33 The structures were solved within Olex2,34 with the ShelXT35 structure solution program using Intrinsic Phasing and refined with the XL36 refinement package using Least Squares minimization. The position of all non-hydrogen atoms were refined

ACS Paragon Plus Environment

7

Crystal Growth & Design 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 46

anisotropically. The hydrogen atoms in the compounds were added in the structure in idealized positions and further refined according to the riding model, Uiso(H) = 1.5Ueq(C) for sp2 carbon and Uiso(H) = 1.2Ueq(C) for sp3 carbons. Additional information concerning measurement and refinement data for compounds 1-4 are found in Table 1 and in the deposited files at Cambridge Crystallographic Data Centre (CCDC). Any request to the CCDC for this material should quote the full literature citation and the reference numbers CCDC 1895216-1895219 for compounds 1-4, respectively.

Table 1. Crystallographic details for 1-4 Compound

1

2

3

4

Formula

C34H34CuN4O6

C31H27CoN3O5

C31H27NiN3O5

C90H74Mn3N8O14

Formula weight

658.20

580.50

580.26

1656.39

Crystal system

tetragonal

monoclinic

monoclinic

monoclinic

Space group

I41/a

Cc

Cc

P21/c

a (Å)

38.7093(3)

13.31352(10)

12.50064(8)

13.74094(10)

b (Å)

38.7093(3)

24.4807(2)

23.31313(15)

18.50072(12)

c (Å)

5.46170(10)

8.57156(7)

8.21615(6)

15.90427(11)

α (o)

90

90

90

90

β (o)

90

99.1319(7)

99.2655(6)

103.3711(6)

γ (o)

90

90

90

90

ACS Paragon Plus Environment

8

Page 9 of 46 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

V (Å3)

8183.9(2)

2758.28(3)

2363.185(19)

3933.54(3)

Z

2

4

4

2

Wavelength (Å)

0.82655

0.82655

0.82655

0.77491

T (K)

100.0

100.0

100.0

100.0

calc(mg m-3)

0.827

1.398

1.631

1.398

. (mm-1)

0.859

1.003

1.098

0.822

F(000)

2744

1204

1208

1714

Theta range(°)

1.7–28.0

1.9–31.1

1.9–29.0

1.6–31.1

Refl. collected

32097

24428

22994

68498

Independent refl.

3549[Rint=0.067] 4962[Rint=0.076] 4200[Rint=0.036] 7921[Rint=0.040]

Completeness to 95  max (%)

98

97

99

Data, restraints, 3549, 288, 283 parameters

4962, 2, 363

4200, 2, 363

7921, 0, 522

Goodness-of-fit on F2

1.57

1.17

1.08

1.10

R, wR [I > 2σ(I)]

0.115, 0.341

0.038, 0.081

0.025, 0.066

0.043, 0.105

R, wR (all data)

0.126, 0.356

0.053, 0.118

0.026, 0.067

0.048, 0.110

Largest diff. peak 1.14, -1.15 and hole (e.Å-3)

0.81, -0.54

0.66, -0.40

0.43, -0.86

CCDC

1895217

1895218

1895219

1895216

RESULTS AND DISCUSSION

ACS Paragon Plus Environment

9

Crystal Growth & Design 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 46

Compounds 1-4 presented three different morphologies, being 2 and 3 isomorphous. Although the structure of 2 have been previously reported,37 herein we give a description of its structure for further comparison with compounds 1, 3 and 4. Additionally, a structure very similar to 1 have been described on literature,38 but, despite they present the same crystallographic parameters, compound 1 have one disordered guest dmf molecule in the asymmetric unit in contrast with the absence of solvent molecules on the previously reported structure. Compounds 1-4 were obtained using different conditions than that described on literature for similar systems, however, compounds 1 and 2 presented the same structural topology of those already reported. This point should be highlighted since a correlation between the structural topology and the temperature of the solvothermal process was established in a previous paper by Xiong and co-workers.38 In this work, different types of structural topologies were obtained using the same conditions, only differing on the metal ion. Such approach is highlighted in Scheme 1, with compound 1 presenting a fivefold interpenetrated lvt framework, while the isomorphous compounds 2 and 3 exhibit a sevenfold interpenetrated dia framework, and finally, compound 4, presenting a twofold interpenetrated dmc framework.

ACS Paragon Plus Environment

10

Page 11 of 46 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

Scheme 1. Distinct structural topologies for a) compound 1: fivefold interpenetrated lvt framework, b) compounds 2 and 3: sevenfold interpenetrated dia framework and c) compound 4: twofold interpenetrated dmc framework.

Description of the structures

ACS Paragon Plus Environment

11

Crystal Growth & Design 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 46

{[Cu(pvb)2]·2(dmf)}n (1). Compound 1 crystallizes in the tetragonal system, I41/a space group, with copper(II) cation siting over a special position with occupation of a half, one pvb anion, and one guest dmf molecule in the asymmetric part of unit cell. The dmf molecule present a positional disorder that was refined as two contributions with a 0.55:0.45 occupancy ratio. The metal center is six-coordinated by two carboxylate O atoms from two pvb ligands and two pyridine N atoms from another two pvb ligands, resulting in a non-regular hexacoordinated geometry (N–Cu–O angles are 92.4(4)o, 87.6(10) and 87.3(6)o, 92.7(6)o ) as shown in Figure 1. The Cu–N and the Cu–O bond lengths are 2.014(9) (Cu1-N1A), 2.030(6) Å (Cu1-N1B), and 1.923(4) (Cu1-O1), 2.696(5) Å (Cu1-O2). If we consider the long Cu1-O2 bond length, the carboxylate moieties are coordinated in a bidentate fashion, giving rise to a highly distorted hexacoordinated environment around the copper ions. Considering the base of non-regular octahedron formed by the oxygen atoms from carboxylate group, the angle formed by those oxygens are 53.72(18)o and 126.28(18)o, which is far from idealized 90o angle expected from a regular octahedron. Using the bond valence model, we can have access to information such as oxidation state, bond length and coordination number.39,40 Thus, the sum of bond valences for Cu1-N1 (0.425v.u), Cu1-N1i (0.425v.u) (i = 1-x, 1-y, 2-z), Cu1-O1ii (0.477v.u) (ii = 1/4+y, 5/4-x, 5/4+z), Cu1-O1iii (0.477v.u) (iii = 3/4-y, -1/4+x, 3/4-z), Cu1-O2iii (0.059v.u) and Cu1-O2ii (0.059v.u) is in good agreement for the expected oxidation state value of 2 for the copper ions in 1. However, due to this highly distorted

ACS Paragon Plus Environment

12

Page 13 of 46 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

octahedral environment, the copper(II) coordination sphere could be treated as a pseudo-square-planar geometry, since this geometry is more related to the formation of an extended structure with a 3D lvt topology.

Figure 1. Perspective drawing of the coordination environment of 1. Ellipsoids are at the 50% probability level. Hydrogen atoms and disordered dmf and pvb atoms were omitted for the sake of clarity.

The 3D lvt topological network produced by the planar 4-connecting metal node extends to form a five-fold interpenetrating 3D structure, as shown in Figure 2. When considering just one lvt net along the a axis, one side length is about 30.5 Å while the

ACS Paragon Plus Environment

13

Crystal Growth & Design 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 46

other is about 23.7 Å. The diagonal distances are 38.5 and 33.7 Å. However, after considering the five-fold interpenetration, the voids along a axis reduces notably, as can be seen in Figure 2 (c and d). One can see the same trend for the 3D lvt net along the direction shown in Figure 2 (a and b), which presents a star-like shape with an internal honeycomb-like channel, the latter also presenting a void reduction after considering the five-fold interpenetration. In the other hand, along the c axis, one lvt net presents a side length of about 15.3 Å, with diagonal distances of 33.7 and 19.2 Å. Despite the void space reduction along a axis when considering the five-fold interpenetration, as stated above, there is no influence of the interpenetration on the void space along the c axis, as can be seen in Figure 2 (e and f).

ACS Paragon Plus Environment

14

Page 15 of 46 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

Figure 2. Lattice topology of 1, highlighting the views showing: a) the void spacing in one net along an arbitrary direction; b) the reduce of void spacing when considering the fivefold interpenetration along an arbitrary direction; c) the void spacing for one net

ACS Paragon Plus Environment

15

Crystal Growth & Design 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 46

along the a axis; d) the reduce of void spacing when considering the fivefold interpenetration along the a axis e) the void spacing for one net along the c axis; f) the maintenance of void spacing when considering the fivefold interpenetration along the c axis.

This arrangement produces large rhombus shaped channels along the c axis, with four dmf molecules per channel. Moreover, the dmf molecules are distributed in two different supramolecular arrays in an alternating mode between neighbor channels, with one group with non-conventional hydrogen bonds between the oxygen atoms of the dmf molecules and the hydrogen atoms from the vinyl group of the pvb moiety (O3A…C7B of 3.039(17) Å and 149.8º). Another group is also present, with nonconventional hydrogen bonds between the oxygen atoms of the dmf molecules and the hydrogen atoms from the pyridyl group of the pvb moiety (O3B…C4B of 3.357(12) and 155.2º), as shown in figure S1. The dmf molecules also present non-conventional hydrogen bonds between molecules of each group, being defined by the bond between the methylene hydrogen of one dmf molecule and the amide oxygen of the neighbor dmf molecule, i.e. O3A…C16Aiv (3.46(2) Å (iv = x, y, 1+z), 149.5 º) and O3A…C17Aiv (3.522(4) Å, 111.11º) for one group and, C16b…O3bv (3.73(6) Å (v = -y+5/4, x-1/4, z-1/4), 122.0º) and C16B…O3bvi (3.65(4) Å (vi = -y+5/4, x-1/4, z+3/4), 114.5º) for the other group (figures S2 and S3). Although the non-conventional hydrogen bonding array leads to

ACS Paragon Plus Environment

16

Page 17 of 46 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

very weak intermolecular forces, they extend through the dmf molecules of each channel, which then forms a supramolecular chain along the c axis in a cooperative way, that may contribute for the assistance in the formation of the observed channel structure of 1, governing factors in directing the packing of the interpenetrated networks and in the modulation of their size and shape. Thus, dmf molecule may play an important role as a solvent template for the formation of such large channels that extend in one direction. {[Ni(pvb)2]·(dmf)}n (3). Compound 3 crystallizes in the monoclinic crystal system, Cc space group, with one nickel(II) cation, two pvb anions and one guest dmf molecule as asymmetric unit. The metal center is hexa-coordinated by two pyridine N atoms from two pvb ligands and four carboxylate O atoms from another two pvb ligands, resulting in a highly distorted octahedral geometry as shown in figure 3. The Ni–N bond lengths are 1.946(3) and 1.956(3) Å, while the Ni–O bond lengths from the two carboxylate coordinated in a bidentate fashion are 1.992(3), 2.005(2), 2.010(2), and 2.036(3) Å.

ACS Paragon Plus Environment

17

Crystal Growth & Design 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 46

Figure 3. Perspective drawing of the coordination environment of 3. Ellipsoids are at the 50% probability level. Hydrogen atoms were omitted for the sake of clarity. This environment gives rise to a tetrahedral 4-connecting metal node, producing a 3D dia topological network with distorted adamantanoid cages as building units, which in turn extends to form a seven-fold interpenetrating 3D structure, as shown in figure 4. The side length of each adamantanoid cage is about 15.1 Å. Following the same trend of compound 1, the interpenetration leads to a reduction of the void spaces that is more pronounced at specific directions, as can be seen in figure 4a and 4b and figure 4c and 4d along the b axis. In this case, the void spaces are smaller than those in compound 1,

ACS Paragon Plus Environment

18

Page 19 of 46 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

with triangular shaped channels along the c axis, containing one dmf molecule per channel.

Figure 4. Lattice topology of 3, highlighting the views showing: a) the void spacing in one net along an arbitrary direction; b) the reduce of void spacing when considering the

ACS Paragon Plus Environment

19

Crystal Growth & Design 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 46

sevenfold interpenetration along an arbitrary direction; c) the void spacing for one net along the b axis; d) the b axis showing the reduce of void spacing when considering the sevenfold interpenetration e) the void spacing for one net along the c axis; f) the maintenance of void spacing when considering the sevenfold interpenetration along the c axis.

The influence of the non-conventional hydrogen bonds between the oxygen atom from a dmf molecule and a hydrogen atom from the carboxylate aromatic ring and the hydrogen from vinyl group of the pvb moiety (shortest interactions C9…O5i 3.298(4)Å, 164.6 º and C6…O6i

3.430(5) Å, 176.0º) (i = x+1, -y+1, z+1/2) as a template in this case

seems to favor a smaller channel structure when compared to 1, since there are less dmf molecules per channel than in 1. Thus, the interpenetration must be favored in this case since there is no hydrogen bonding between dmf molecules in a cooperative way, as reflected by the increase of the degree of interpenetration from five to seven.

{[Mn3(pvb)6]n·2dmf} (4). Compound 4 crystallizes in the monoclinic system, P21/c space group, with one and a half Mn(II) cations, three pvb anions, and one guest dmf molecule as asymmetric unit. The building unit is quite different from the complexes discussed above since it consists of a linear trinuclear manganese(II) bridged through the carboxylate moiety of six pvb anions, with four of them in a 2-1:1 mode and two

ACS Paragon Plus Environment

20

Page 21 of 46 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

in a 2-1:2 mode. On the peripheral Mn(II), there are four pvb ligands, two on each side, coordinated by the pyridinic N atom, forming a highly distorted octahedral geometry (Mn1–O bond lengths of 2.2730(14) Å, 2.2263(15) Å, 2.0843(15) Å, 2.1416(14) Å and Mn1–N bond lengths of 2.2064(17) and 2.3113(18) Å). The central Mn(II) ion is six coordinated by carboxylic O atoms, forming an octahedral geometry (Mn2–O bond lengths of 2.2236(14), 2.2236(14), 2.2007(14), 2.2007(14), 2.1226(14) and 2.1226(14) Å, as shown in Figure 5.

ACS Paragon Plus Environment

21

Crystal Growth & Design 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 46

ACS Paragon Plus Environment

22

Page 23 of 46 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

Figure 5. Perspective drawing of the coordination environment of 4, showing a) the Mn(II) ion surroundings and b) the Mn(II) coordination environment with atom labels for the non-hydrogen and non-carbon atoms for best visualization. Ellipsoids are at the 50% probability level. Hydrogen atoms were omitted for the sake of clarity. Symmetry codes: (i) -1+x, 1/2-y, -1/2+z; (ii) x, y, -1+z; (iii) -x, 1-y, -z; (iv) 1-x, 1/2+y, 1/2-z; (v) -x, 1-y, 1z.

The Mn(II) trinuclear unit can be regarded as a 6-connecting node (two of them being double stranded), producing a 3D dmc topological framework, which in turn extends to form a two-fold interpenetrating 3D structure, as shown in Figure 6. The interpenetration in this case also leads to a reduction of the void space that is more pronounced at specific directions, as can be seen in figure 6. In this case, the void spaces are the smallest in comparison to those in compounds 1-3 since there is a clearly reduction of void spacing when considering the interpenetration for all directions, including that along the c axis.

ACS Paragon Plus Environment

23

Crystal Growth & Design 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 46

Figure 6. Lattice topology of 4, highlighting the views showing: a) the void spacing in one net along an arbitrary direction; b) the reduce of void spacing when considering the twofold interpenetration along an arbitrary direction; c) the void spacing for one net along the a axis; d) the reduce of void spacing when considering the twofold

ACS Paragon Plus Environment

24

Page 25 of 46 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

interpenetration along the a axis e) the void spacing for one net along the c axis; f) the reduce of void spacing when considering the twofold interpenetration along the c axis.

As shown for 3, the tridimensional arrangement of the topological lattices in 4 produces smaller channels along the c axis when compared to 1, and also with different shape. The same trend observed in 3 regarding the influence of the non-conventional hydrogen bonds as a template is observed in 4. The bond between the oxygen atom from a dmf molecule and hydrogen atoms from the aromatic ring from carboxylate group and vinyl group of the pvb moiety (shortest interactions C32…O7vi 3.519(3), 169.4º and C37…O7vi 3.271(3), 165.8 º, vi = -1+x, y, z) seems to favor a smaller channel structure since there are less dmf molecules per channel than in 1. However, the degree of interpenetration is lower than 1-3, probably due to the different nature of the nodes since in this case they are formed by the more rigid trinuclear structure bridged by carboxylate moieties. Besides, one of the pyridine group of the pvb ligand is not coordinated.

Magnetic properties The magnetic susceptibility of crushed polycrystalline samples of 1-4 were investigated in the temperature range of 2-300 K under an applied dc magnetic field of 1

ACS Paragon Plus Environment

25

Crystal Growth & Design 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 26 of 46

kOe. The corresponding MT versus T curve, M being the magnetic susceptibility per copper(II), cobalt(II) or nickel(II) unit or per trimanganese(II) unit, are shown in Figures 7-10, respectively. In those 3D structures, the separation distances between two adjacent M(II) (M = Cu, Ni, Co, Mn) across the pvb ligand are 15.2941(1) Å for 1, 15.383(3) Å for 2, and 14.582(3) Å for 3. For compound 4, since the building unit is quite different from compounds 1-3, the main separation distances that should be considered for the study of magnetic properties are that between two adjacent intramolecular Mn(II) in the trinuclear system, of 3.5292(6) Å, those between two peripheral intratrinuclear Mn(II), Mn1...Mn1iii (iii = -x, 1-y, -z) of 7.0584(13) Å, and those between two adjacent intertrinuclear Mn(II), across the pvb ligand, of Mn1…Mn2vii (vii = 1-x, -1/2+y, 1/2-z) 14.055(3), Mn1…Mn1viii (viii = 1+x, 1/2-y, 1/2+z) 15.575(4), Mn1…Mn1ii (ii = x, y, -1+z) 15.904(3), of Mn1…Mn2ii 15.469(3) Å (Figure S4). Those structural parameters indicate weak magnetic coupling exchanged by the pvb bridges for compounds 1-3 and a significant magnetic coupling exchanged by the carboxylate bridges in 2-1:1 and 2-

1:2 modes for intratrinuclear interactions for compound 4. Regarding that, the magnetic properties of compounds 1-3 may be analyzed by a simple Curie-Weiss model, while compound 4 may be analyzed by a linear trinuclear model (vide infra). At room temperature, the MT value for 1 is 0.41 cm3 mol-1 K and remains constant upon cooling until approximately 25 K, which is consistent to a magnetically isolated spin doublet (0.41 cm3 mol-1 K for SCu = 1/2) with g = 2.10. At lower temperatures MT

ACS Paragon Plus Environment

26

Page 27 of 46 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

further decreases to 0.39 cm3 mol−1 K at 2.0 K. This behavior is typical of a very weak antiferromagnetic interaction between the copper(II) ions in the three-dimensional network.

Figure 7. Temperature dependence of MT for 1 collected under an applied magnetic field of 1 kOe. Solid line represents the theoretical fitting (see text). As pointed above, this system may be analyzed through a simple Curie-Weiss law expression (Equation 1),41 since the molecular structure of 1 show very long Cu–Cu distances across the pvb bridge, with a θ parameter to account for the very weak magnetic interactions between the metal ions.

ACS Paragon Plus Environment

27

Crystal Growth & Design 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

𝑁 𝑔2 𝛽2

𝜒 = 3𝑘(𝑇 ― 𝜃) 𝑆(𝑆 + 1)

Page 28 of 46

(1)

where S is the spin operator for each S = 1/2 Cu(II) center, g is the average Landé factor of the copper(II) ions and N, β and k have their usual meanings. For this model, leastsquares best-fit parameters are  = -0.09 K, g = 2.10, and R = 5.3 × 10−6 (R being the agreement factor defined as ∑[(χMT)obs − (χMT)calc]2/∑[(χMT)obs]2). The theoretical curve (solid line in Figure 7) have fitted properly the experimental data, with the very small θ value reflecting the poor mediation of magnetic interactions induced by the extended pvb bridge. For compound 2, the MT value at room temperature is 2.87 cm3 mol-1 K and remains constant upon cooling until approximately 100 K, which is greater than that expected for a magnetically isolated spin quartet (1.875 cm3 mol-1 K for SCo = 3/2) with g = 2.00, indicating that a significant orbital contribution may be involved. At lower temperatures MT further decreases continuously to 1.76 cm3 mol−1 K at 2.0 K. This behavior is compatible to a system with S ≥ 1 spin ground state, which can be subject to a zero-field splitting (ZFS) that arises through coupling of the ground state with excited states via spin–orbit coupling.

ACS Paragon Plus Environment

28

Page 29 of 46 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

Figure 8. Temperature dependence of MT for 2 collected under an applied magnetic field of 1 kOe. Solid line represents the theoretical fitting (see text).

Considering the long Co–Co distances across the pvb bridge in 2, this system was analyzed through a simple Curie-Weiss law expression (Equation 1). However the theoretical curve (solid line in Figure S10) did not fit properly the experimental data (g = 2.49, θ = -5.3 K, and the agreement factor R = 1.0 10-3). Thus, taking into account that the main contribution to the decrease of the MT values at lower temperatures is due to ZFS effects, the following expressions for the magnetic susceptibility for S = 3/2 are applicable (Equations 2 and 3).41

ACS Paragon Plus Environment

29

Crystal Growth & Design 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

𝑁 𝑔2 𝛽2 1 + 9 𝑒 ―2𝐷/𝑘𝑇 𝑘𝑇 4(1 + 𝑒 ―2𝐷/𝑘𝑇)

(2)

𝑁 𝑔2 𝛽2 4 + (3𝑘𝑇/𝐷) (1 ― 𝑒 ―2𝐷/𝑘𝑇) 𝑘𝑇 4(1 + 𝑒 ―2𝐷/𝑘𝑇)

(3)

𝜒∥ = 𝜒⊥ =

Page 30 of 46

where g is the average Landé factor of the cobalt(II) ions, and D is the magnitude of the ZFS. The average molar magnetic susceptibility for a powder sample is given by Equation 4.

𝜒′ =

𝜒 ∥ + 2𝜒 ⊥ 3

(4)

For this model, least-squares best-fit parameters are D = 37.2 cm-1, g = 2.48, and the agreement factor R = 5.0 × 10−5. The theoretical curve (solid line in Figure 8) have fitted properly the experimental data, with D value lower than the typical average values found in literature for octahedral cobalt(II) complexes, in the range of 56.4 to 103.10 cm1.42–46

Despite the lower D value for 2 in comparison with the typical ones, values in the

range of approximately 25 to 50 cm-1 are also reported.47,48 This behavior could be attributed to the very distorted octahedral geometry of the cobalt(II) environment.42 Compound 3 presents a behavior similar to that of compound 2, with a MT value at room temperature of 1.26 cm3 mol-1 K that remains constant upon cooling until approximately 25 K, which is greater than that expected for a magnetically isolated nickel(II) center (1.0 cm3 mol-1 K for SNi = 1) with g = 2.00, indicating that an orbital contribution may be involved as well as observed in compound 2. At lower temperatures MT further decreases continuously to 0.60 cm3 mol−1 K at 2.0 K. This

ACS Paragon Plus Environment

30

Page 31 of 46 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

behavior is compatible to a system with S ≥ 1 spin ground state, which can be subject to a zero-field splitting (ZFS).

Figure 9. Temperature dependence of MT for 3 collected under an applied magnetic field of 1 kOe. Solid line represents the theoretical fitting (see text). Following the same steps for 2, when considering the long Ni–Ni distances across the pvb bridge in 3, this system was also analyzed through a simple Curie-Weiss law expression (Equation 1). However the theoretical curve (solid line in Figure S11) did not fit properly the experimental data (g = 2.32, θ = -1.6 K, and the agreement factor R = 1.6 10-3). As observed for compound 2, the main contribution to the decrease of the MT

ACS Paragon Plus Environment

31

Crystal Growth & Design 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 32 of 46

values at lower temperatures may be attributed to ZFS effects. Thus, the following expressions for the magnetic susceptibility for S = 1 are applicable (Equations 5 and 6).41

𝜒∥ = 𝜒⊥ =

𝑁 𝑔2 𝛽2 2 𝑒 ―𝐷/𝑘𝑇 𝑘𝑇 1 + 2𝑒 ―𝐷/𝑘𝑇

𝑁 𝑔2 𝛽2 (2𝑘𝑇/𝐷) (1 ― 𝑒 ―𝐷/𝑘𝑇) 𝑘𝑇 1 + 2𝑒 ―𝐷/𝑘𝑇

(5) (6)

where g is the average Landé factor of the nickel(II) ions, and D is the magnitude of the ZFS. The average molar magnetic susceptibility for a powder sample is given by Equation 4. For this model, least-squares best-fit parameters are D = 6.4 cm-1, g = 2.28, and the agreement factor R = 1.8 × 10−4. The theoretical curve (solid line in Figure 9) have fitted properly the experimental data, with D value in the range of the average values found in literature for octahedral nickel(II) complexes (from -22 to +12 cm-1).48–52 For compound 4, the MT value at room temperature of 12.72 cm3 mol-1 K is smaller than that expected for three magnetically isolated manganese(II) centers (13.125 cm3 mol-1 K for SMn = 5/2) with g = 2.00. In addition, upon cooling, the MT values gradually decrease to 11.61 cm3 mol-1 K at approximately 100 K, then rapidly decrease to 4.22 cm3 mol−1 K at 2.0 K. This behavior indicates a dominant antiferromagnetic coupling associated to a total spin ground state S = 5/2 at low temperatures (spin-only value of 4.125 cm3 mol-1 K).

ACS Paragon Plus Environment

32

Page 33 of 46 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

Figure 10. Temperature dependence of MT for 4 collected under an applied magnetic field of 1 kOe. Solid line represents the theoretical fitting (see text). The magnetic properties of 4 may be analyzed through the isotropic Heisenberg Spin Hamiltonian for a linear trinuclear cluster (Equation 7).

𝐻 = ― 𝐽1(𝑆𝑀𝑛2 ∙ 𝑆𝑀𝑛1 + 𝑆𝑀𝑛2 ∙ 𝑆𝑀𝑛1𝑖𝑖𝑖) ― 𝐽2(𝑆𝑀𝑛1 ∙ 𝑆𝑀𝑛1𝑖𝑖𝑖)

(7)

where J1 denotes the exchange coupling parameter between the central (Mn2) and peripheral manganese(II) ions (Mn1 and Mn1iii, where Mn1iii is the corresponding peripheral atom generated by symmetry; symmetry code: -x, 1-y, -z), and J2 is the coupling parameter between the two peripheral manganese(II) ions. Assuming that there is no magnetic interaction between the two peripheral Mn(II) ions, since this is a

ACS Paragon Plus Environment

33

Crystal Growth & Design 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 34 of 46

known behavior for most of the known linear Mn(II)3 complexes, and considering the large Mn1–Mn1iii separation of 7.0585(4) Å, J2 was assumed to be negligible. Thus, the expression of the magnetic susceptibility derived through this Hamiltonian by the application of the van Vleck equation41,53 to the Kambe’s vector coupling scheme,54 is given by equation 8.

𝜒=

𝑁 𝑔2 𝛽2 𝐴 𝑘𝑇 𝐵

(8)

where: A = 52.5 + 1020 x27.5 + 682.5 x20 + 682.5 x25 + 429 x13.5 + 429 x18.5 + 429 x22.5 + 247.5 x8 + 247.5 x13 + 247.5 x17 + 247.5 x20 + 126 x3.5 + 126 x8.5 + 126 x12.5 + 126 x15.5 + 126 x17.5 + 52.5 x5 + 52.5 x9 + 52.5 x12 + 52.5 x14 + 52.5 x15 + 15 x2.5 + 15 x6.5 + 15 x9.5 + 15 x11.5 + 1.5 x5 + 1.5 x8; B = 6 + 16 x27.5 + 14 x20 + 14 x25 + 12 x13.5 + 12 x18.5 +12 x22.5 + 10 x8 + 10 x13 + 10 x17 + 10 x20 + 8 x3.5 + 8 x8.5 + 8 x12.5 + 8 x15.5 + 8 x17.5 + 6 x5 + 6 x9 +6 x12 + 6 x14 + 6 x15 + 4 x2.5 + 4 x6.5 + 4 x9.5 +4 x11.5 + 2 x5 + 2 x8; x = exp (-J/kT).

For this model, least-squares best-fit parameters are J = -2.5 cm-1, g = 2.01 and R = 3.0 × 10−5. The theoretical curve (solid line in Figure 10) have fitted properly the experimental data, with J value similar to the average values found in literature (J values in the range of -0.95 to -5.6 cm-1) for linear trimanganese(II) bridged by carboxylate units.55–64 Magnetization measurements as a function of the applied magnetic field were also performed for 1-4 at 2.0 K (Figure S13). The MS values of 1.0 (1), 2.1 (2), 1.7 (3), and 5.2

ACS Paragon Plus Environment

34

Page 35 of 46 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

N (4) at 5.0 T are in agreement with the expected values according to the points discussed above.

Thermal stabilities TGA experiments were conducted to investigate the thermal stabilities of 1-4. As shown in Figure 11, two main steps of weight losses were observed for 1 and 2 on the temperature range 25-500 ºC, while for compounds 3 and 4 more steps where observed in the same temperature range. The first step weight loss of 1 starts at 31.2 ºC, extending until 154.1 ºC, corresponding to the losses of three dmf and two water molecules (Calcd: 33.3%, observed 33.7% for 1). The molecular formula for 1 was defined as 1·dmf·2H2O, since according to SCXRD two dmf molecules are present on the crystal lattice. For 2, the first step weight loss starts at 107.7 ºC, extending until 185.1 ºC, corresponding to the loss of one dmf molecule (Calcd: 12.6%, observed 13.1% for 2). For 3, the first step weight loss starts at 55.0 ºC, extending until 204.8 ºC, corresponding to the losses of one dmf molecule and half water molecule. (Calcd: 13.9%, observed 14.3% for 3). The molecular formula for 3 was defined as 3·0.5H2O, since according to SCXRD one dmf molecule is present on the crystal lattice. Finally, for 4, the first step weight loss starts at 24.7 ºC, extending until 239.2 ºC, corresponding to the losses of two dmf and one water molecules (Calcd: 9.8%, observed 9.7% for 4). The molecular formula for 4 was defined as 4·H2O, since according to SCXRD two dmf molecules are present on the crystal

ACS Paragon Plus Environment

35

Crystal Growth & Design 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 36 of 46

lattice. Further weight losses steps for 1-4 correspond to the decomposition of the framework. The DTG curves for more detailed information on the thermal analysis are presented in the Supporting Information (Figures S5-S8).

Figure 11. TGA curves of 1-4. PXRD measurements were performed on the samples of 1-4 which exhibit diffraction patterns in agreement with those calculated from the single crystal diffraction data, as shown in Figure S9.

CONCLUSIONS

ACS Paragon Plus Environment

36

Page 37 of 46 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

A family of four three-dimensional M(II)-MOFs based on Hpvb ligand with different topologies, degrees of interpenetration, pore sizes and shapes of general formula {[Mx(pvb)2x]·y(dmf)}n were solvothermally obtained. Using the same synthesis conditions, only changing the metal ion enforces subtle changes in the coordination modes, leading to pseudo-square planar (copper compound, 1) octahedral or highly distorted octahedral geometries (cobalt, nickel and manganese compounds, 2, 3 and 4, respectively). This feature, along with template effects induced by the solvent, play an important role in the formation of distinctive structural topologies, as could be seen for compounds 1-4 since 1 exhibits a fivefold interpenetrated lvt framework, while compounds 2 and 3 exhibit a sevenfold interpenetrating dia framework, and finally, compound 4 with a twofold interpenetrated dmc framework. Moreover, the magnetic properties of 1-4 have been investigated and discussed in detail, along with the corresponding , J and D parameters related to their structural characteristics. For compounds 1-3 the very long metal-metal distances imposed by the pvb bridge results in very weak antiferromagnetic interaction. For this reason, these systems were analyzed by a simple Curie-Weiss model in the case of 1, or a ZFS model in the case of 2 and 3. The coupling between the metal centers in 4 is stronger in comparison to compounds 1-3 since the coupling pathway is done by carboxylate bridges instead of the very long pvb pathway, affording short metal-metal separation. Finally, the procedures described here might be generally applicable for the obtainment of variant

ACS Paragon Plus Environment

37

Crystal Growth & Design 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 38 of 46

topological structures by the use of different metal ions in the aspect of constructing other metal-organic complexes.

ASSOCIATED CONTENT Supporting Information. The Supporting Information is available free of charge on the ACS Publications website at DOI: Additional figures for the related compounds, bond lengths and angles, TGA and DrTGA curves, PXRD patterns, magnetic data are provided. CCDC numbers for the structures determined in this work are 1895216, 1895217, 1895218, and 1895219 (the data can be obtained free of charge from The Cambridge Crystallographic Data Centre via www.ccdc.cam.ac.uk/getstructures). AUTHOR INFORMATION Corresponding Author *email: [email protected]. Tel: +55 19 3521-3490 ORCID: Wdeson P. Barros: 0000-0002-7261-1308 Marcos A. Ribeiro: 0000-0002-9350-6419

ACS Paragon Plus Environment

38

Page 39 of 46 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

Kleber R. Pirota: 0000-0002-1467-4415

Author Contributions The manuscript was written through contributions of all authors. All authors have given approval to the final version of the manuscript. Funding Sources This research was financed by Fundação de Amparo à Pesquisa do Estado de São Paulo (FAPESP) (Grant 2015/22379-7) and in part by the Coordenação de Aperfeiçoamento de Pessoal de Nível Superior - Brasil - CAPES (Finance Code 001). Notes The authors declare no competing financial interest. ACKNOWLEDGMENT The authors are deeply grateful to the Fundação de Amparo à Pesquisa do Estado de São Paulo (FAPESP) for financial support (Grant 2015/22379-7) Coordenação de Aperfeiçoamento de Pessoal de Nível Superior (CAPES) is acknowledged for the fellowship to YD. We also thank the Brazilian Synchrotron Light Laboratory (LNLS) for the MX2 beamline time and Dr. Ana Zeri and Dr. Andrey F. Z. Nascimento for the assistance in X-ray diffraction experiments.

ACS Paragon Plus Environment

39

Crystal Growth & Design 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 40 of 46

ABBREVIATIONS Hpvb, Bis {4-[2-(4-pyridyl)ethenyl]} benzoic acid pre-ligand; MOF, metal-organic framework; TGA, thermogravimetric analysis; DrTGA, derivative thermogravimetric analysis; SCXRD, single crystal X-ray diffraction; PXRD, powder X-ray diffraction. REFERENCES (1)

(2)

(3)

(4)

(5)

(6)

(7)

(8)

(9)

Lu, W.; Wei, Z.; Gu, Z.-Y.; Liu, T.-F.; Park, J.; Park, J.; Tian, J.; Zhang, M.; Zhang, Q.; Gentle III, T.; et al. Tuning the Structure and Function of Metal–Organic Frameworks via Linker Design. Chem Soc Rev 2014, 43 (16), 5561–5593. https://doi.org/10.1039/C4CS00003J. Schneemann, A.; Bon, V.; Schwedler, I.; Senkovska, I.; Kaskel, S.; Fischer, R. A. Flexible Metal–Organic Frameworks. Chem Soc Rev 2014, 43 (16), 6062–6096. https://doi.org/10.1039/C4CS00101J. Cadiau, A.; Adil, K.; Bhatt, P. M.; Belmabkhout, Y.; Eddaoudi, M. A MetalOrganic Framework-Based Splitter for Separating Propylene from Propane. Science 2016, 353 (6295), 137–140. https://doi.org/10.1126/science.aaf6323. Rigamonti, L.; Cotton, C.; Nava, A.; Lang, H.; Rüffer, T.; Perfetti, M.; Sorace, L.; Barra, A.-L.; Lan, Y.; Wernsdorfer, W.; et al. Diamondoid Structure in a MetalOrganic Framework of Fe 4 Single-Molecule Magnets. Chem. - Eur. J. 2016, 22 (38), 13705–13714. https://doi.org/10.1002/chem.201601383. Sheberla, D.; Bachman, J. C.; Elias, J. S.; Sun, C.-J.; Shao-Horn, Y.; Dincă, M. Conductive MOF Electrodes for Stable Supercapacitors with High Areal Capacitance. Nat. Mater. 2017, 16 (2), 220–224. https://doi.org/10.1038/nmat4766. Chui, S. S. A Chemically Functionalizable Nanoporous Material [Cu3(TMA)2(H2O)3]n. Science 1999, 283 (5405), 1148–1150. https://doi.org/10.1126/science.283.5405.1148. Li, H.; Eddaoudi, M.; O’Keeffe, M.; Yaghi, O. M. Design and Synthesis of an Exceptionally Stable and Highly Porous Metal-Organic Framework. Nature 1999, 402 (6759), 276–279. https://doi.org/10.1038/46248. Lee, J.; Farha, O. K.; Roberts, J.; Scheidt, K. A.; Nguyen, S. T.; Hupp, J. T. Metal– Organic Framework Materials as Catalysts. Chem. Soc. Rev. 2009, 38 (5), 1450. https://doi.org/10.1039/b807080f. Ohmann, R.; Vitali, L.; Kern, K. Actuated Transitory Metal−Ligand Bond As Tunable Electromechanical Switch. Nano Lett. 2010, 10 (8), 2995–3000. https://doi.org/10.1021/nl1014348.

ACS Paragon Plus Environment

40

Page 41 of 46 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

(10) Agarwal, R. A.; Aijaz, A.; Sañudo, C.; Xu, Q.; Bharadwaj, P. K. Gas Adsorption and Magnetic Properties in Isostructural Ni(II), Mn(II), and Co(II) Coordination Polymers. Cryst. Growth Des. 2013, 13 (3), 1238–1245. https://doi.org/10.1021/cg3016904. (11) Liu, X.; Cen, P.; Li, H.; Ke, H.; Zhang, S.; Wei, Q.; Xie, G.; Chen, S.; Gao, S. SolventInduced Syntheses, Crystal Structures, Magnetic Properties, and Single-Crystalto-Single-Crystal Transformation of Azido-Cu(II) Coordination Polymers with 2Naphthoic Acid as Co-Ligand. Inorg. Chem. 2014, 53 (15), 8088–8097. https://doi.org/10.1021/ic5010769. (12) Giacobbe, C.; Lavigna, E.; Maspero, A.; Galli, S. Elucidating the CO 2 Adsorption Mechanisms in the Triangular Channels of the Bis(Pyrazolate) MOF Fe 2 (BPEB) 3 by in Situ Synchrotron X-Ray Diffraction and Molecular Dynamics Simulations. J. Mater. Chem. A 2017, 5 (32), 16964–16975. https://doi.org/10.1039/C7TA04501H. (13) Tian, D.; Chen, Q.; Li, Y.; Zhang, Y.-H.; Chang, Z.; Bu, X.-H. A Mixed Molecular Building Block Strategy for the Design of Nested Polyhedron Metal-Organic Frameworks. Angew. Chem. Int. Ed. 2014, 53 (3), 837–841. https://doi.org/10.1002/anie.201307681. (14) Sen, S.; Nair, N. N.; Yamada, T.; Kitagawa, H.; Bharadwaj, P. K. High Proton Conductivity by a Metal–Organic Framework Incorporating Zn 8 O Clusters with Aligned Imidazolium Groups Decorating the Channels. J. Am. Chem. Soc. 2012, 134 (47), 19432–19437. https://doi.org/10.1021/ja3076378. (15) Hosseini, M. W. Reflexion on Molecular Tectonics. CrystEngComm 2004, 6 (56), 318. https://doi.org/10.1039/b412125m. (16) Hosseini, M. W. Molecular Tectonics: From Simple Tectons to Complex Molecular Networks. Acc. Chem. Res. 2005, 38 (4), 313–323. https://doi.org/10.1021/ar0401799. (17) Wuest, J. D. Engineering Crystals by the Strategy of Molecular Tectonics. Chem. Commun. 2005, No. 47, 5830. https://doi.org/10.1039/b512641j. (18) Eddaoudi, M.; Moler, D. B.; Li, H.; Chen, B.; Reineke, T. M.; O’Keeffe, M.; Yaghi, O. M. Modular Chemistry: Secondary Building Units as a Basis for the Design of Highly Porous and Robust Metal−Organic Carboxylate Frameworks. Acc. Chem. Res. 2001, 34 (4), 319–330. https://doi.org/10.1021/ar000034b. (19) Yaghi, O. M.; O’Keeffe, M.; Ockwig, N. W.; Chae, H. K.; Eddaoudi, M.; Kim, J. Reticular Synthesis and the Design of New Materials. Nature 2003, 423 (6941), 705– 714. https://doi.org/10.1038/nature01650. (20) Ockwig, N. W.; Delgado-Friedrichs, O.; O’Keeffe, M.; Yaghi, O. M. Reticular Chemistry: Occurrence and Taxonomy of Nets and Grammar for the Design of Frameworks. Acc. Chem. Res. 2005, 38 (3), 176–182. https://doi.org/10.1021/ar020022l. (21) Zhang, Y.-B.; Zhou, H.-L.; Lin, R.-B.; Zhang, C.; Lin, J.-B.; Zhang, J.-P.; Chen, X.-M. Geometry Analysis and Systematic Synthesis of Highly Porous Isoreticular

ACS Paragon Plus Environment

41

Crystal Growth & Design 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(22)

(23) (24)

(25) (26) (27)

(28)

(29)

(30)

(31)

(32)

Page 42 of 46

Frameworks with a Unique Topology. Nat. Commun. 2012, 3 (1). https://doi.org/10.1038/ncomms1654. Liu, Y.-L.; Wu, Y.-P.; Li, D.-S.; Dong, W.-W.; Zhou, C.-S. Two Interpenetrating CuII/NiII-Coordinated Polymers Based on an Unsymmetrical Bifunctional N/OTectonic: Syntheses, Structures and Magnetic Properties. J. Solid State Chem. 2015, 223, 38–43. https://doi.org/10.1016/j.jssc.2014.05.022. Amabilino, D. B.; Smith, D. K.; Steed, J. W. Supramolecular Materials. Chem. Soc. Rev. 2017, 46 (9), 2404–2420. https://doi.org/10.1039/C7CS00163K. Cook, T. R.; Zheng, Y.-R.; Stang, P. J. Metal–Organic Frameworks and SelfAssembled Supramolecular Coordination Complexes: Comparing and Contrasting the Design, Synthesis, and Functionality of Metal–Organic Materials. Chem. Rev. 2013, 113 (1), 734–777. https://doi.org/10.1021/cr3002824. Earnshaw, A. Introduction to Magnetochemistry; Academic Press: London; New York, 1968. Bain, G. A.; Berry, J. F. Diamagnetic Corrections and Pascal’s Constants. J. Chem. Educ. 2008, 85 (4), 532–536. https://doi.org/10.1021/ed085p532. Sharma, M. K.; Lama, P.; Bharadwaj, P. K. Reversible Single-Crystal to SingleCrystal Exchange of Guests in a Seven-Fold Interpenetrated Diamondoid Coordination Polymer. Cryst. Growth Des. 2011, 11 (4), 1411–1416. https://doi.org/10.1021/cg101731j. Guimarães, B. G.; Sanfelici, L.; Neuenschwander, R. T.; Rodrigues, F.; Grizolli, W. C.; Raulik, M. A.; Piton, J. R.; Meyer, B. C.; Nascimento, A. S.; Polikarpov, I. The MX2 Macromolecular Crystallography Beamline: A Wiggler X-Ray Source at the LNLS. J. Synchrotron Radiat. 2009, 16 (1), 69–75. https://doi.org/10.1107/S0909049508034870. Winn, M. D.; Ballard, C. C.; Cowtan, K. D.; Dodson, E. J.; Emsley, P.; Evans, P. R.; Keegan, R. M.; Krissinel, E. B.; Leslie, A. G. W.; McCoy, A.; et al. Overview of the CCP 4 Suite and Current Developments. Acta Crystallogr. D Biol. Crystallogr. 2011, 67 (4), 235–242. https://doi.org/10.1107/S0907444910045749. Winter, G. Xia2 : An Expert System for Macromolecular Crystallography Data Reduction. J. Appl. Crystallogr. 2010, 43 (1), 186–190. https://doi.org/10.1107/S0021889809045701. Winter, G.; Waterman, D. G.; Parkhurst, J. M.; Brewster, A. S.; Gildea, R. J.; Gerstel, M.; Fuentes-Montero, L.; Vollmar, M.; Michels-Clark, T.; Young, I. D.; et al. DIALS : Implementation and Evaluation of a New Integration Package. Acta Crystallogr. Sect. Struct. Biol. 2018, 74 (2), 85–97. https://doi.org/10.1107/S2059798317017235. Evans, P. R.; Murshudov, G. N. How Good Are My Data and What Is the Resolution? Acta Crystallogr. D Biol. Crystallogr. 2013, 69 (7), 1204–1214. https://doi.org/10.1107/S0907444913000061.

ACS Paragon Plus Environment

42

Page 43 of 46 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

(33) Evans, P. Scaling and Assessment of Data Quality. Acta Crystallogr. D Biol. Crystallogr. 2006, 62 (1), 72–82. https://doi.org/10.1107/S0907444905036693. (34) 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 (2), 339–341. https://doi.org/10.1107/S0021889808042726. (35) Sheldrick, G. M. SHELXT – Integrated Space-Group and Crystal-Structure Determination. Acta Crystallogr. Sect. Found. Adv. 2015, 71 (1), 3–8. https://doi.org/10.1107/S2053273314026370. (36) Sheldrick, G. M. Crystal Structure Refinement with SHELXL. Acta Crystallogr. Sect. C Struct. Chem. 2015, 71 (1), 3–8. https://doi.org/10.1107/S2053229614024218. (37) Elsaidi, S. K.; Mohamed, M. H.; Wojtas, L.; Chanthapally, A.; Pham, T.; Space, B.; Vittal, J. J.; Zaworotko, M. J. Putting the Squeeze on CH 4 and CO 2 through Control over Interpenetration in Diamondoid Nets. J. Am. Chem. Soc. 2014, 136 (13), 5072–5077. https://doi.org/10.1021/ja500005k. (38) Xiong, G.; Wang, Y.; Zhao, B.; You, L.; Ren, B.; He, Y.; Wang, S.; Sun, Y. Temperature-Tuned Topologies and Interpenetrations of Two 3D Porous Copper(II)-Organic Frameworks and Gas Adsorption Behaviors. Inorganica Chim. Acta 2018, 471, 180–185. https://doi.org/10.1016/j.ica.2017.10.039. (39) Brown, I. D.; Altermatt, D. Bond-Valence Parameters Obtained from a Systematic Analysis of the Inorganic Crystal Structure Database. Acta Crystallogr. B 1985, 41 (4), 244–247. https://doi.org/10.1107/S0108768185002063. (40) Nfor, E. N.; You, X.-Z.; Wei, L.; Iniama, G. (Acetato-κ O )(2-Acetylpyridine 4Methylthiosemicarbazonato-κ 3 N , N ′, S )Copper(II) Monohydrate. Acta Crystallogr. Sect. E Struct. Rep. Online 2006, 62 (2), m379–m381. https://doi.org/10.1107/S1600536806002364. (41) Kahn, O. Molecular Magnetism; VCH: New York, NY, 1993. (42) Ding, Z.-Y.; Meng, Y.-S.; Xiao, Y.; Zhang, Y.-Q.; Zhu, Y.-Y.; Gao, S. Probing the Influence of Molecular Symmetry on the Magnetic Anisotropy of Octahedral Cobalt( II ) Complexes. Inorg. Chem. Front. 2017, 4 (11), 1909–1916. https://doi.org/10.1039/C7QI00547D. (43) Hudák, J.; Boča, R.; Dlháň, Ľ.; Kožíšek, J.; Moncoľ, J. Structure and Magnetism of Mono-, Di-, and Trinuclear Benzoato Cobalt(II) Complexes. Polyhedron 2011, 30 (7), 1367–1373. https://doi.org/10.1016/j.poly.2011.02.047. (44) Świtlicka-Olszewska, A.; Palion-Gazda, J.; Klemens, T.; Machura, B.; Vallejo, J.; Cano, J.; Lloret, F.; Julve, M. Single-Ion Magnet Behaviour in Mononuclear and Two-Dimensional Dicyanamide-Containing Cobalt( II ) Complexes. Dalton Trans. 2016, 45 (25), 10181–10193. https://doi.org/10.1039/C6DT01160H. (45) Titiš, J.; Hudák, J.; Kožíšek, J.; Krutošíková, A.; Moncol’, J.; Tarabová, D.; Boča, R. Structural, Spectral and Magnetic Properties of Carboxylato Cobalt(II) Complexes with Heterocyclic N-Donor Ligands: Reconstruction of Magnetic Parameters from

ACS Paragon Plus Environment

43

Crystal Growth & Design 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(46)

(47)

(48) (49)

(50)

(51)

(52)

(53)

(54) (55)

Page 44 of 46

Electronic Spectra. Inorganica Chim. Acta 2012, 388, 106–113. https://doi.org/10.1016/j.ica.2012.03.036. Vallejo, J.; Castro, I.; Ruiz-García, R.; Cano, J.; Julve, M.; Lloret, F.; De Munno, G.; Wernsdorfer, W.; Pardo, E. Field-Induced Slow Magnetic Relaxation in a SixCoordinate Mononuclear Cobalt(II) Complex with a Positive Anisotropy. J. Am. Chem. Soc. 2012, 134 (38), 15704–15707. https://doi.org/10.1021/ja3075314. Herchel, R.; Váhovská, L.; Potočňák, I.; Trávníček, Z. Slow Magnetic Relaxation in Octahedral Cobalt(II) Field-Induced Single-Ion Magnet with Positive Axial and Large Rhombic Anisotropy. Inorg. Chem. 2014, 53 (12), 5896–5898. https://doi.org/10.1021/ic500916u. Boča, R. Zero-Field Splitting in Metal Complexes. Coord. Chem. Rev. 2004, 248 (9– 10), 757–815. https://doi.org/10.1016/j.ccr.2004.03.001. Mašlejová, A.; Ivaniková, R.; Svoboda, I.; Papánková, B.; Dlháň, L.; Mikloš, D.; Fuess, H.; Boča, R. Structural Characterization and Magnetic Properties of Hexakis(Imidazole)Nickel(II) Bis(Formate), Bis(Chloroacetate), Bis(2Chloropropionate) and Hexakis(1-Methyl-Imidazole)Nickel(II) Chloride Dihydrate. Polyhedron 2006, 25 (8), 1823–1830. https://doi.org/10.1016/j.poly.2005.11.026. Burkhardt, A.; Plass, W. Modeling the Coordination Chemistry of Chitosan: Synthesis and Characterization of a Nickel(II) Complex with a 2-Aminoglucose Schiff-Base Ligand. Inorg. Chem. Commun. 2008, 11 (3), 303–306. https://doi.org/10.1016/j.inoche.2007.12.006. Titiš, J.; Boča, R.; Dlháň, L.; Ďurčeková, T.; Fuess, H.; Ivaniková, R.; Mrázová, V.; Papánková, B.; Svoboda, I. Magnetostructural Correlations in Heteroleptic Nickel(II) Complexes. Polyhedron 2007, 26 (7), 1523–1530. https://doi.org/10.1016/j.poly.2006.11.054. Panja, A.; Jana, N. Ch.; Adak, S.; Brandão, P.; Dlháň, L.; Titiš, J.; Boča, R. The Structure and Magnetism of Mono- and Di-Nuclear Ni( II ) Complexes Derived from {N 3 O}-Donor Schiff Base Ligands. New J. Chem. 2017, 41 (8), 3143–3153. https://doi.org/10.1039/C7NJ00158D. Van Vleck, J. H. The Theory of Electric and Magnetic Susceptibilities.; The International series of monographs on physics; Oxford University Press: New York, 1932. Kambe, K. On the Paramagnetic Susceptibilities of Some Polynuclear Complex Salts. J. Phys. Soc. Jpn. 1950, 5 (1), 48–51. https://doi.org/10.1143/JPSJ.5.48. Menage, S.; Vitols, S. E.; Bergerat, P.; Codjovi, E.; Kahn, O.; Girerd, J. J.; Guillot, M.; Solans, X.; Calvet, T. Structure of the Linear Trinuclear Complex Hexakis(Acetato)Bis(2,2’-Bipyridine)Trimanganese (II). Determination of the J Electron-Exchange Parameter through Magnetic Susceptibility and High-Field

ACS Paragon Plus Environment

44

Page 45 of 46 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

(56)

(57)

(58)

(59)

(60)

(61)

(62)

(63)

(64)

Magnetization Measurements. Inorg. Chem. 1991, 30 (12), 2666–2671. https://doi.org/10.1021/ic00012a020. Rardin, R. L.; Poganiuch, P.; Bino, A.; Goldberg, D. P.; Tolman, W. B.; Liu, S.; Lippard, S. J. Synthesis and Characterization of Novel Trinuclear Iron(II) and Manganese(II) Carboxylate Complexes: Structural Trends in Low Valent Iron and Manganese Carboxylates. J. Am. Chem. Soc. 1992, 114 (13), 5240–5249. https://doi.org/10.1021/ja00039a041. Reynolds, R. A.; Dunham, W. R.; Coucouvanis, D. Kinetic Lability, Structural Diversity, and Oxidation Reactions of New Oligomeric, Anionic Carboxylate−Pyridine Complexes. Inorg. Chem. 1998, 37 (6), 1232–1241. https://doi.org/10.1021/ic971537p. Bu, X.-H.; Tong, M.-L.; Xie, Y.-B.; Li, J.-R.; Chang, H.-C.; Kitagawa, S.; Ribas, J. Synthesis, Structures, and Magnetic Properties of the Copper(II), Cobalt(II), and Manganese(II) Complexes with 9-Acridinecarboxylate and 4Quinolinecarboxylate Ligands. Inorg. Chem. 2005, 44 (26), 9837–9846. https://doi.org/10.1021/ic050886d. Kloskowski, M.; Pursche, D.; Hoffmann, R.-D.; Pöttgen, R.; Läge, M.; Hammerschmidt, A.; Glaser, T.; Krebs, B. Novel Trinuclear MnII/MnII/MnII Complexes – Crystal Structures and Catalytic Properties. Z. Für Anorg. Allg. Chem. 2007, 633 (1), 106–112. https://doi.org/10.1002/zaac.200600262. Escuer, A.; Cordero, B.; Solans, X.; Font-Bardia, M.; Calvet, T. Manganese(II) Trimeric Systems Derived from Pyridyldioximato Ligands: Synthesis, Crystal Structure, and Magnetic Characterization. Eur. J. Inorg. Chem. 2008, 2008 (32), 5082–5087. https://doi.org/10.1002/ejic.200800548. Tan, Q.-H.; Wang, Y.-Q.; Liu, H.-T.; Liu, Z.-L. Two New Carboxylate–Oxygen Bridged Trinuclear M(II) (MMn and Co) Compounds with Zwitterionic Dicarboxylate Ligands: Crystal Structures and Magnetism. Inorg. Chem. Commun. 2015, 58, 67–70. https://doi.org/10.1016/j.inoche.2015.06.008. Escriche-Tur, L.; Font-Bardia, M.; Albela, B.; Corbella, M. Determination of ZFS Parameters from the EPR Spectra of Mono-, Di- and Trinuclear Mn II Complexes: Impact of Magnetic Coupling. Dalton Trans. 2017, 46 (8), 2699–2714. https://doi.org/10.1039/C6DT04012H. Zhou, J.; Wang, X.; Yan, T.; Wang, T.; Wang, D.-W.; Yang, J.-R.; Xie, M.-J.; Zhao, Q.-H. Syntheses, Structures, and Magnetic Properties of Two MOFs Based on VShaped Carboxylate and Flexible Spacer Ligand: From Binuclear Co(II) with α-Po Net to Trinuclear Mn(II) with 416.611.8 Topology. Inorg. Chem. Commun. 2017, 86, 35–38. https://doi.org/10.1016/j.inoche.2017.09.013. Cui, L.; Wang, Q.-H.; Duan, J.-W.; Wang, K.; Wang, Y.-Y. Construction of a Magnetic Chiral Coordination Polymer Based on a Semi-Rigid Carboxylic Acid. J. Coord. Chem. 2018, 71 (3), 421–427. https://doi.org/10.1080/00958972.2018.1444161.

ACS Paragon Plus Environment

45

Crystal Growth & Design 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 46 of 46

FOR TABLE OF CONTENTS ONLY Title: Influence of the metal ion in the topology and interpenetration of pyridylvinyl(benzoate) based metal-organic frameworks Authors: Dezotti, Yuri; Ribeiro, Marcos A.; Pirota, Kleber R.; Barros, Wdeson P.

Synopsis: The influence of different metal ions and solvent template effects on the structural topologies and magnetic properties of a family of four M(II)-MOFs of general formula {[Mx(pvb)2x]·y(dmf)}n (M=Cu, 1; M=Co, 2; M=Ni, 3; M=Mn, 4), based on the bis{4-[2-(4-pyridyl)ethenyl]} benzoic acid (Hpvb) ligand were investigated.

ACS Paragon Plus Environment

46