Dicarboxylate anion-dependent assembly of Ni(II) coordination polymers with 4,4′-dipyridyl sulfide

7
Dicarboxylate anion-dependent assembly of Ni(II) coordination polymers with 4,4 0 -dipyridyl sulfideLu-Fang Ma, ab Li-Ya Wang, * a Jiang-Liang Hu, a Yao-Yu Wang, * b Stuart R. Batten c and Jian-Ge Wang a Received 11th September 2008, Accepted 8th December 2008 First published as an Advance Article on the web 21st January 2009 DOI: 10.1039/b815916e Four new coordination polymers, [Ni(isop)(bps)(H 2 O)] 2n $5H 2 O(1), [Ni(O 2 N–BDC)(bps)(H 2 O)] 2n (2), [Ni(OH–BDC)(bps) 2 (H 2 O)] n (3) and {[Ni 2 (H 2 O)(tbip) 2 (bps) 2 ]} n (4) (bps ¼ 4,4 0 -dipyridyl sulfide, H 2 isop¼ isophthalic acid, O 2 N–BDC ¼ 5-nitroisophthalic acid, OH–BDC ¼ 5-hydroxyisophthalic acid, H 2 tbip ¼ 5-tert-butyl isophthalic acid), have been isolated from parallel hydrothermal reactions of Ni(OAc) 2 $4H 2 O and bps with isop, O 2 N–BDC, OH–BDC and H 2 tbip, respectively. Compound 1 is an unusual 2D double layered supramolecular motif generated by hydrogen bonding interactions of two single 2D (4,4) networks. Compound 2 has a highly puckered 2D (4,4) sheet and further stack via O/O intermolecular contacts to form a 3D supramolecular structure. Compound 3 features a 2D rectangular grid layer with terminal bps ligands projecting into the interlayer space, leading to a highly interdigitated packing motif and further expansion to generate a 3-D network through interlayer C–H/p stacking interactions. Compound 4 contains [Ni 2 (H 2 O)] 4+ dimers bridged by pairs of ligands into a single 3D diamond network with alternately left-handed and right-handed helical channels. As compared to isop, the coexistence of electron-donating (–OH, –C(CH 3 ) 3 ) or electron-withdrawing (–NO 2 ) groups in the dicarboxylate derivatives has a significant effect on the molecular self-assembly. Variable-temperature magnetic susceptibility measurements reveals the existence of weak ferromagnetic interactions between the nickel centers in 4. Introduction Crystal engineering based on metal–organic frameworks has recently attracted considerable interest, owing to their potential applications as functional materials in fluorescent sensing, molecular magnetism, gas sorption, optoelectronic devices and so on. 1–4 The diversity in the framework structures of such materials greatly depends on the metal centers and the structure of the spacer ligands, as well as on the reaction pathway. 5,6 Among diverse elegant efforts to find key factors in their devel- opment, the choice of bridging ligand backbone is worthy of close attention as a rational design strategy. As a result, many types of bridging ligands, such as dpe (1,2-bi(4-pyr- idyl)ethene), bpa (1,2-bi(4-pyridyl)ethane) and bpp (1,3-di(4- pyridyl)propane), with different carbon backbones between the two 4-pyridyl rings, have been extensively used in the construc- tion of diverse multidimensional architectures. 7 However, 4,4 0 - dipyridyl sulfide (dps), which possesses a non-linear backbone due to the bend around the sulfur atom, has seldom been reported. 8 On the other hand, benzene-1,3-dicarboxylic acid (isophthalic acid, H 2 isop) and its derivatives with special conformations, such as with a 120 angle between two carboxylic groups, present versatile coordination modes, can yield pre- determined networks and have been widely utilized to construct coordination polymers. 9,10 Yaghi has reported the synthesis of a porous metal–organic polyhedron which is constructed from 12 paddle-wheel units bridged by m-isophthalic acid to give a large metal–carboxylate polyhedron. 11 Zaworotko has demonstrated how the use of metal–organic secondary building units (SBUs) that are linked by an angular m-isophthalic acid ligand can generate nanoscale SBUs with curvature. 12 Compared to m-isophthalic acid, coexistent noncoordinated groups as electron-withdrawing and electron-donating groups have a profound impact on the electron density of whole ligands, and thereby different physical phenomena can be produced. 13 With the aim of preparing new materials with interesting structural topologies and physicochemical properties, we chose 4,4 0 -dipyridyl sulfide along with benzene-1,3-dicarboxylic acid and its derivatives (including derivatives also containing electron-donating (–OH, –C(CH 3 ) 3 ) or electron-withdrawing groups (–NO 2 ) groups) to construct new metal coordination polymers. In the present paper, four different coordination polymers, [Ni(isop)(bps)(H 2 O)] 2n $5H 2 O (1), [Ni(O 2 N– BDC)(bps)(H 2 O)] 2n (2), [Ni(OH–BDC)(bps) 2 (H 2 O)] n (3) and {[Ni 2 (H 2 O)(tbip) 2 (bps) 2 ]} n (4), are reported. Experimental Materials and physical measurements All reagents used in the syntheses were of analytical grade. Elemental analyses for carbon, hydrogen and nitrogen atoms a College of Chemistry and Chemical Engineering, Luoyang Normal University, Luoyang, 471022, P. R. China. E-mail: [email protected] b Key Laboratory of Synthetic and Natural Functional Molecule Chemistry of Ministry of Education, Department of Chemistry, Northwest University, Xi’an, 710069, P. R. China. E-mail: [email protected] c School of Chemistry, Monash University, Victoria, 3800, Australia † Electronic supplementary information (ESI) available: X-Ray crystallographic files in CIF format, additional figures, PXRD pattern of 4. CCDC reference numbers 701845–701848. For ESI and crystallographic data in CIF or other electronic format see DOI: 10.1039/b815916e This journal is ª The Royal Society of Chemistry 2009 CrystEngComm, 2009, 11, 777–783 | 777 PAPER www.rsc.org/crystengcomm | CrystEngComm

Transcript of Dicarboxylate anion-dependent assembly of Ni(II) coordination polymers with 4,4′-dipyridyl sulfide

PAPER www.rsc.org/crystengcomm | CrystEngComm

Dicarboxylate anion-dependent assembly of Ni(II) coordination polymerswith 4,40-dipyridyl sulfide†

Lu-Fang Ma,ab Li-Ya Wang,*a Jiang-Liang Hu,a Yao-Yu Wang,*b Stuart R. Battenc and Jian-Ge Wanga

Received 11th September 2008, Accepted 8th December 2008

First published as an Advance Article on the web 21st January 2009

DOI: 10.1039/b815916e

Four new coordination polymers, [Ni(isop)(bps)(H2O)]2n$5H2O (1), [Ni(O2N–BDC)(bps)(H2O)]2n (2),

[Ni(OH–BDC)(bps)2(H2O)]n (3) and {[Ni2(H2O)(tbip)2(bps)2]}n (4) (bps ¼ 4,40-dipyridyl sulfide,

H2isop¼ isophthalic acid, O2N–BDC ¼ 5-nitroisophthalic acid, OH–BDC ¼ 5-hydroxyisophthalic

acid, H2tbip ¼ 5-tert-butyl isophthalic acid), have been isolated from parallel hydrothermal reactions

of Ni(OAc)2$4H2O and bps with isop, O2N–BDC, OH–BDC and H2tbip, respectively.

Compound 1 is an unusual 2D double layered supramolecular motif generated by hydrogen bonding

interactions of two single 2D (4,4) networks. Compound 2 has a highly puckered 2D (4,4) sheet and

further stack via O/O intermolecular contacts to form a 3D supramolecular structure.

Compound 3 features a 2D rectangular grid layer with terminal bps ligands projecting into the

interlayer space, leading to a highly interdigitated packing motif and further expansion to generate

a 3-D network through interlayer C–H/p stacking interactions. Compound 4 contains [Ni2(H2O)]4+

dimers bridged by pairs of ligands into a single 3D diamond network with alternately left-handed and

right-handed helical channels. As compared to isop, the coexistence of electron-donating (–OH,

–C(CH3)3) or electron-withdrawing (–NO2) groups in the dicarboxylate derivatives has a significant

effect on the molecular self-assembly. Variable-temperature magnetic susceptibility measurements

reveals the existence of weak ferromagnetic interactions between the nickel centers in 4.

Introduction

Crystal engineering based on metal–organic frameworks has

recently attracted considerable interest, owing to their potential

applications as functional materials in fluorescent sensing,

molecular magnetism, gas sorption, optoelectronic devices and

so on.1–4 The diversity in the framework structures of such

materials greatly depends on the metal centers and the structure

of the spacer ligands, as well as on the reaction pathway.5,6

Among diverse elegant efforts to find key factors in their devel-

opment, the choice of bridging ligand backbone is worthy

of close attention as a rational design strategy. As a result,

many types of bridging ligands, such as dpe (1,2-bi(4-pyr-

idyl)ethene), bpa (1,2-bi(4-pyridyl)ethane) and bpp (1,3-di(4-

pyridyl)propane), with different carbon backbones between the

two 4-pyridyl rings, have been extensively used in the construc-

tion of diverse multidimensional architectures.7 However, 4,40-

dipyridyl sulfide (dps), which possesses a non-linear backbone

due to the bend around the sulfur atom, has seldom been

reported.8 On the other hand, benzene-1,3-dicarboxylic acid

(isophthalic acid, H2isop) and its derivatives with special

aCollege of Chemistry and Chemical Engineering, Luoyang NormalUniversity, Luoyang, 471022, P. R. China. E-mail: [email protected] Laboratory of Synthetic and Natural Functional Molecule Chemistryof Ministry of Education, Department of Chemistry, Northwest University,Xi’an, 710069, P. R. China. E-mail: [email protected] of Chemistry, Monash University, Victoria, 3800, Australia

† Electronic supplementary information (ESI) available: X-Raycrystallographic files in CIF format, additional figures, PXRD patternof 4. CCDC reference numbers 701845–701848. For ESI andcrystallographic data in CIF or other electronic format see DOI:10.1039/b815916e

This journal is ª The Royal Society of Chemistry 2009

conformations, such as with a 120� angle between two carboxylic

groups, present versatile coordination modes, can yield pre-

determined networks and have been widely utilized to construct

coordination polymers.9,10 Yaghi has reported the synthesis of

a porous metal–organic polyhedron which is constructed

from 12 paddle-wheel units bridged by m-isophthalic acid to give

a large metal–carboxylate polyhedron.11 Zaworotko has

demonstrated how the use of metal–organic secondary building

units (SBUs) that are linked by an angular m-isophthalic acid

ligand can generate nanoscale SBUs with curvature.12

Compared to m-isophthalic acid, coexistent noncoordinated

groups as electron-withdrawing and electron-donating groups

have a profound impact on the electron density of whole ligands,

and thereby different physical phenomena can be produced.13

With the aim of preparing new materials with interesting

structural topologies and physicochemical properties, we chose

4,40-dipyridyl sulfide along with benzene-1,3-dicarboxylic acid

and its derivatives (including derivatives also containing

electron-donating (–OH, –C(CH3)3) or electron-withdrawing

groups (–NO2) groups) to construct new metal coordination

polymers. In the present paper, four different coordination

polymers, [Ni(isop)(bps)(H2O)]2n$5H2O (1), [Ni(O2N–

BDC)(bps)(H2O)]2n (2), [Ni(OH–BDC)(bps)2(H2O)]n (3) and

{[Ni2(H2O)(tbip)2(bps)2]}n (4), are reported.

Experimental

Materials and physical measurements

All reagents used in the syntheses were of analytical grade.

Elemental analyses for carbon, hydrogen and nitrogen atoms

CrystEngComm, 2009, 11, 777–783 | 777

were performed on a Vario EL III elemental analyzer. The

infrared spectra (4000�400 cm�1) were recorded by using a KBr

pellet on an Avatar� 360 E. S. P. IR spectrometer. Variable-

temperature magnetic susceptibilities were measured using

a MPMS-7 SQUID magnetometer. Diamagnetic corrections

were made with Pascal’s constants for all constituent atoms.

Preparation of complexes 1–4

[Ni(isop)(bps)(H2O)]2n$5H2O (1). A mixture of H2isop

(0.1 mmol, 16.5 mg), bps (0.1 mmol, 18.8 mg), Ni(OAc)2$4H2O

(24.8 mg, 0.1 mmol), KOH (0.1 mmol, 5.6 mg) and H2O 15 mL

was placed in a Teflon-lined stainless steel vessel, heated to 160�C for 3 d, and then cooled to room temperature over 24 h. Blue

block crystals of 1 were obtained. Yield: 0.024 g, 50%. Elemental

analysis (%): calcd for C36H38N4Ni2O15S2 C 45.60, H 4.04,

N 5.91; found C 45.68, H 4.16, N 5.87. IR (cm�1): 3287 m, 3098

w, 2448 m, 1590 s nas(COO�), 1541 m ns(COO�), 1480 m, 1388 s,

1217 m, 1155 w, 1063 m, 815 m, 744 m, 724 m, 532 w, 497 m.

[Ni(O2N–BDC)(bps)(H2O)]2n (2). 2 was synthesized in

a similar way as that described for 1, except that H2isop was

replaced by O2N–BDC. Yield: 0.018 g, 38%. Elemental analysis

(%): calcd for C36H26N6Ni2O14S2 C 45.60, H 2.76, N 8.86; found

C 45.57, H 2.71, N 8.74. IR (cm�1): 3635 m, 3103 m, 3063 w, 2475

m, 1623 s nas(COO�), 1598 m ns(COO�), 1460 m, 1420 m, 1355 s,

1211 m, 1026 m, 915 m, 887 m, 822 m, 787 m, 541 m, 498 m.

[Ni(OH–BDC)(bps)2(H2O)]n (3). 3 was synthesized in a similar

way as that described for 1, except that H2isop was replaced by

OH–BDC. Yield: 0.021 g, 33%. Elemental analysis (%): calcd for

C28H22N4NiO6S2 C 53.10, H 3.50, N 8.85; found C 53.05, H 3.47,

N 8.73. IR (cm�1): 3188 m, 3101 m, 1588 s nas(COO�), 1479 m

ns(COO�), 1411 m, 1368 m, 1273 s, 1214 m, 1062 m, 970 m,

809 m, 782 m, 722m, 540 m.

{[Ni2(H2O)(tbip)2(bps)2]}n (4). 4 was synthesized in a similar

way as that described for 1, except that H2isop was replaced by

Table 1 Crystallographic data for complexes 1–4

Compound 1 2

Formula C36H38N4Ni2O15S2 C36H26N6

FW 948.24 948.17Temperature 291(2) 291(2)Crystal system triclinic monocliniSpace group P-1 P2/n

a ¼ 10.2688(10)Unit cell b ¼ 11.1859(10) a ¼ 14.358

c ¼ 18.4444(17) b ¼ 8.2781Dimensions/�A, � a ¼ 80.255(10) c ¼ 15.898

b ¼ 83.122(1) b ¼ 93.735g ¼ 89.7290(10)

V/�A3 2072.8(3) 1885.6(3)Z 2 2r/g cm�3 1.519 1.670F(000) 980 968Flack parametersR1, wR2 [I > 2s (I)] 0.0371, 0.0969 0.0261, 0.0R1,wR2 (all data) 0.0466, 0.1036 0.0297, 0.0Residuals/e �A�3 0.644, -0.529 0.251, -0.5

778 | CrystEngComm, 2009, 11, 777–783

H2tbip. Yield: 0.014 g, 29%. Elemental analysis (%): calcd for

C44H42N4Ni2O9S2 C 55.49, H 4.45, N 5.88; found C 55.41, H

4.54, N 5.83. IR (cm�1): 3429 m, 2960 m, 2912 m, 1638 s

nas(COO�), 1588 m ns(COO�), 1545 m, 1416 m, 1369 s, 1215,

811 m, 777 m, 715 m.

X-Ray crystallography

Single crystal X-ray diffraction analyses of 1–4 were carried out

on a Bruker SMART APEX II CCD diffractometer equipped

with a graphite monochromated Mo Ka radiation (l ¼ 0.71073�A) by using the f/u scan technique at room temperature.

The structures were solved by direct methods with SHELXS-97.

The hydrogen atoms were assigned with common isotropic

displacement factors and included in the final refinement by use

of geometrical restraints. In 3, 48 restraints were used in the

refinement. The restraints are needed to restrict the large Ueq of

the pyridine ring (C14, C15, C16, C17, C18 and N2). In 4, the

carbon atoms of tert-butyl groups have departed into two parts

and their site occupation was refined by the programme. The

restraints are needed to restrict the large Ueq of tert-butyl

groups. The Flack parameter of 4 is refined to near zero (see

Table 1), which reveals that the absolute structure is correct.14 A

full-matrix least-squares refinement on F2 was carried out using

SHELXL-97. The crystallographic data and selected bond

lengths for 1–4 are listed in Tables 1–2. Crystallographic data for

the structural analysis have been deposited with the Cambridge

Crystallographic Data Center. CCDC reference numbers:

701845–701848.

Results and discussion

Description of crystal structures

[Ni(isop)(bps)(H2O)]2n$5H2O (1). Complex 1 features an

unusual 2D bilayer architecture generated by hydrogen-bonding

interactions of two single-layered networks. The asymmetric unit

of 1 contains two Ni2+ cations, two isop anions, two bps ligands,

two coordinated water molecules and five lattice water

3 4

Ni2O14S2 C28H22N4NiO6S2 C44H42N4Ni2O9S2

633.33 952.36291(2) 291(2)

c monoclinic monoclinicP2(1)/c Cc

a ¼ 15.3099(10)4(14) a ¼ 10.117(4) b ¼ 27.9468(18)(8) b ¼ 19.692(8) c ¼ 15.7112(10)1(15) c ¼ 15.346(6) b ¼ 114.862(1)(1) b ¼ 106.672(4)

2929(2) 6099.2(7)4 41.436 1.0371304 1976

0.003(14)681 0.0296, 0.0753 0.0543, 0.1458705 0.0350, 0.0795 0.0591, 0.150626 0.574, -0.470 0.985, -0.508

This journal is ª The Royal Society of Chemistry 2009

Fig. 1 (a) Coordination environment of Ni (II) ion in 1. The hydrogen atoms

2D layer of compound 1. (c) The 2D supramolecular bilayer motif in 1 assembl

Table 2 Selected bond lengths (�A) for 1–4a

1Ni(1)–O(8) 2.0015(19) Ni(2)–O(11)#2 1.9937(19)Ni(1)–N(1)#1 2.061(2) Ni(2)–N(3)#1 2.067(2)Ni(1)–N(2) 2.096(2) Ni(2)–N(4) 2.093(2)Ni(1)–O(7) 2.1019(19) Ni(2)–O(6) 2.104(2)Ni(1)–O(12) 2.1153(18) Ni(2)–O(14) 2.1069(18)Ni(1)–O(13) 2.1357(19) Ni(2)–O(15) 2.1668(19)2Ni(1)–N(2)#1 2.0369(15) Ni(2)–O(3)#4 2.0266(13)Ni(1)–N(2)#2 2.0369(15) Ni(2)–O(3) 2.0266(13)Ni(1)–O(1)#3 2.0462(12) Ni(2)–O(7)#4 2.1231(14)Ni(1)–O(1) 2.0462(12) Ni(2)–O(7) 2.1231(14)Ni(1)–O(2) 2.2317(13) Ni(2)–N(1) 2.1299(16)Ni(1)–O(2)#3 2.2317(13) Ni(2)–N(1)#4 2.1299(15)3Ni(1)–O(3)#1 2.0324(14) Ni(1)–N(3) 2.1081(18)Ni(1)–O(1) 2.0629(13) Ni(1)–N(4)#2 2.1180(16)Ni(1)–O(6) 2.0772(14) Ni(1)–N(1) 2.1313(18)4Ni(1)–O(8)#1 2.029(3) Ni(2)–O(2) 2.030(3)Ni(1)–O(1) 2.035(3) Ni(2)–O(7)#1 2.043(3)Ni(1)–O(5) 2.078(3) Ni(2)–O(4)#2 2.066(3)Ni(1)–O(9) 2.103(3) Ni(2)–N(3) 2.087(4)Ni(1)–N(2)#2 2.109(4) Ni(2)–N(4)#3 2.091(4)Ni(1)–N(1) 2.110(4) Ni(2)–O(9) 2.102(3)

a Symmetry transformations used to generate equivalent atoms: #1 x + 1,y, z, #2 x, y + 1, z� 1 for 1; #1�x + 2,�y + 2,�z + 2, #2 x� 1/2,�y + 2,z + 1/2, #3�x + 3/2, y, �z + 5/2, #4�x + 3/2, y, �z + 3/2 for 2, #1 �x +1, y� 1/2,�z + 1/2, #2 x� 1, y, z for 3, #1 x + 1/2,�y + 1/2, z + 1/2, #2 x,�y, z � 1/2, #3 x � 1/2, �y + 1/2, z � 1/2 for 4.

This journal is ª The Royal Society of Chemistry 2009

molecules. As shown in Fig. 1a, both nickel atoms are octahe-

drally coordinated by two nitrogen atoms of two bps ligands and

three oxygen atoms of two isop ligands as well as one water

molecule. The Ni–O and Ni–N bond lengths are in the range of

1.994(2)–2.167(2) and 2.061(2)–2.096(2) �A, respectively. The isop

ligands show both bis-monodentate and bis-chelating modes

(Scheme 1a and 1b) to bridge the Ni (II) ions to form a chain with

alternating ligand and metal types. These chains are cross-linked

by bridging bps ligands along the b axis to generate a 2D (4,4)

rectangular grid layer (Fig. 1b). Several kinds of hydrogen

and free water are omitted for clarity. Symmetry codes: A, 1 + x, y, z. (b)

ed by hydrogen bonds of adjacent nets encapsulating (H2O)10 clusters (d).

Scheme 1 The varied coordination modes of benzene-1,3-dicarboxylate

and derivatives observed in 1–4.

CrystEngComm, 2009, 11, 777–783 | 779

bonding are present: (a) hydrogen bonding between coordinated

water molecules and coordinated carboxylate oxygen atoms with

O/O distances of 2.742(3) and 2.757(3) �A, respectively;

(b) hydrogen bonding between free water molecules with O/O

distances in the range of 2.519(17)–3.034(9) �A; (c) hydrogen

bonding between free water molecules and uncoordinated

carboxylate oxygen atoms with O/O distances of 2.732(5) and

2.816(5) �A. The adjacent two 2D layers are further packed into

a double layered structure by O–H/O hydrogen bonds (a)

between the two single layers. Free water molecules are located in

the channels of the bilayers by forming a (H2O)10 cluster (b) and

are hydrogen bonded (c) to uncoordinated carboxylate oxygen

atoms of the host network (Fig. 1c and 1d).

[Ni(O2N–BDC)(bps)(H2O)]2n (2). The fundamental building

unit of 2 contains two nickel ions, two O2N–BDC ligands, two

bps ligands and two coordinated water molecules. The two Ni

atoms lie on twofold axes. The Ni1 center is coordinated by four

oxygen atoms from two O2N–BDC ligands and two nitrogen

atoms from two bps ligands. Ni2 adopts a distorted octahedral

[N2O4] coordination environment, and is coordinated by two

carboxylic oxygen atoms from two O2N–BDC ligands, two water

molecules and two nitrogen atoms from two bps ligands, as

shown in Fig. 2a. The Ni–O and Ni–N bond lengths are in the

range of 2.0276(14)–2.2315(13) and 2.0369(15)–2.1286(16) �A,

respectively. Every bps ligand adopts a bis-monodentate

bridging mode and links neighboring Ni(II) ions, resulting in

a 1D chain. The adjacent chains are connected by O2N–BDC

Fig. 2 (a) Coordination environment of Ni (II) ion in 2. Symmetry codes: A, 1

� x, y, 1.5 � z. (b) and (c) View of a 2D layer in 2. (d) 3D supramolecular n

780 | CrystEngComm, 2009, 11, 777–783

ligands (Scheme 1c) to give a highly puckered 2D (4,4) sheet

(Fig. 2b and 2c). Adjacent 2D layers are connected by O/O

intermolecular contacts (O4/O5, 3.015 �A; O4/O6, 3.025 �A)

between the uncoordinated carboxylate oxygens and the NO2

groups of O2N–BDC ligands, leading to a 3D supramolecular

network structure (Fig. 2d).

[Ni(OH–BDC)(bps)2(H2O)]n (3). The asymmetric unit of 3

consists of one Ni(II) ion, one OH–BDC, one bridging bps, one

monodentate bps and one coordinated water molecule. Each

Ni(II) center is in a distorted octahedral geometry. The six atoms

coordinated to Ni(II) ion come from two nitrogen atoms of two

bridging bps, one nitrogen atom from a terminal bps, two oxygen

atoms from two OH–BDC and one water molecule. The Ni–O

and Ni–N bond lengths are in the range of 2.0324(14)–2.0772(14)

and 2.1081(18)–2.1313(18) �A, respectively. Each OH–BDC acts

as a bis-monodentate ligand to bridge adjacent Ni(II) ions to

form a uniform chain. These chains are further linked by

bridging bps ligands along the c axis to generate a 2D rectangular

(4,4) grid layer (Fig. 3b and 3c). The terminal bps ligands sticking

outward from neighboring layers are deeply interdigitated to give

the 3D architecture shown in Fig. 3d. As a result, there exist

C–H/p stacking interactions between C14 of the terminal bps

and the benzene ring of OH–BDC ligands, with a distance of

3.722 �A, which connect the adjoining sheets.

{[Ni2(H2O)(tbip)2(bps)2]}n (4). The crystal structure of 4

exhibits a 3D network consisting of dinuclear nickel [Ni2(H2O)]4+

.5� x, y, 2.5� z; B, 2� x, 2� y, 2� z; C,�0.5 + x, 2� y, 0.5 + z; D, 1.5

etwork structure of 2.

This journal is ª The Royal Society of Chemistry 2009

Fig. 3 (a) Coordination environment of Ni (II) ion in 3. The hydrogen atoms are omitted for clarity. A,�1 + x, y, z; B, 1� x,�0.5 + y, 0.5� z. (b) and

(c) 2D layer of 3. (d) A schematic presentation of the interdigitation. Different colors in (d) are used to distinguish the layers of different directions.

moieties, coordinated water molecule, tbip and bps. The two

Ni(II) ions in 4 are both six-coordinated with a distorted octa-

hedral geometry (Fig. 4a). The six atoms coordinated to each

Ni(II) ion come from two nitrogen atoms of two bps ligands and

three oxygen atoms from three separate tbip ligands, as well as

one bridging water molecule. The two carboxylic groups of both

unique tbip ligands exhibit two kinds of coordination modes,

namely monodentate and bis-monodentate coordinating modes

(Scheme 1e).

The water bridge in each [Ni2(H2O)]4+ dimer is reinforced by

two bis-monodentate carboxylate groups from two tbip ligands

which also bridge the metal atoms, and this dimer acts as the

node of the overall network. Each dimer is not only bonded to

the two tbip ligands that provide the bridging carboxylates, but

also to further tbip ligands which bond in a monodentate fashion

(one per metal) and four bps ligands (Fig. 4a). For the mono-

dentate carboxylates, however, the connectivity is further

reinforced by hydrogen bonding interactions between the water

ligands and the uncoordinated carboxylate oxygens. Each water

ligand makes two such interactions. Thus there are eight bridging

ligands radiating out from each dimer, however these connect to

only four other dimers. This is because dimers are connected by

pairs of ligands—one bps bridge and one tbip bridge (Fig. 4b).

An overall 4-connected net is formed which has the diamond

topology (Fig. 4c). A feature of the diamond net is the hexagonal

and square channels; in this structure the hexagonal channels

are quite crowded due to the ligand conformations, however

the square channels are more apparent (Figs. 4d, 4e). As for the

diamond net, adjoining square channels have alternating

chirality. (Each tbip and bps in 4 in double bridging mode to

connect two adjacent nickel(II) ions forming unique left-handed

This journal is ª The Royal Society of Chemistry 2009

and right-handed 1-D chains.) Although many helical coordi-

nation polymers have been reported in the literature,15 the single

helix chain constructed from an intersectant 1-D two-stranded

chains in compound 4 is rare (Fig. S1, ESI†). The effective free

volume of 4 was calculated by PLATON analysis as 36.6% of the

crystal volume (2229.3 out of the 6099.2 �A3 unit cell volume).

Magnetic properties

The magnetic susceptibilities, cM, of 4 were measured in the

2–300 K temperature range, and are shown as cMT and cM

versus T plots in Fig. 5. The cMT values increase with the

decrease of the temperature, typical of the presence of intra-

dimeric ferromagnetic interaction. The decrease of cMT may be

attributed to an intermolecular antiferromagnetic interaction

and/or the zero-field splitting term of the Ni(II) ion.16 The

experimental cMT value at 300 K is 2.43 cm3 K mol�1, which is

slightly larger than the spin-only value (2.0 cm3 K mol�1)

expected for two coupled high-spin Ni(II) ions. The temperature

dependence of the reciprocal susceptibilities (1/cM) obeys the

Curie–Weiss law above 50 K with q ¼ 8.45 K, C ¼ 2.65 and R ¼1.84 � 10�4. The positive q value supports the presence of overall

ferromagnetic interactions in 4.

According to the structure of 4, it could be presumed that the

main magnetic interactions between the Ni(II) centers might

happen between water bridged Ni(II) ions whereas the super-

echange interactions between Ni(II) ions through the tbip and bps

bridge can be ignored because of the long distances of Ni/Ni

separation. The magnetic susceptibility data were fitted assuming

that the water bridges between Ni (II) ions form a binuclear

unit with exchange constant J and then tbip and bps connect the

CrystEngComm, 2009, 11, 777–783 | 781

Fig. 4 (a) Coordination environment of Ni(II) ion in 4. Symmetry codes: A, x, �y, 0.5 + z; B, 0.5 + x, 0.5� y, 0.5 + z; C, x, �y, �0.5 + z; D, �0.5 + x,

0.5� y,�0.5 + z. (b) The connecting of each cluster to four others via the eight bridging ligands. For clarity, tertiary butyl groups are omitted for clarity

and only water protons are shown, and striped bonds represent hydrogen bonds. (c) The overall diamond topology; one of the characteristic adamantane

cavities is highlighted in black. Node represent the Ni dimers. (d) The framework viewed along the [100] direction, showing two types of helical channels

which are alternately arranged along the a axis (L: left-handed helical channel, R: right-handed helical channel). (e) Space-filling view of the 3D structure

of 4 along the a axis.

Fig. 5 Temperature dependence of cMT and cM for 4. Open points are

the experimental data, and the solid line represents the best fit.

binuclear units to form a 3D network with an exchange constant

zJ0. The susceptibility data were thus approximately analyzed by

an isotropic dimer mode of spin S ¼ 1.17

782 | CrystEngComm, 2009, 11, 777–783

The least-squares analysis of magnetic susceptibilities data led

to J¼ 0.56 cm�1, g¼ 2.38, zJ0 ¼ �0.07 cm�1 and R¼ 5.12� 10�4

(the agreement factor defined as R ¼P

[(cM)obs � (cM)calc]2/

P[(cM)obs]

2. The J value indicates a weakly ferromagnetic

interaction between the two Ni(II) ions bridged by a water

molecule. The smaller negative zJ0 value can be attributed to

a very weak antiferromagnetic interaction between Ni(II) ions

through tbip and bps bridges.

Conclusion

Four nickel coordination polymers with three diverse 2D layer

structures and a 3D diamond network assembled from 4,40-

dipyridyl sulfide and isop as well as its three dicarboxylate

derivatives have been prepared and structurally characterized.

Structural comparisons indicate that the coexistent groups of

organic ligands on the dicarboxylate ligand play an important

role in governing the structures of 1–4. In addition, weak

hydrogen-bonding and intermolecular C–H/p stacking inter-

actions play a crucial role in linking the low-dimensional

networks into higher-dimensional supramolecular structures.

The successful preparation of the four different complexes

provides a valuable approach for the construction of other

This journal is ª The Royal Society of Chemistry 2009

coordination polymers with diverse structures via the introduc-

tion of various uncoordinated substituted groups on the bridging

ligands. Variable-temperature magnetic susceptibility measure-

ments revealed the existence of weak ferromagnetic interactions

between the nickel centers in 4.

Acknowledgements

This work was supported by the Natural Science Foundation of

China (Nos. 20771054 and 20771090) and Henan tackle key

problem of science and technology (Nos. 072102270030 and

072102270034) and the Foundation of Education Committee of

the Henan province (Nos. 2006150017 and 2008A150018).

References

1 (a) S. R. Batten and R. Robson, Angew. Chem., Int. Ed., 1998, 37,1460; (b) Z. B. Han, X. N. Cheng and X. M. Chen, Cryst. GrowthDes., 2005, 5, 695; (c) J. J. Perry, G. J. McManus andM. J. Zaworotko, Chem. Commun., 2004, 2534; (d) J. P. Zhang,S. L. Zheng, X. C. Huang and X. M. Chen, Angew. Chem., Int. Ed.,2004, 43, 206; (e) S. Hu, J. C. Chen, M. L. Tong, B. Wang,Y. X. Yan and S. R. Batten, Angew. Chem., Int. Ed., 2005, 44,5471; (f) S. Kitagawa, S. Noro and T. Nakamura, Chem. Commun.,2006, 701.

2 (a) M. Eddaoudi, J. Kim, N. Rosi, D. Vodak, J. Wachter,M. O’Keeffe and O. M. Yaghi, Science, 2002, 295, 469; (b) L. Pan,K. M. Adams, H. E. Hernandez, X. T. Wang, C. Zheng, Y. Hattoriand K. Kaneko, J. Am. Chem. Soc., 2003, 125, 3062; (c)G. L. Zheng, J. F. Ma, Z. M. Su, L. K. Yan, J. Yang, Y. Y. Li andJ. F. Liu, Angew. Chem., Int. Ed., 2004, 43, 2409; (d) L. F. Ma,L. Y. Wang, X. K. Huo, Y. Y. Wang, Y. T. Fan, J. G. Wang andS. H. Chen, Cryst. Growth Des., 2008, 8, 620; (e) L. F. Ma,Y. Y. Wang, L. Y. Wang, J. Q. Liu, Y. P. Wu, J. G. Wang,Q. Z. Shi and S. M. Peng, Eur. J. Inorg. Chem., 2008, 693.

3 (a) L. Y. Zhang, J. P. Zhang, Y. Y. Lin and X. M. Chen, Cryst.Growth Des., 2006, 6, 1684; (b) Q. Fang, G. Zhu, M. Xue, J. Sun,F. Sun and S. Qiu, Inorg. Chem., 2006, 45, 3582; (c) Z. B. Han,X. N. Cheng and X. M. Chen, Cryst. Growth Des., 2005, 5, 695; (d)G. F�erey, F. Millange, M. Morcrette, C. Serre, M. L. Doublet,J. M. Greneche and J. M. Tarascon, Angew. Chem., Int. Ed., 2007,46, 3259.

4 (a) G. L. Zheng, J. F. Ma, J. Yang, Y. Y. Li and X. R. Hao, Chem.–Eur. J., 2004, 10, 3761; (b) J. F. Ma, J. Yang, G. L. Zheng, L. Li andJ. F. Liu, Inorg. Chem., 2003, 42, 7531; (c) G. F�erey, F. Millange,M. Morcrette, C. Serre, M. L. Doublet, J. M. Greneche andJ. M. Tarascon, Angew. Chem., Int. Ed., 2007, 46, 3259; (d)X. M. Gao, D. S. Li, J. J. Wang, F. Fu, Y. P. Wu, H. M. Hu andJ. W. Wang, CrystEngComm, 2008, 10, 479; (e) K. Biradha,M. Sarkar and L. Rajput, Chem. Commun., 2006, 4169.

5 (a) D. M. J. Dobe, C. H. Benison, A. J. Blake, D. Fenske,M. S. Jackson, R. D. Kay, W. Li and M. Schr€oder, Angew. Chem.,Int. Ed., 1999, 38, 1915; (b) M. L. Tong, X. M. Chen, B. H. Ye andS. W. Ng, Inorg. Chem., 1998, 37, 2645; (c) J. W. Cheng,S. T. Zheng and G. Y. Yang, Dalton Trans., 2007, 4059; (d)Y. Q. uang, Z. L. Shen, T. Okamura, Y. Wang, X. F. Wang,W. Y. Sun, J. Q. Yu and N. Ueyama, Dalton Trans., 2008, 204; (e)H. Gudbjarlson, K. M. Poirier and M. J. Zaworotko, J. Am. Chem.

This journal is ª The Royal Society of Chemistry 2009

Soc., 1999, 121, 2599; (f) S. Noro, S. Kitagawa, M. Kondo andK. Seki, Angew. Chem., Int. Ed., 2000, 39, 2081.

6 (a) D. M. Shin, I. S. Lee, Y. K. Chung and M. S. Lah, Chem.Commun., 2003, 1036; (b) X. H. Bu, W. Chen, W. F. Hou, M. Du,R. H. Zhang and F. Brisse, Inorg. Chem., 2002, 41, 3477; (c)M. Du, X. J. Jiang and X. J. Zhao, Inorg. Chem., 2006, 45, 3998;(d) S. W. Keller and S. Lopez, Inorg. Chem., 1999, 38, 1883; (e)O. M. Yaghi, H. Li, C. David, D. Richardson and T. L. Groy, Acc.Chem. Res., 1998, 31, 474.

7 (a) X. M. Gao, D. S. Li, J. J. Wang, F. Fu, Y. P. Wu, H. M. Hu andJ. W. Wang, CrystEngComm, 2008, 10, 479; (b) B. R. Bhogala,S. Basavoju and A. Nangia, CrystEngComm, 2005, 7, 551; (c)W. J. Zhuang, X. J. Zheng, L. C. Li, D. Z. Liao, H. Ma andL. P. Jin, CrystEngComm, 2007, 9, 653; (d) D. R. Xiao,E. B. Wang, H. Y. An, Y. G. Li, Z. M. Su and C. Y. Sun, Chem.–Eur. J., 2006, 12, 6528; (e) M. Xue, G. S. Zhu, Y. J. Zhang,Q. R. Fang, I. J. Hewitt and S. L. Qiu, Cryst. Growth Des., 2008, 8,427; (f) W. G. Lu, L. Jiang and T. B. Lu, Cryst. Growth Des., 2008,8, 986.

8 (a) Z. M. Hao and X. M. Zhang, Cryst. Growth Des., 2007, 7, 64; (b)Y. Y. Niu, Z. J. Li, Y. L. Song, M. S. Tang, B. L. Wu and X. Q. Xin,J. Solid State Chem., 2006, 179, 4003; (c) O. S. Jung, S. H. Park,D. C. Kim and K. M. Kim, Inorg. Chem., 1998, 37, 610; (d) Z. Niand J. J. Vittal, Cryst. Growth Des., 2001, 1, 195.

9 (a) L. Pan, B. Parker, X. Y. Huang, D. H. Oison, J. Y. Lee and J. Li,J. Am. Chem. Soc., 2006, 128, 4180; (b) H. Chun, J. Am. Chem. Soc.,2008, 130, 800; (c) M. Du, X. J. Jiang and X. J. Zhao, Inorg. Chem.,2007, 46, 3984; (d) J. F. Ma, J. Y., G. L. Zheng, L. Li and J. F. Liu,Inorg. Chem., 2003, 42, 7531.

10 (a) Y. F. Zhou, F. L. Jiang, D. Q. Yuan, B. L. Wu, R. H. Wang,Z. Z. Lin and M. C. Hong, Angew. Chem., Int. Ed., 2004, 43, 5665;(b) S. C. Manna, E. Zangrando, J. Ribas and N. R. Chaudhuri,Eur. J. Inorg. Chem., 2008, 1400; (c) C. B. Ma, C. N. Chen,Q. T. Liu, D. Z. Liao and L. C. Li, Eur. J. Inorg. Chem., 2008,1865; (d) X. J. Li, R. Cao, D. F. Sun, W. H. Bi, Y. Q. Wang, X. L.and M. C. Hong, Cryst. Growth Des., 2004, 4, 775.

11 M. Eddaoudi, Jaheon Kim, J. B. Wachter, H. K. Chae, M. O’Keeffeand O. M. Yaghi, J. Am. Chem. Soc., 2001, 123, 4368.

12 S. A. Bourne, J. J. Lu, A. Mondal, B. Moulton and M. J. Zaworotko,Angew. Chem., Int. Ed., 2001, 40, 2111.

13 (a) X. J. Li, R. Cao, W. H. Bi, Y. Q. Wang, Y. L. Wang, X. L. andZ. G. Cao, Cryst. Growth Des., 2005, 5, 1651; (b)G. S. Papaefstathiou and L. R. MacGillivray, Coord. Chem. Rev.,2003, 246, 169.

14 (a) H. D. Flack and G. Bernardinelli, Acta Cryst., 1999, A55,908; (b) H. D. Flack and G. Bernardinelli, J. Appl.Crystallogr., 2000, 33, 1143; (c) P. G. Jones, Acta Crystallogr.,Sect. A, 1986, 42, 57.

15 (a) Y. Cui, S. J. Lee and W. B. Lin, J. Am. Chem. Soc., 2003, 125,6014; (b) R. H. Wang, Y. F. Zhou, Y. Q. Sun, D. Q. Yuan, L. Han,B. Y. Lou, B. L. Wu and M. C. Hong, Cryst. Growth Des., 2005, 5,251; (c) L. Han, M. C. Hong, R. H. Wang, J. H. Luo, Z. Z. Linand Daqiang Yuan, Chem. Commun., 2003, 2580; (d) X. Shi,G. S. Zhu, S. L. Qiu, K. L. Huang, J. H. Yu and R. R. Xu, Angew.Chem., Int. Ed., 2004, 43, 6482.

16 H. Z. Kou, S. Hishiya and O. Sato, Inorg. Chim. Acta, 2008, 361,2396.

17 (a) K. K. Nanda, A. W. Addison, N. Paterson, E. Sinn,L. K. Thompson and U. Sakaguchi, Inorg. Chem., 1998, 37, 1028;(b) A. P. Ginsberg, R. L. Martin, R. W. Brookes andR. C. Sherwood, Inorg. Chem., 1972, 11, 2884; (c) D. M. Dugganand D. N. Hendrickson, Inorg. Chem., 1974, 13, 2929.

CrystEngComm, 2009, 11, 777–783 | 783