Coordination Polymers with Tetrafluoroterephthalate as Bridging Ligand

Oct 5, 2011 - ... which are connected by the tfBDC2– ligand to form a three-dimensional (3D) network. Upon heating it decomposes at approximately 20...
0 downloads 0 Views 1MB Size
ARTICLE pubs.acs.org/crystal

Coordination Polymers with Tetrafluoroterephthalate as Bridging Ligand Christiane Seidel, Ruth Ahlers, and Uwe Ruschewitz* Department of Chemistry, University of Cologne, Greinstrasse 6, D-50939 Cologne, Germany

bS Supporting Information ABSTRACT: Seven nonporous coordination polymers with tetrafluoroterephthalate (tfBDC2) as a bridging ligand were synthesized and structurally characterized. Homoleptic 3 ∞ [Tl2(tfBDC)] (P1, Z = 1, 1) contains six-coordinated Tl centers, which are connected by the tfBDC2 ligand to form a three-dimensional (3D) network. Upon heating it decomposes at approximately 200 °C. Nine-coordinated PbO9 polyhedra are 2 [Pb(tfBDC)(H2O)3] 3 1/2 H2O (P1, Z = 1, 2), found in ∞ which forms a layer-like structure. These layers are held together by hydrogen bonds involving water molecules. These water molecules are released by heating to 70100 °C. II II 1 ∞[M (tfBDC)(H2O)4] (P1, Z = 1; M = Zn (3), Co (4), Ni II (5)) contains almost undistorted M O6 octahedra with the carboxylate groups of the tfBDC ligands in trans coordination so that a chain-like structural unit is formed. These chains are further connected by hydrogen bonds. Upon heating, these water molecules are 3 [Mn2(tfBDC)2(DMF)2(EtOH)] (P21, Z = 2, 7) is isostructural to the known Zn released in two steps at 100 and 200 °C. ∞ compound (6). Both compounds contain dimeric [MIIO6]2 units, which are connected by tfBDC2 ligands to form a 3D network. All compounds are unprecedented in the crystal chemistry of coordination polymers with nonfluorinated terephthalate (BDC2) as bridging ligand. This is mainly because the planar conformation found for BDC2 is energetically less favorable for perfluorinated tfBDC2.

’ INTRODUCTION Coordination polymers and especially their porous congeners, which are frequently termed metal-organic frameworks (MOFs),1 are the focus of many research groups worldwide. The interest in this class of compounds is mainly based on their easy synthetic accessibility and potential applications.24 One of these applications, their potential as hydrogen storage materials, is of considerable interest and has led to numerous publications.510 In this context, MOFs containing fluorinated ligands are receiving increasing attention.1115 One of the most frequently used linker ligands for the construction of MOFs is the dianion of terephthalic acid (H2BDC), from which two of the most prominent members of this class of compounds, MOF-51 and MIL-53,16 are constructed. For MOFs with its perfluorinated counterpart, that is, 2,3,5,6-tetrafluoroterephthalic acid (H2tfBDC), theoretical investigations predict superior H2 adsorbing properties.17 By contrast, a different computational study concludes that the interaction between H2 and fluoroaromatic compounds may be even weaker compared with their nonfluorinated analogues.18 The beneficial properties of fluorinated linkers are however supported by experimental results for a MOF based on a triazolate with CF3 substituents, which shows excellent gas storage capacities for O2 and H2.11 Similar results were obtained very recently for a partially fluorinated MOF based on H2tfBDC with a highly interpenetrated framework r 2011 American Chemical Society

structure.19 A slightly enhanced adsorption enthalpy of 6.2 kJ/mol compared with other non-fluorinated MOFs was obtained. But despite these promising results to date, only a few coordination polymers or MOFs employing perfluorinated linking terephthalates (tfBDC2) have been reported.14,1935 Most of them are nonporous. In many cases, it was found that additional non-fluorinated coligands, for example, imidazole, 2,20 -bipyridine, or 4,40 -bipyridine, are needed, which is probably due to the enhanced acidity of H2tfBDC compared to non-fluorinated H2BDC.32 Furthermore, in fluorinated H2tfBDC and its coordination polymers, the carboxylate groups are twisted out of the plane of the benzene ring. The respective torsion angles are considerably enlarged (4560°) compared to coordination polymers with BDC2 ligands, where torsion angles near 0° are observed.32 This was attributed to an electrostatic repulsion between the fluorine atoms on the ring and the oxygen atoms of the carboxylate groups as well as a decrease in aromatic character of the carboxylate group due to the electron-withdrawing nature of the fluorine atoms.32,36 Thus, a porous perfluorinated MOF isostructural to the well-known nonfluorinated congeners is still unknown. Received: July 27, 2011 Revised: September 15, 2011 Published: October 05, 2011 5053

dx.doi.org/10.1021/cg200974h | Cryst. Growth Des. 2011, 11, 5053–5063

Crystal Growth & Design Recently, we have reported an improved synthesis of tetrafluoroterephthalic acid (H2tfBDC) starting from readily available 1,2,4,5-tetrafluorobenzene.37 With gram quantities of this ligand in hand, we have started to synthesize new coordination polymers with tetrafluoroterephthalate as a bridging ligand. Some of these results will be presented in the following.

’ EXPERIMENTAL SECTION General Remarks. Tetrafluoroterephthalic acid (H2tfBDC) and the corresponding ammonium salt were prepared according to the procedures described in the literature.37 TlCH3COO (Merck), Pb(CH3COO)2 3 3H2O, Zn(NO3)2 3 4H2O, Zn(CH3COO)2 3 2H2O, Ni(CH3COO)2 3 4H2O (all from Fisher Scientific), Co(CH3COO)2 3 4 H2O (Riedel de Haen), NaOH (Riedel de Haen), Mn(NO3)2 3 4H2O (Fluka), N,N0 -dimethylformamide (KMF Laborchemie Handels GmbH), and ethanol (Biesterfeld) were used as purchased without any further purification. Triethylamine (Acros Organics) was distilled prior to the experiment. Caution! Thallium compounds are highly toxic. Therefore adequate safety precautions have to be made before handling thallium compounds. Syntheses. 3∞[Tl2(tfBDC)] (1). In a beaker, 1 mL of 1 M aqueous NaOH was added to a solution of 136.1 mg (0.57 mmol) of diammonium tetrafluoroterephthalate37 in 5 mL of deionized water, until a clear solution was obtained. Subsequently, 263.4 mg (1.0 mmol) of Tl(CH3COO) in 2 mL of deionized water were added. The beaker was sealed with a perforated foil. After three weeks, long crystalline needles had been formed, from which single crystals of 1 were isolated. Elemental analysis for C8O4F4Tl2 (644.84): calcd C, 14.90%; found C, 14.96%. Purity was additionally checked by X-ray powder diffraction (XRPD) (see Supporting Information, Figure S1). Some weak additional reflections were detected. 2 ∞[Pb(tfBDC)(H2O)3] 3 1/2H2O (2). In a beaker 27.5 mg (0.072 mmol) of Pb(CH3COO)2 3 3H2O were dissolved in 10 mL of deionized water and added to a solution of 15.9 mg (0.067 mmol) of tetrafluoroterephthalic acid in 4 mL of deionized water. The beaker was sealed with a perforated foil. After four weeks the solvent has been evaporated completely and a large block-shaped crystal of 2 remained. Elemental analysis for C16H14F8O15Pb2 (1012.66): calcd C, 18.98%, H, 1.39%; found C, 19.10%, H, 0.42%. Purity was additionally checked by XRPD (see Supporting Information, Figure S2). Both suggest that no single-phase sample was obtained. 1 ∞[Zn(tfBDC)(H2O)4] (3). In a mechanochemical synthesis, 96.5 mg (0.44 mmol) of Zn(CH3COO)2 3 2H2O and 121.4 mg (0.51 mmol) of H2tfBDC were ground intensively in an agate mortar, until no further evaporation of acetic acid was detected (1020 min). After recrystallization in a mixture of ethanol and water (2:1, v:v), colorless crystals of 3 were obtained. Alternatively, 3 can be synthesized under solvothermal conditions reacting 450.3 mg (1.72 mmol) of Zn(NO3)2 3 4H2O and 133.9 mg (0.56 mmol) of H2tfBDC in 5.3 mL of ethanol. The solution of the starting materials was transferred to a glass tube and heated for 24 h at 90 °C with a heating rate of 5 °C/h and a cooling rate of 2 °C/h. A microcrystalline powder was obtained, which was used to investigate the properties of 3. Elemental analysis for C8O8H8F4Zn (373.51): calcd C, 25.72%, H, 2.16%; found C, 25.38%, H, 1.94%. Purity was additionally checked by XRPD (see Supporting Information, Figure S3). Both investigations confirm that a single-phase sample was obtained. 1 ∞[Co(tfBDC)(H2O)4] (4). 4 was synthesized by grinding 88.8 mg (0.36 mmol) of Co(CH3COO)2 3 4H2O and 118.9 mg (0.5 mmol) of H2tfBDC. Powder diffraction data confirm that the synthesis of pure 4 in a polycrystalline form can be accomplished by this mechanochemical

ARTICLE

route (see Supporting Information, Figure S4). To obtain single crystals for a crystal structure analysis, the resulting powder was recrystallized from a mixture of ethanol and water (2:1, v:v). XRPD data state that a single-phase sample was obtained, but the elemental analysis does not provide a convincing agreement. Elemental analysis for C8O8H8F4Co (367.07): calcd C, 26.18%, H, 2.20%; found C, 29.31%, H, 2.42%. 1 ∞[Ni(tfBDC)(H2O)4] (5). In a beaker, 127.3 mg (0.51 mmol) of Ni(CH3COO)2 3 4H2O and 60.5 mg (0.25 mmol) of H2tfBDC were dissolved in 10 mL of deionized water. The beaker was sealed with a perforated foil. After several weeks, a green crystalline powder of 5 precipitated. Its crystal structure was confirmed by comparison of its XRPD data to those of 3 and 4. The respective lattice parameters are given below (Table 3). These XRPD data confirm the purity of the sample (see Supporting Information, Figure S5). Elemental analysis for C8O8H8F4Ni (366.83): calcd C, 26.19%, H, 2.20%; found C, 27.93%, H, 2.87%. 3 ∞[M2(tfBDC)2(DMF)2(EtOH)] (M = Zn (6), Mn (7)). One leg of an H-shaped tube was filled with a solution of 40.2 mg (0.15 mmol) of Zn(NO3)2 3 4 H2O or 36.3 mg (0.15 mmol) Mn(NO3)2 3 4H2O and 26.2 mg (0.11 mmol) of H2tfBDC in 1 mL of a mixture of ethanol and DMF (3:1, v:v). The other leg of the H-shaped tube was filled with 70 μL of triethylamine (0.5 mmol) dissolved in 5 mL of ethanol and DMF (3:1, v:v). Both tubes were closed. After a few days, single crystals of 6 and 7 suitable for an X-ray single crystal structure precipitated by slow diffusion of the base into the acidsalt solution. Elemental analysis for Zn2C24H20F8N2O11 (795.16): calcd C, 36.25%, H, 2.53%, N, 3.52%; found C, 34.60%, H, 3.39%, N, 4.43%; for Mn2C24H20F8N2O11 (774.30): calcd C, 37.23%, H, 2.60%, N, 3.62%; found C, 36.34%, H, 2.98%, N, 4.37%. Purity was additionally checked by XRPD (see Supporting Information, Figures S6 and S7). X-ray Single Crystal Structure Analysis. Single crystals of 14 and 67 were isolated as described above and mounted in sealed glass capillaries on a Stoe IPDS I or Stoe IPDS II single crystal diffractometer (T ≈ 293 K, MoKα radiation). For data collection and reduction, the Stoe program package38 was applied. The structural models were solved using SIR-9239 and completed using difference Fourier maps calculated with SHELXL-97,40 which was also used for final refinements. All programs were run under the WinGX system.41 All non-hydrogen atoms were refined anisotropically. For each compound, the treatment of hydrogen atoms is described below. Details of all single crystal structure analyses42 are given in Table 1. Selected interatomic distances and angles are listed in Table 2. X-ray Powder Diffraction. XRPD data were collected on a Huber G670 with a germanium monochromator, CuKα1 radiation and an imaging plate detector at room temperature. Cobalt containing compounds were analyzed using MoKα1 radiation, also using a Huber G670 diffractometer. Samples were sealed in capillaries (Ø 0.20.5 mm) or placed between two foils as a flat sample showing two additional reflections of the foil at 2θ ≈ 21.5° and 2θ ≈ 23.7°. Exposure times vary from 20 to 30 min for a measurement of a flat sample to 120800 min for measurements in capillaries. Within the WinXPow software suite,43 the recorded patterns were compared with theoretical patterns calculated from the obtained single crystal structure data. Thermoanalytical Investigations. Differential thermal analyses (DTA) and thermogravimetry (TG) investigations were performed on compounds 13 (sample mass: 19.9 mg (1), 14.3 mg (2), 21.1 mg (3), Al2O3 containers) in the temperature ranges 20400 °C (1, 2) and 20500 °C (3) with the heating rate 10 °C/min in an argon atmosphere. The instrument (Netzsch STA 409C) is housed in a glovebox (M. Braun, Garching/Germany, N2 atmosphere). Infrared Spectroscopy. IR measurements were carried out on solid KBr pellets using a Bruker IFS 66v/S with a Nernst globar. 5054

dx.doi.org/10.1021/cg200974h |Cryst. Growth Des. 2011, 11, 5053–5063

7.353(1) Å 8.839(1) Å 10.447(2) Å

colorless needles

0.22 3 0.18 3 0.17 mm P1 (No. 2), 1

3.718(1) Å

6.760(2) Å 10.202(3) Å

crystal description

crystal size

space group; Z

unit cell

244.7(1) Å3

colorless needles

6 e h e 6, 8 e k e 8,

Stoe IPDS II, MoKα 293(2) K 59.5° 10 e h e 10, 12 e k e 12, 14 e l e 14

Stoe IPDS II, MoKα 293(2) K

59.5°

5 e h e 4,

8 e k e 9,

14 e l e 14

2θmax

index ranges

5055

0.051 1.83/1.35

0.071 1327/83/0

1.05

0.042

0.113

1.17/2.57

R(int) data/parameters/restraints

GOF = Sall

R [F2 > 2σ(F2)]

wR(F2)

ΔFmax/ΔFmin

Flack x

2899 with I > 2σ(I)

1119 with I > 2σ(I)

significant reflections

0.025

1.04

0.066 3335/213/6

11636/3335

2927/1327

reflections collected/independent

0.63/0.77

0.078

0.034

1.05

0.069 1290/114/4

1096 with I > 2σ(I)

3489/1290

12 e l e 12

56.2°

Stoe IPDS I, MoKα 293(2)

numerical

diffractometer, radiation temperature

2.135 g 3 cm

numerical

numerical

4.376 g 3 cm

3

290.6(1) Å3

76.35(2)°

88.35(2)°

73.02(2)°

6.420(1) Å 9.384(2) Å

5.194(1) Å

0.62 3 0.22 3 0.12 mm P1 (No. 2), 1

absorption correction

2.807 g 3 cm

599.0(2) Å3

92.82(3)° 3

68.10(1)°

92.74(2)°

3

76.53(1)° 73.96(1)°

106.78(2)°

0.3 3 0.2 3 0.2 mm P1 (No. 2), 1

373.51 g 3 mol1

ZnC8H8F4O8

3

calc density

volume

1012.65 g 3 mol1

644.82 g 3 mol1 colorless blocks

Pb2C16H14F8O15

Tl2C8F4O4

formula weight

2

formula

1

Table 1. Details of X-ray Single Crystal Structure Determination of Compounds 14, 6, and 7

3

1.54/1.08

0.208

0.076

1.04

0.134 1299/113/4

1065 with I > 2σ(I)

3479/1299

12 e l e 12

8 e k e 8,

6 e h e 6,

56.2°

Stoe IPDS I, MoKα 293(2)

2.109 g 3 cm numerical

289.1(1) Å3

76.90(3)°

88.67(3)°

72.66(3)°

6.409(2) Å 9.396(3) Å

5.169(1) Å

0.44 3 0.13 3 0.03 mm P1 (No. 2), 1

pink plates

367.07 g 3 mol1

CoC8H8F4O8

4

0.96/0.54 0.00(4)

0.007(13)

0.132

0.056

0.90

0.099 7373/439/1

4624 with I > 2σ(I)

0.47/0.56

0.111

0.047

0.93

0.094 7231/438/1

5463 with I > 2σ(I)

14882/7373

12 e l e 12 14213/7231

25 e k e 25, 25 e k e 25,

11 e h e 11,

56.30°

Stoe IPDS I, MoKα 293(2)

1.638 g 3 cm3 numerical

1569.9(3) Å3

90.00°

101.82(1)°

90.00°

19.512(2) Å 9.742(1) Å

8.438(1) Å

0.34 3 0.21 3 0.15 mm P21 (No. 4), 2

brown blocks

774.30 g 3 mol1

Mn2C24H20F8N2O11

7

12 e l e 12

10 e h e 10,

56.30°

Stoe IPDS I, MoKα 293(2)

numerical

1.741 g 3 cm

3

1516.7(3) Å3

90.00°

101.73(2)°

90.00°

19.402(2) Å 9.697(1) Å

8.234(1) Å

0.78 3 0.29 3 0.17 mm P21 (No. 4), 2

colorless needles

795.16 g 3 mol1

Zn2C24H20F8N2O11

6

Crystal Growth & Design ARTICLE

dx.doi.org/10.1021/cg200974h |Cryst. Growth Des. 2011, 11, 5053–5063

2.817(3) 2.906(5)

Pb1O8B

Pb1O3

5056

126.5(11)

115.4(10)

118.0(9)

O1AC1-O1B

O1BC1C2

1.531(14)

C1C2

O1AC1-C2

1.270(14)

1.226(15)

C1O1A

C1O1B

C1O1A

1.237(5)

117.4(4) 124.7(4) 117.3(4) 118.1(4)

O8AC8-O8B

O8AC8-C5 O8BC8C5

117.5(3)

O1BC1C2

O1AC1-C2

125.1(4)

1.529(5) 1.533(6)

C1C2

C5C8

O1AC1-O1B

1.242(6) 1.255(6)

1.245(5)

C8O8A C8O8B

C1O1B

4.6234(8)

4.4842(7)

2.808(4)

Pb1O1B

2.693(3)

2.680(4)

2.628(4)

2.592(3) 2.608(3)

2.778(4)

2

Pb1O1B

Pb1Pb1

3.005(9)

Tl1O1B

Pb1O1A

Pb1Pb1

2.989(10)

Tl1O1A

Pb1O4

Pb1O2

3.882(1)

2.915(10)

Tl1O1A

Pb1O1A Pb1O8A

3.718(1) 2x

2.880(11)

Tl1O1B

Tl1Tl1

2.825(9) 2.855(10)

Tl1O1B Tl1O1A

1

O1BC1C2

O1AC1-C2

O1AC1-O1B

C1C2

C1O1B

C1O1A

Zn1O1A

Zn1O2 Zn1O3

3

117.0(3)

117.0(2)

126.0(3)

1.514(4)

1.251(3)

1.258(4)

2.131(2) 2x

2.079(2) 2x 2.095(2) 2x

O1BC1C2

O1AC1-C2

O1AC1-O1B

C1C2

C1O1B

C1O1A

Co1O1A

Co1O3 Co1O2

Table 2. Selected Interatomic Distances [Å] < 5 Å and Angles [°] for Compounds 14, 6, and 7 4

117.3(4)

115.6(4)

127.1(4)

1.519(6)

1.242(6)

1.259(5)

2.115(3) 2x

2.080(4) 2x 2.103(3) 2x

2.174(4)

Zn2O1

C7O1

C3C8O4 O3C8O4

C3C8O3

O1C7O2

C6C7O2

C6C7O1

C3C8

C6C7

C8O3 C8O4

C7O2

115.3(5) 116.6(4) 128.1(5)

C3C8O4 O3C8O4

C3C8O3

O1C7O2

C6C7O2 123.2(5)

C6C7O1 120.1(5)

C3C8

C6C7

C8O3 C8O4

C7O2

C7O1

Mn1Mn2

Mn2O1

Mn2O7

Mn2O9 Mn2O6

Mn2O4

Mn2O11

Mn1O2

Mn1O1

Mn1O10

Mn1O5

Mn1O3 Mn1O8

116.6(5)

1.512(7)

1.514(7)

1.239(6) 1.236(7)

1.214(7)

1.272(6)

3.4426(9)

2.101(4)

Zn2O6 Zn1Zn2

2.085(4) 2.086(4)

2.063(4)

2.028(6)

2.633(4)

2.098(3)

2.060(3)

2.047(3)

2.007(4) 2.047(4)

Zn2O4 Zn2O7

Zn2O9

Zn2O11

Zn1O2

Zn1O5

Zn1O8

Zn1O1

Zn1O3 Zn1O10

6

7

118.5(6) 126.0(6)

115.4(6)

122.7(6)

120.5(6)

116.8(5)

1.511(8)

1.508(8)

1.244(8) 1.224(8)

1.227(7)

1.249(7)

3.596(1)

2.192(4)

2.163(4)

2.138(5) 2.152(4)

2.134(4)

2.107(7)

2.502(4)

2.203(4)

2.159(5)

2.142(3)

2.110(4) 2.122(3)

Crystal Growth & Design ARTICLE

dx.doi.org/10.1021/cg200974h |Cryst. Growth Des. 2011, 11, 5053–5063

Crystal Growth & Design

’ RESULTS AND DISCUSSION 3 ∞[Tl2(tfBDC)]

(1). 1 crystallizes in the triclinic space group P1 as colorless needles. Elemental analysis states a high purity of the sample, but powder diffraction data (Figure S1, Supporting Information) show some weak additional reflections. The coordination sphere of the thallium cation is shown in Figure 1. Taking bond distances up to 3.005 Å into account (the nextnearest TlO distance is at 3.679 Å), an unsymmetric polyhedron with coordination number 6 results. One part of the sphere remains uncoordinated, which is typical for a stereochemically active lone pair. The TlO6 polyhedra are located in layers parallel to (110) (Figure 2). Within these layers, short TlTl distances of 3.718(1) Å and 3.882(1) Å are found. In Tl2ADC (ADC2: acetylenedicarboxylate) the shortest TlTl distances range from 3.928 to 3.992 Å.47 These TlO layers are further connected by tfBDC2 ligands to form a 3D coordination polymer. Using the bond valence sum formalism,44 a valence V(Tl) = 0.83 is calculated for Tl1. This is close to the expected value for monovalent Tl. The distances and angles within the tfBDC ligand are as expected (Table 2). The ligand exhibits ci symmetry; both carboxylate groups are parallel to each other. The planes spanned by the aryl moiety and the carboxylate groups enclose an angle of 43.0(6)°. It is remarkable that in the crystal structure of 1 all aryl rings of the tfBDC ligand are coplanar to each other. In several publications it was shown that compounds containing a cation with a stereochemically active lone pair show an enhanced probability to crystallize in noncentrosymmetric space groups.4547 1 however crystallizes in centrosymmetric space group P1. DTA/TG investigations on 1 show that the compound is stable up to almost 200 °C (Figure 3). Between 200 and 300 °C,

ARTICLE

a weight loss of approximately 32% occurs. Assuming that Tl2O remains a theoretical weight loss of 34% is calculated. This seems to be in reasonable agreement. By heating 1 in an argon atmosphere for several hours we obtained a black residue. Its X-ray powder pattern is not in agreement with Tl2O or any other thallium oxide. As we have not been able to identify the residue, the decomposition of 1 remains unclear. Comparing the IR spectrum recorded on our sample (Figure S8, Supporting Information) with previously published data,48 a good agreement can be found. This states that already in this work from 1972 1 was successfully synthesized. But that time no structural details were given. IR48: 1590vs (br), 1450s, 1370vs (br), 1279w, 1264vs, 1000 and 993vs, 900m, 795s, 750s, and 675w. IR[this work]: 1670m, 1587s, 1485w, 1450m, 1429m, 1367s, 1255w, 1045w, 989s, 919w, 889w, 796m, 733s, 650m, 465m. 2 ∞[Pb(tfBDC)(H2O)3] 3 1/2 H2O (2). 2 crystallizes in the triclinic space group P1 as colorless blocks. The lead cation is surrounded by nine oxygen atoms (Figure 4): three belong to water molecules (PbO: 2.628 Å, 2.680 Å, 2.906 Å), two to monodentately (PbO: 2.592 Å, 2.778 Å), and four to bidentately chelating coordinating carboxylate groups (PbO: 2.608 Å, 2.693 Å, 2.808 Å, 2.817 Å). The next-nearest PbO distance is at 4.314 Å. Using the bond valence sum formalism,44 a valence V(Pb) = 1.79 is calculated for Pb1 using PbO distances < 3.0 Å.

Figure 3. DTA (blue) and TG (black) curves of 3∞[Tl2(tfBDC)] (1).

Figure 1. ORTEP diagram of the coordination sphere around the thallium cation in 3∞[Tl2(tfBDC)] (1) showing 50% probability thermal ellipsoids and the atom-numbering scheme. TlO connections are drawn as broken lines.

Figure 2. Packing diagram of 1 in a view along [010] (Tl: blue, C: white, O: red, F: green). TlO connections are drawn as broken lines.

Figure 4. ORTEP diagram of the coordination sphere around the lead cation in 2∞[Pb(tfBDC)(H2O)3] 3 1/2 H2O (2) showing 50% probability thermal ellipsoids and the atom-numbering scheme. PbO connections are drawn as broken lines. 5057

dx.doi.org/10.1021/cg200974h |Cryst. Growth Des. 2011, 11, 5053–5063

Crystal Growth & Design

ARTICLE

2 ∞[Pb(tfBDC)-

Figure 5. Packing diagram of 2 (Pb: blue, C: white, O: red, F: green, H: white small balls). PbO connections are drawn as broken lines.

Figure 6. DTA (blue) and TG (black) curves of (H2O)3] 3 1/2 H2O (2).

This is close to the expected value for divalent Pb. The perfluoroterephthalate linkers connect the lead cations to a layer-like structure, which is shown in Figure 5. Within these layers characteristic PbO zigzag chains are found, which run along [100]. The shortest PbPb distances within these chains are 4.4842(7) Å and 4.6234(8) Å. The layers are held together by hydrogen bonds involving the water molecules O3 (O3 3 3 3 O8B = 2.812(5) Å) and O4 (O4 3 3 3 O8B = 2.797(7) Å). The distances within the tfBDC linker are as expected. Some selected distances are given in Table 2. The carboxylate groups are almost perpendicular to each other with an angle of 85.3(5)° between the planes of both groups. Between the planes spanned by the aryl moiety and the carboxylate groups angles of 62.6(3)° and 22.3(5)° are found. Remarkably, like in 1, all aryl rings of the tfBDC ligand are coplanar in the crystal structure of 2. As already mentioned the coordination sphere of the lead cation is filled by three water molecules. Despite the heavy scattering lead cation their hydrogen atoms could be unambiguously located from difference Fourier maps. They were refined with fixed OH distances (0.96 Å) and no further constraints. A remaining electron density was ascribed to another noncoordinating water molecule (O5), which is located on a special position with 1 symmetry (Wyckoff site 4a). Its large thermal ellipsoid (Figure 4) indicates a disorder around this site. No hydrogen atoms could be assigned to this oxygen atom. It seems to be weakly bonded via a hydrogen bridge to water molecule O2 (O5 3 3 3 O2 = 3.117(5) Å). A single crystal structure analysis at low temperatures (170 K) did not “freeze” the disorder of O5 significantly. Although powder diffraction data (Figure S2, Supporting Information) and elemental analysis are not in agreement with a high purity of the sample (we assume that by grinding the obtained crystals the sample starts losing crystal water), we performed a DTA/TG investigation on 2 (Figure 6). This shows an endothermal mass loss of approximately 11% between 70 and 100 °C. A theoretical mass loss of 10.7% is calculated for the loss of three water molecules and 12.5% for the loss of 3.5 water molecules. According to these data, a loss of three water molecules must be assumed indicating that the sample under investigation already has lost 1/2 water molecule compared to 2. This confirms our assumption that ground crystals of 2 readily lose 1/2 water molecule explaining the discrepancies in the elemental analysis and the XRPD investigations. A weak exothermal DTA signal is observed at 285 °C and a stronger exothermal

signal is observed at 330 °C. Both are accompanied by a mass loss of approximately 15%, which is due to a decomposition of the ligand. For a 2-fold decarboxylation of the ligand, a theoretical mass loss of 17% is calculated. For a complete decomposition to PbO, a remaining mass of 44% is calculated. Obviously heating to temperatures above 400 °C is necessary to achieve a complete decomposition of 2. 1 ∞[M(tfBDC)(H2O)4] (M = Zn (3), Co (4), Ni (5)). 35 crystallize in the triclinic space group P1 as colorless needles (3), pink plates (4), or a green crystalline powder (5). From XRPD data (Figure S5 in the Supporting Information), it can be concluded that 5 is isostructural to 3 and 4, whose crystal structures were solved from single crystal data. The respective lattice parameters are given in Table 3. In the single crystal structure investigations of 3 and 4, all hydrogen atoms of the water molecules were located from difference Fourier maps and refined with fixed OH distances (0.96 Å). No further constraints were applied. As 35 are isostructural, only the crystal structure of 3 will be discussed in detail. In 3 the zinc cation occupies a special position with site symmetry 1 (Wyckoff site 1a). Its coordination sphere is shown in Figure 7. Zn is surrounded by six oxygen atoms to form an almost undistorted octahedron. Four water molecules (ZnO: 2.079 Å, 2x; 2.095 Å, 2x) and two monodentately coordinating carboxylate groups of two tfBDC ligands (ZnO: 2.131 Å, 2x) form the first coordination sphere of Zn2+. The carboxylate groups of the tfBDC ligands are trans coordinating. Thus chains are formed, which run along [111] (Figure 8). These chains are packed in form of a (distorted) hexagonal rod packing. The chains are linked by hydrogen bonds among each other. The shortest hydrogen bonds (O1A 3 3 3 O3 = 2.823(3) Å) connect the chains to layers. Longer hydrogen bonds (O1B 3 3 3 O3 = 2.939(3) Å) connect these layers to a 3D structure. In some respects, the crystal structure of 3 is similar to the recently published crystal structure of 1∞[Zn(tfBDC)(MeOH)4].34 Here also a polymeric structure with a chain-like structural arrangement is found, but due to a cis coordination of the carboxylate groups at the central zinc cation 1D zigzag chains are formed. The distances and angles within the tfBDC ligand are as expected (Table 2). Both carboxylate groups are coplanar. They enclose an angle of 47.5(2)° (47.2(3)° for 4) with the aryl moiety. Like in 1 and 2 these aryl rings are coplanar in the crystal structure of 3. The MO distances (M = Zn, Co) as well the lattice parameters (Table 3) in 3 and 4 are very similar reflecting 5058

dx.doi.org/10.1021/cg200974h |Cryst. Growth Des. 2011, 11, 5053–5063

Crystal Growth & Design

ARTICLE

Table 3. Unit Cell Parameters of Compounds 35 3

4

5

method a

single crystal, Stoe IPDS I, Mo Kα 5.194(1) Å

single crystal, Stoe IPDS I, Mo Kα 5.169(1) Å

powder, Huber G670, Cu Kα1 5.155(2) Å

b

6.420(1) Å

6.409(2) Å

6.368(2) Å

c

9.384(2) Å

9.396(3) Å

9.358(3) Å

α

73.02(2)°

72.66(3)°

72.76(2)°

β

88.35(2)°

88.67(3)°

88.65(2)°

γ

76.35(2)°

76.90(3)°

76.78(3)°

volume

290.6(1) Å3

289.1(1) Å3

285.3(1) Å3

Figure 7. ORTEP diagram of the coordination sphere around the zinc cation in 1∞[Zn(tfBDC)(H2O)4] (3) showing 50% probability thermal ellipsoids and the atom-numbering scheme. ZnO connections are drawn as broken lines. Figure 9. DTA (blue) and TG (black) curves of (H2O)4] (3).

Figure 8. Packing diagram of 3 in a view along [010] (Zn: blue, C: white, O: red, F: green, H: white small balls). ZnO connections are drawn as white broken lines. Hydrogen bonds are emphasized as red broken lines.

the comparable ionic radii of Zn2+ and Co2+ (high spin). For 5 a distinctively smaller unit cell is found (Table 3), again in agreement with the smaller ionic radii of Ni2+ compared to Zn2+ and Co2+ (h.s.). As 3 and 4 can be synthesized in a mechanochemical reaction without any solvent, it is apparent that water is incorporated from the crystal water of the starting materials Zn(CH3COO)2 3 2H2O and Co(CH3COO)2 3 4H2O. DTA/TG investigations were performed on compound 3 (Figure 9). Two endothermal signals are observed at 100 and 200 °C accompanied by a mass loss of 12.5% and 4%, respectively. For the loss of one water molecule, a theoretical mass loss of 4.8% is calculated. Thus, one can expect that in the first step at 100 °C approximately three water molecules are released and in the next step at 200 °C another water molecule. The calculated relative mass for the potential anhydrous compound Zn(tfBDC) is 80.7%, which is in good agreement with the relative mass

1 ∞[Zn(tfBDC)-

detected at 200 °C. Between 350 and 400 °C, an exothermal signal with a mass loss of approximately 22% is observed with a remaining relative mass of 61.3%. For ZnO, a relative mass of 21.8% is calculated. This shows that up to 500 °C no complete decomposition of 3 occurs. For a 2-fold decarboxylation of 3 a theoretical mass loss of 23.6% is calculated, which is in reasonable agreement with the mass loss observed between 350 and 400 °C. To understand this decomposition in more detail, further experiments, which include the analysis of the evolved gases via mass spectroscopy, are necessary. But these efforts seem to be worthwhile, as the loss of coordinating water molecules must lead to a structural rearrangement, which might result in coordination polymers with frameworks of higher dimensionality. Therefore we investigated two samples heated at 125 °C (3a) and 225 °C (3b) by XRPD and elemental analysis. In the XRPD patterns we observed a reasonable crystallinity of these samples (Figure 10). 3a shows sharp reflections, whereas the reflections of 3b are significantly broadened. Nonetheless, we were able to index the pattern of 3b with a C-centered monoclinic unit cell.49 Currently, we are trying to elucidate the crystal structure of 3b with different powder diffraction techniques. However, the data of 3a could not be indexed up to now. It cannot be excluded that this sample is not single-phase. Elemental analysis (Table 4) shows a good agreement between measured and calculated values for a composition Zn(tfBDC) 3 2.5 H2O (3a) and anhydrous Zn(tfBDC) (3b). For the latter, these results are in good agreement with the results of the DTA/TG analysis (see above). However, for compound 3a an obvious discrepancy is found in these two measurements, for which no reasonable explanation can be given at the moment. 5059

dx.doi.org/10.1021/cg200974h |Cryst. Growth Des. 2011, 11, 5053–5063

Crystal Growth & Design

ARTICLE

Figure 10. Comparison of selected powder diffraction patterns (background corrected) of 3: as synthesized (Huber 670, flat sample, 20 min; black, below), after heating at 125 °C (Huber 670, capillary, 120 min; red, middle) and after heating at 225 °C (Huber 670, capillary, 120 min; blue, above).

Table 4. Elemental Analysis of 1∞Zn(tfBDC)(H2O)4 (3) after Heating at 125 and 225 °C; Comparison with Several Potential Dehydration Products sample as synthesized

heated at 125 °C

heated at 225 °C

found C (%)

found H (%)

composition

calcd C (%)

calcd H (%)

25.38

1.94

Zn(tfBDC) 3 4 H2O Zn(tfBDC) 3 3.5 H2O

25.72

2.16

26.36

1.94

Zn(tfBDC) 3 3 H2O Zn(tfBDC) 3 2.5 H2O

27.03 27.73

1.70 1.45

Zn(tfBDC) 3 2 H2O Zn(tfBDC) 3 1.5 H2O

28.47

1.19

29.25

0.92

Zn(tfBDC) 3 H2O

30.07

0.63

Zn(tfBDC) 3 1/2 H2O

30.95

0.32

Zn(tfBDC)

31.87

0

27.66

31.10

3 ∞[M2(tfBDC)2(DMF)2(EtOH)]

1.22

0.06

(M = Zn (6), Mn (7)). During preparation of this manuscript, we became aware of a very recent publication of Cheng et al.,34 in which the crystal structure of 6 was described. This structural analysis is in agreement with our former results already presented at a conference.50 Therefore, we will restrict our brief discussion on the crystal structure of 7, which is isostructural to 6. 6 and 7 crystallize in monoclinic acentric space group P21 as colorless needles and light brown blocks, respectively. All hydrogen atoms of dimethylformamide as well as methyl and methylene hydrogen atoms of ethanol were placed on calculated positions and refined with a “riding” model with fixed CH distances (CH: 0.93 Å, CH2: 0.97 Å, CH3: 0.96 Å). The position of the hydrogen atom of the hydroxyl group of ethanol was located from difference Fourier maps and refined with a “riding” model with a fixed OH distance (0.82 Å). No further constraints were applied. In both compounds, one of the two DMF

molecules is disordered. It was refined with a split position for C22H22 and C22AH22A with 60%:40%. 7 is an inversion twin and was refined accordingly. In Figure 11 an ORTEP plot of 7 is shown. Both symmetry independent manganese cations occupy general position 2a. Mn1 is coordinated by five oxygen atoms with MnO distances ranging from 2.110(4) Å to 2.203(4) Å (Table 2) stemming from the carboxylate groups of four different tfBDC2 linkers and one ethanol molecule. An additional oxygen atom with a MnO distance of 2.502(4) Å stems from one of the carboxylate groups mentioned above. Thus the latter is coordinating in a chelating mode. The coordination polyhedron around Mn1 is somewhere in-between an octahedron and a trigonal bipyramid with an additional oxygen atom capping one of the edges of the equatorial trigonal plane (OMnO: 106.0(2)°, 109.0(2)°, 145.0(2)°). Mn2 is coordinated by six oxygen atoms (MnO: 2.107(7)2.192(4) Å) forming a slightly distorted octahedron 5060

dx.doi.org/10.1021/cg200974h |Cryst. Growth Des. 2011, 11, 5053–5063

Crystal Growth & Design

Figure 11. ORTEP diagram of the coordination sphere around the manganese cations in 3∞[M2(tfBDC)2(DMF)2(EtOH)] (7) showing 50% probability thermal ellipsoids and the atom-numbering scheme. MnO connections are drawn as broken lines. Solvent molecules (DMF, EtOH) as well as tfBDC ligands not belonging to the asymmetric unit are shown in a wire representation. The disorder of one DMF molecule (C22H22/C22AH22A = 60%:40%) has been omitted.

(OMnO: 86.1(2)98.6(2)°). These oxygen atoms stem from carboxylate groups of four different ftBDC2 linkers and two DMF molecules. Both MnO6 octahedra are bridged by carboxylate groups to form a dimer (Mn1Mn2 = 3.596(1) Å). More details of the coordination modes of the bridging carboxylates are given in ref 34. These dimers are connected within the (010) plane and perpendicular to it via tfBDC2 linkers to form a 3D framework (Figure 12). The topology of this framework is discussed elsewhere.34 Elemental analysis and X-ray powder diffraction data reveal that 6 and 7 were obtained as samples of modest purity. No further investigations were performed on these compounds up to now.

’ CONCLUSION We have presented the crystal structures of six new coordination polymers with tetrafluoroterephthalate (tfBDC2) as a bridging ligand. Although none of them are porous, 3 ∞[Tl2(tfBDC)] (1) represents one of the very rare examples of a homoleptic coordination polymer with a perfluorinated linker.32 3D structures are found for 1, 6, and 7, whereas 2D (2) and 1D structures (35) are found for the others. No obvious trend of the resulting dimensionality of the framework from the synthetic conditions was found, for example, 1 is synthesized in H2O, whereas 6 and 7 are obtained in EtOH/DMF. Similarly, the ratio M/tfBDC in the starting mixture is not well reflected in the resulting coordination polymer, for example, 1∞[Zn(tfBDC)(H2O)4] (3) (M/tfBDC = 1:1) is obtained from a starting solution with a M/tfBDC = 0.86:1 as well as 3.07:1. Thus, it is still challenging to obtain a coordination polymer with a specific dimensionality just by rational design.

ARTICLE

Figure 12. Packing diagram of 7 in a view along [100] (Mn: light blue, C: white, O: red, F: green, N: blue, H: white small balls). MnO connections are drawn as white broken lines and MnO6 polyhedra are emphasized.

But probably the most important goal is the synthesis of porous coordination polymers (MOFs) with perfluorinated linkers. It has already been stated that the crystal chemistry of nonfluorinated terephthalate (BDC2) cannot be transferred to tfBDC2.32,36 In fluorinated tfBDC2 and its coordination polymers the carboxylate groups are twisted out of the plane of the benzene ring and the respective torsion angles are considerably enlarged (4560°) compared to coordination polymers with BDC2 ligands, where torsion angles near 0° are observed.32 This was attributed to an electrostatic repulsion between the fluorine atoms on the ring and the oxygen atoms of the carboxylate groups as well as a decrease in aromatic character of the carboxylate group due to the electron-withdrawing nature of the fluorine atoms.32,36 Preliminary gas-phase calculations (0 K)51 on BDC2 and tfBDC2 using TURBOMOLE 5.10 with the BP86 functional and def2-TZVP basis sets52 confirm these findings and assumptions. For BDC2 a planar conformation is calculated as the most stable one, whereas for tfBDC2 conformations with torsion angles > 0° are more stable. In fact, an orthogonal conformation (torsion angle = 90°) was calculated to be the most stable one. To further confirm these findings, we have searched the Cambridge Structural Database (CSD) for all compounds containing either BDC2 or tfBDC2. An analysis of the torsion angles between the carboxylate groups and the plane of the benzene ring is shown in Figure 13. It is obvious that much more compounds are known with BDC2 than with tfBDC2. But despite the much smaller number of compounds with tfBDC2 the trend is clear: in compounds containing BDC2 most of the torsion angles found are close to zero. In compounds containing tfBDC2 however significantly larger torsion angles are found (>30°). But there is no clear preference of any of these larger torsion angles. Therefore, it must be assumed that the total energies of these different conformations are close to each other, and depending on the packing restrictions within a crystal structure different torsion angles > 30° can be formed. The results of this work  shown in violet in Figure 13 (right)  fit 5061

dx.doi.org/10.1021/cg200974h |Cryst. Growth Des. 2011, 11, 5053–5063

Crystal Growth & Design

ARTICLE

Figure 13. Torsion angles between carboxylate groups and benzene rings in compounds containing terephthalate, BDC2 (left) or tetrafluoroterephthalate, tfBDC2 (right); the results shown in green were extracted from the Cambridge Structural Database (CSD). Results shown in violet are from this work.

nicely into these trends. All these results confirm that there is a distinct difference between the crystal chemistry of terephthalates and tetrafluoroterephthalates. It is obvious that the concepts of synthesizing MOFs with BDC2 as bridging ligands cannot be transferred to its perfluorinated congener. Furthermore it must be considered that the acidity of H2tfBDC is significantly higher than that of H2BDC. Therefore, it will be a big challenge to find the right synthetic conditions for the preparation of a perfluorinated MOF.

’ ASSOCIATED CONTENT

bS

Supporting Information. Experimental and simulated X-ray powder diffraction patterns of compounds 17, IR spectrum of compound 1, and X-ray crystallographic files in cif format for compounds 14, 6, 7. This material is available free of charge via the Internet at http://pubs.acs.org.

’ AUTHOR INFORMATION Corresponding Author

*Fax: (+49)221-470-3933. Phone: (+49)221-470-3285. E-mail: [email protected].

’ ACKNOWLEDGMENT We thank Dr. Ingo Pantenburg and Mrs. Ingrid M€uller for collecting X-ray single crystal data, Dr. J€org-Martin Neud€orfl for helpful discussions and advices concerning the X-ray single crystal structure analysis of compounds 6 and 7, Mr. Peter Kliesen for recording DTA/TG and IR data, Mr. Horst Schumacher for recording some of the X-ray powder diffraction data, and Mrs. Silke Kremer for elemental analysis. ’ REFERENCES (1) Li, H.; Eddaoudi, M.; O’Keeffe, M.; Yaghi, O. M. Nature 1999, 402, 276–279. (2) Mueller, U.; Schubert, M.; Teich, F.; Puetter, H.; Schierle-Arndt, K.; Pastre, J. J. Mater. Chem. 2006, 16, 626–636. (3) Ferey, G. Chem. Soc. Rev. 2008, 37, 191–214.

(4) Czaja, A. U.; Trukhan, N.; Mueller, U. Chem. Soc. Rev. 2009, 38, 1284–1293. (5) Rowsell, J. L. C.; Yaghi, O. M. Angew. Chem., Int. Ed. 2005, 44, 4670–4679. (6) Collins, D. J.; Zhou, H.-C. J. Mater. Chem. 2007, 17, 3154–3160. (7) Morris, R. E.; Wheatley, P. S. Angew. Chem., Int. Ed. 2008, 47, 4966–4981. (8) van den Berg, A. W. C.; Arean, C. O. Chem. Commun. 2008, 668–681. (9) Eberle, U.; Felderhoff, M.; Sch€uth, F. Angew. Chem., Int. Ed. 2009, 48, 6608–6630. (10) Ma, S.; Zhou, H.-C. Chem. Commun. 2010, 46, 44–53. (11) Yang, C.; Wang, X.; Omary, M. A. J. Am. Chem. Soc. 2007, 129, 15454–15455. (12) Yang, C.; Wang, X.; Omary, M. A. Angew. Chem., Int. Ed. 2009, 48, 2500–2505. (13) Fischer, R. A.; W€oll, C. Angew. Chem., Int. Ed. 2008, 47, 8164–8168. (14) Hulvey, Z.; Falcao, E. H. L.; Eckert, J.; Cheetham, A. K. J. Mater. Chem. 2009, 19, 4307–4309. (15) Uemura, K.; Maeda, A.; Maji, T. K.; Kanoo, P.; Kita, H. Eur. J. Inorg. Chem. 2009, 2329–2337. (16) Loiseau, T.; Serre, C.; Huguenard, C.; Fink, G.; Taulelle, F.; Henry, M.; Bataille, T.; Ferey, G. Chem.—Eur. J. 2004, 10, 1373–1382. (17) Zhang, L.; Wang, Q.; Liu, Y.-C. J. Phys. Chem. B 2007, 111, 4291–4295. (18) H€ubner, O.; Gl€oss, A.; Fichtner, M.; Klopper, W. J. Phys. Chem. A 2004, 108, 3019–3023. (19) Hulvey, Z.; Sava, D. A.; Eckert, J.; Cheetham, A. K. Inorg. Chem. 2011, 50, 403–405. (20) Chisholm, M. H.; Wilson, P. J.; Woodward, P. M. Chem. Commun. 2002, 566–567. (21) Kitaura, R.; Iwahori, F.; Matsuda, R.; Kitagawa, S.; Kubota, Y.; Takata, M.; Kobayashi, T. C. Inorg. Chem. 2004, 43, 6522–6524. (22) Liu, S. Q.; Konaka, H.; Kuroda-Sowa, T.; Maekawa, M.; Suenaga, Y.; Ning, G. L.; Munakataa, M. Acta Crystallogr. 2004, E60, m1504–m1506. (23) Ito, M.; Onaka, S. Inorg. Chim. Acta 2004, 357, 1039–1040. (24) Chun, H.; Dybtsev, D. N.; Kim, H.; Kim, K. Chem.—Eur. J. 2005, 11, 3521–3529. (25) Chen, B.; Yang, Y.; Zapata, F.; Qian, G.; Luo, Y.; Zhang, J.; Lobkovsky, E. B. Inorg. Chem. 2006, 45, 8882–8886. (26) Yoon, J. H.; Choi, S. B.; Oh, Y. J.; Seo, M. J.; Jhon, Y. H.; Lee, T.-B.; Kim, D.; Choi, S. H.; Kim, J. Catal. Today 2007, 120, 324–329. (27) Zhu, E.-J.; Liu, Q.; Chen, Q.; He, M.-Y.; Chen, S.-C.; Huang, H.-X.; Yang, Q. Wuji Huaxue Xuebao (Chin. J. Inorg. Chem.) 2008, 24, 1428–1433. 5062

dx.doi.org/10.1021/cg200974h |Cryst. Growth Des. 2011, 11, 5053–5063

Crystal Growth & Design

ARTICLE

(28) Zheng, C.-G.; Hong, J.-Q.; Zhang, J.; Wang, C. Acta Crystallogr. 2008, E64, m879. (29) Zheng, C.-G.; Zhang, J.; Hong, J.-Q.; Li, S. Acta Crystallogr. 2008, E64, m965. (30) Hulvey, Z.; Ayala, E.; Furman, J. D.; Forster, P. M.; Cheetham, A. K. Cryst. Growth Des. 2009, 9, 4759–4765. (31) Hulvey, Z.; Ayala, E.; Cheetham, A. K. Z. Anorg. Allg. Chem. 2009, 635, 1753–1757. (32) Hulvey, Z.; Furman, J. D.; Turner, S. A.; Tang, M.; Cheetham, A. K. Cryst. Growth Des. 2010, 10, 2041–2043. (33) MacNeill, C. M.; Day, C. S.; Marts, A.; Lachgar, A.; Noftle, R. E. Inorg. Chim. Acta 2011, 365, 196–203. (34) Cheng, M.-L.; Zhu, E.; Liu, Q.; Chen, S.-C.; Chen, Q.; He, M.-Y. Inorg. Chem. Commun. 2011, 14, 300–303. (35) He, M.-Y.; Xu, H.; Lu, L.-D.; Yang, X.-J.; Wang, X. Z. Kristallogr. New Cryst. Struct. 2009, 224, 619–620. (36) Wang, Z.; Kravtsov, V. C.; Walsh, R. B.; Zaworotko, M. J. Cryst. Growth Des. 2007, 7, 1154–1162. (37) Orthaber, A.; Seidel, C.; Belaj, F.; Pietschnig, R.; Ruschewitz, U. Inorg. Chem. 2010, 49, 9350–9357. (38) Stoe, IPDS manual; Stoe & Cie GmbH: Germany; X-Red 1.22Stoe Data Reduction Program; Stoe & Cie GmbH: Germany, 2001. (39) Altomare, A.; Cascarano, G.; Giacovazzo, C.; Gualardi, A. J. Appl. Crystallogr. 1993, 26, 343–350. (40) Sheldrick, G. M. SHELXL-97: A Program for Crystal Structure Refinement; University of G€ottingen: Germany, 1997; release 972; Sheldrick, G. M. Acta Crystallogr. 2008, A64, 112–122. (41) WinGX-Version 1.64.04, An Integrated System of Windows Programs for the Solution, Refinement and Analysis of Single Crystal X-Ray Diffraction Data; Department of Chemistry, University of Glasgow: UK, 19972002; Farrugia, L. J. J. Appl. Crystallogr. 1999, 32, 837–838. (42) Crystallographic data (excluding structure factors) for the structures reported in this paper have been deposited with the Cambridge Crystallographic Data Centre as supplementary publication no. CCDC-823594 (ν3∞[Tl2(tfBDC)], 1), CCDC-823593 (2∞[Pb(tfBDC)(H2O)3] 3 1/2 H2O, 2), CCDC-823595 (1∞[Zn(tfBDC)(H2O)4], 3), CCDC-823591 (1∞[Co(tfBDC)(H2O)4], 4), CCDC-823596 (3∞[Zn2(tfBDC)2(DMF)2(EtOH)], 6), and CCDC-823592 (3∞[Mn2(tfBDC)2(DMF)2(EtOH)], 7). Copies of the data can be obtained free of charge on application to CCDC, 12 Union Road, Cambridge CB2 1EZ, UK [fax.: (internat.) + 44 1223/336-033; e-mail: [email protected]]. (43) Win XPOW, version 1.04 (07-Jan-1999); Stoe & Cie GmbH: Darmstadt, Germany. (44) Brese, N. E.; O’Keeffe, M. Acta Crystallogr. 1991, B47, 192–197. (45) Ra, H.-S.; Ok, K. M.; Halasyamani, P. S. J. Am. Chem. Soc. 2003, 125, 7764–7765. (46) Ok, K. M.; Halasyamani, P. S. Angew. Chem., Int. Ed. 2004, 43, 5489–5491. (47) Ahlers, R.; Ruschewitz, U. Solid State Sci. 2009, 11, 1058–1064. (48) Albrecht, H. B.; Deacon, G. B. Aust. J. Chem. 1972, 25, 57–65. (49) Visser, J. W. J. Appl. Crystallogr. 1969, 2, 89–95. (50) Seidel, C.; Ruschewitz, U. 2nd International Conference on MetalOrganic Frameworks and Open Framework Compounds, Marseille/France, 5.-8. September 2010, Poster P 8. (51) Weißmann, D.; Dolg, M., private communications. (52) Ahlrichs, R.; B€ar, M.; H€aser, M.; Horn, H.; K€olmel, C. Chem. Phys. Lett. 1989, 162, 165–169. Dirac, P. A. M. Proc. R. Soc. London A 1929, 123, 714–733. Slater, J. C. Phys. Rev. 1951, 81, 385–390. Vosko, S.; Wilk, L.; Nusair, M. Can. J. Phys. 1980, 58, 1200–1211. Becke, A. D. Phys. Rev. B 1988, 38, 3098–3100. Perdew, J. P. Phys. Rev. B 1986, 33, 8822–8824. Weigend, F.; Ahlrichs, R. Phys. Chem. Chem. Phys. 2005, 7, 3297–3305.

5063

dx.doi.org/10.1021/cg200974h |Cryst. Growth Des. 2011, 11, 5053–5063