Sei sulla pagina 1di 14

Volume 18 Number 8 28 February 2016 Pages 12531464

CrystEngComm
www.rsc.org/crystengcomm

PAPER
Javier Cepeda, Antonio Rodrguez-Diguez et al.
Controlling interpenetration for tuning porosity and luminescence
properties of flexible MOFs based on biphenyl-4,4-dicarboxylic acid

CrystEngComm
PAPER

Cite this: CrystEngComm, 2016, 18,


1282

Controlling interpenetration for tuning porosity


and luminescence properties of flexible MOFs
based on biphenyl-4,4-dicarboxylic acid
Beln Fernndez,a Garikoitz Beobide,b Ignacio Snchez,a Francisco CarrascoMarn,a Jos M. Seco,c Antonio J. Calahorro,a Javier Cepeda*cd
and Antonio Rodrguez-Diguez*a
Four new compounds based on zincII) or cadmiumII) metal ions and elongated dicarboxylate and
bipyridine ligands, namely, {[Cd34-bpdc)3H2O)2]DMF}n (1), {[Zn34-bpdc)3-bpdb)]5DMF}n (2), {[Zn24bpdc)2-bpdb)]7DMF}n (3), and {[Zn44-bpdc)3DMF)4-O)H2O)]7DMF3H2O}n (4), (where bpdc =
biphenyl-4,4-dicarboxylate, bpdb = 1,4-bis4-pyridyl)-2,3-diaza-1,3-butadiene, DMF = dimethylformamide)
have been synthesised under solvothermal conditions and structurally characterised by single crystal X-ray
diffraction. The crystal structures range from 2D (in 1) to 3D (2, 3, and 4) systems according to the coordination mode acquired by the bpdc ligand and the presence of an ancillary linker. Compound 1 consists of
stacked Cd-bpdc neutral layers containing isolated small voids. The coordination of the bpdb ligand (2 and

Received 18th October 2015,


Accepted 11th January 2016

3) or the formation of a tetrahedral Zn4O cluster (in 4) generates highly open 3D architectures that share
the structural feature of being doubly interpenetrated. A careful computational analysis on the crystal structures permits unravelling their void systems. Moreover, characterising the photoluminescence emission of

DOI: 10.1039/c5ce02036k

the compounds at variable excitation wavelengths provides an opportunity to couple the luminescence response with their porosity, which could signify the potential utility of these materials as photofluorescent

www.rsc.org/crystengcomm

sensors for small adsorbates.

1. Introduction
One major point concerning the sustainable development of
our society deals with the achievement of porous adsorbents
with an improved performance for environmentally friendly
and economically favourable separation, capture, and storage
of small gas molecules at an industrial level.1 In this sense,
extensive research has been performed to produce selective
adsorbents for purifying H2 from CO2 in the syngas resulting
from a gasification process.2 The efforts carried out over the
last two decades have mainly focused on the design of new
a

Departamento de Qumica Inorgnica, Universidad de Granada, 18071,


Granada, Spain. E-mail: antonio5@ugr.es; Tel: +34958240442
b
Departamento de Qumica Inorgnica, Facultad de Ciencia y Tecnologa,
Universidad del Pas Vasco/Euskal Herriko Unibertsitatea, UPV/EHU, Apdo. 644,
48080, Bilbao, Spain
c
Departamento de Qumica Aplicada, Facultad de Qumica, Universidad del Pas
Vasco/Euskal Herriko Unibertsitatea, UPV/EHU, 20018, San Sebastin, Spain.
E-mail: javier.cepeda@ehu.eus
d
Departamento de Ciencia y Tecnologa de Polmeros, Facultad de Qumica,
Universidad del Pas Vasco/Euskal Herriko Unibertsitatea, UPV/EHU, 20018, San
Sebastin, Spain
Electronic supplementary information (ESI) available: CCDC 14316981431701
contain the supplementary crystallographic data for this paper. For ESI and crystallographic data in CIF or other electronic format see DOI: 10.1039/c5ce02036k

1282 | CrystEngComm, 2016, 18, 12821294

metalorganic frameworks (MOFs), also called porous coordination polymers (PCPs), due to their structural diversity as a
consequence of their modular nature as well as their functional properties, which span diverse fields of application besides gas adsorption.3 In general, to explore the optimal structure for storage or separation performance, modulation of the
pore size/type has been considered as the best choice of a rational design.4 On the one hand, the modulation may be directed from a synthetic point of view, where the alteration of
the chemical composition, functionality, and molecular dimensions of the framework is addressed for a particular underlying topology.5 The large family of highly porous frameworks synthesised from the (3,24)-connected network
possessing the rht topology6 or that consisting of
4,8-connected networks with the scu topology are good examples for precise pore design and control.7 On the other hand,
MOFs with interpenetrated structures have demonstrated
great potential, since their pore apertures can be tuned to the
size of the gas molecules.8 Although this phenomenon has
been considered as a drawback for a long time because of the
associated decrease in the available void space and many
strategies were developed to avoid it,2e,9 the adsorption of
small molecules (e.g. H2, CO2, CH4) can benefit from interpenetration.10 This is because the number of adsorption sites

This journal is The Royal Society of Chemistry 2016

CrystEngComm

and their selectivity towards the target molecules is increased.


For instance, Chen et al. showed how doubly and triply interpenetrated Cu-MOFs had a significant uptake of CO2 at 195 K
and a very low uptake of N2 at 77 K.11 Moreover, the interpenetrated MOF-508b also exhibited an unprecedented capacity to separate and remove CO2 from binary CO2/N2 and CO2/
CH4 as well as ternary CO2/N2/CH4 mixtures.12
Notwithstanding the potential of interpenetrated frameworks for capture of small molecules, they also provide excellent case studies for the investigation of structural dynamism
given that they are inclined to transform in response to a critical amount or specific characteristics of guest molecules.13
This fact tends to afford isotherms with unusual shapes that
offer outstanding possibilities for the design of good adsorbents for a pressureswing process, in which not only a
high selectivity but also a high deliverable capacity are mandatory.14 Dynamic effects can arise either locally from flexible
ligands and/or flexibility of the coordination geometry of the
metal ions,15 or from the cooperative movement of the framework, for instance the sliding of interpenetrated networks
during guest accommodation,16 although they can also occur
simultaneously and contribute to one another.
Taking these considerations into account, we selected
biphenyl-4,4-dicarboxylate (bpdc) and 1,4-bis4-pyridyl)-2,3diaza-1,3-butadiene (bpdb) as linkers due to their high potential to build open architectures as a result of the length of
the spacer and the opposed disposition of the donor sets
while they also introduce certain flexibility.17 Additionally,
the first one permits diversified coordination modes (monodentate, bis-chelating interdinuclear mode, bis-bridging
intradinuclear mode and bis--O,O--O,O bridging/chelating
intradinuclear) of carboxylate groups, whereas the second
one imposes a predominant coordination pattern (Scheme 1).
Therefore, their linkage to metal atoms with a high coordination plasticity, such as cadmiumII) and zincII), opens up the
way to framework dynamics. As a result, four new compounds with intriguing frameworks have been synthesised
and characterised. Most of them possess open architectures
despite the occurrence of interpenetration, one of which exhibits a stepped adsorption of N2. Photoluminescence measurements have been performed in order to verify the ligand
centred emission or the ligand-to-metal charge transfer
mechanism associated with d10 transition metal based
MOFs.18 Moreover, the emission of the materials in response
to different excitation wavelengths has been analysed in order to study their potential utility as sensors for gas
adsorbates.

2. Experimental section
All reagents and solvents were commercially available and
used directly without any further purification.
2.1. Synthesis of compounds
2.1.1. Synthesis of {[Cd34-bpdc)3H2O)2]DMF}n (1). 2 mL
of a DMF solution containing 0.05 mmol of CdNO3)24H2O

This journal is The Royal Society of Chemistry 2016

Paper

Scheme 1 Coordination modes of the ligands in compounds 14: (a)


4-bpdc-2O,O:O:2O,O:O; (b) 4-bpdc-O:O:O:O; (c)
-bpdb-N:N.

(0.015 g) were added to another solution (2 mL of DMF) of


0.05 mmol of biphenyl-4,4-dicarboxylic acid (H2bpdc). The reaction mixture was heated in a closed vial at 95 C in oven
for 12 h. Single crystals of compound 1 were obtained. Yield:
75% based on Cd. Anal. calc. for C45H35Cd3NO15 (%): C,
46.31; H, 3.02; Cd, 28.90; N, 1.20. Found: C, 46.07; H, 3.25;
Cd, 29.02; N, 1.38.
2.1.2. Synthesis of {[Zn34-bpdc)3-bpdb)]5DMF}n (2). A
DMF solution (2 mL) containing 0.0261 g of ZnNO3)24H2O
(0.1 mmol) was added to 1 mL of a DMF solution of H2bpdc
(0.0242 g, 0.1 mmol). The resulting solution was sonicated
for 10 minutes, and then 0.0105 g of 1,4-bis4-pyridyl)-2,3diaza-1,3-butadiene (bpdb) (0.057 mmol) dissolved in 2 mL of
DMF was added. The reaction mixture was heated at 95 C
for 24 h and left to cool down to room temperature. Light yellow single crystals of compound 2 were observed. Yield: 80%
based on Zn. Anal. calc. for C69H69N9O17Zn3 (%): C, 55.53; H,
4.66; N, 8.45; Zn, 13.14. Found: C, 55.35; H, 4.43; N, 8.59; Zn,
13.35.
2.1.3. Synthesis of {[Zn24-bpdc)2-bpdb)]7DMF}n (3).
Light yellow single crystals of compound 3 were formed following the same procedure detailed for compound 2 but
heating the reaction mixture for 48 h at 95 C. Yield: 70%
based on Zn. Anal. calc. for C61H75N11O15Zn2 (%): C, 54.96;
H, 5.67; N, 11.56; Zn, 9.81. Found: C, 55.08; H, 5.55; N, 11.82;
Zn, 9.76.
2.1.4. Synthesis of {[Zn44-bpdc)3DMF)4-O)H2O)]
7DMF3H2O}n (4). Colourless single crystals of compound 4
were obtained by a similar procedure to that described for 1,
although zinc nitrate tetrahydrate was employed instead of

CrystEngComm, 2016, 18, 12821294 | 1283

Paper

CrystEngComm

cadmium nitrate to prepare the reaction mixture and it was


heated up to 95 C for 72 h. Yield: 65% based on Zn. Anal.
calc. for C66H88N8O25Zn4 (%): C, 47.90; H, 5.36; N, 6.77; Zn,
15.80. Found: C, 47.98; H, 5.25; N, 6.82; Zn, 15.56.

2.2. Single-crystal X-ray diffraction


X-ray data collection of suitable single crystals were done at
100(2) K on a Bruker VENTURE area detector equipped with
graphite monochromated Mo-K radiation ( = 0.71073 ) by
applying the -scan method. Data reduction was performed
with the APEX2 (ref. 19) software and data were corrected for
absorption using SADABS.20 Crystal structures were solved by
direct methods using the SIR97 program21 and refined by
full-matrix least-squares on F 2 including all reflections using
anisotropic displacement parameters by means of the WINGX
crystallographic package.22 All hydrogen atoms were located
in difference Fourier maps and included as fixed contributions riding on attached atoms with isotropic thermal displacement parameters 1.2 times or 1.5 times those of their
parent atoms for the organic ligands and the water molecules, respectively. During the refinement of all compounds,
the electron density at the voids was subtracted from the reflection data by the SQUEEZE procedure as implemented in
the PLATON program23 due to the presence of disordered solvent molecules. Details of the structure determination and refinement of all compounds are summarized in Table 1. X-ray
powder diffraction (XRPD) patterns were collected on a Phillips X'PERT powder diffractometer with Cu-K radiation ( =
1.5418 ) over the range 5 < 2 < 50 with a step size of
0.02 and an acquisition time of 2.5 s per step at 25 C (see
the ESI). Indexation of the diffraction profiles was
conducted by means of the FULLPROF program (patternmatching analysis)24 on the basis of the space group and the
cell parameters measured on single crystal X-ray diffraction.

2.3. Physical measurements


Elemental analyses (C, H, and N) were performed on a Fisons
Instruments EA-1008 analyser. The metal content was determined by inductively coupled plasma (ICP-AES) analysis with
a Horiba Yobin Yvon Activa spectrometer. The IR spectra
(KBr pellets) were recorded on a ThermoNicolet IR 200
spectrometer in the 4000400 cm1 spectral region. All equipment is sited at the Centre of Scientific Instrumentation of
the University of Granada. Thermogravimetric and Differential Thermal Analyses (TG/DTA) were performed on a TA Instruments SDT 2960 thermal analyser in a synthetic air atmosphere (79% N2/21% O2) with a heating rate of 5 C min1
(see Tables S1S4 in the ESI).
2.4. Gas adsorption isotherms
Volumetric N2 gas adsorption isotherms were obtained at 77
K using a Quantachrome Autosorb-1 equipment. Approximately 50 mg of the corresponding solid product was transferred to a pre-weighed sample tube and evacuated under dynamic vacuum, 107 torr, at 383 K on the gas adsorption
apparatus. The sample tube was reweighed to obtain a consistent mass for the degassed modified product. Helium was
used for dead volume determination. All gases used were of
99.999% purity.
2.5. Photoluminescence measurements
A Varian Cary-Eclipse Fluorescence Spectrofluorimeter was
used to obtain the fluorescence spectra. The spectrofluorimeter was equipped with a xenon discharge lamp (peak power
equivalent to 75 kW), CzernyTurner monochromators, and
an R-928 photomultiplier tube which is red sensitive (even
900 nm) with manual or automatic voltage controlled using
the Cary Eclipse software for a Windows 95/98/NT system.
The photomultiplier detector voltage was 700 V and the

Table 1 Crystallographic data and structure refinement details of all compounds

Compound

C69H69N9O17Zn3
C61H75N11O15Zn2
Chem. form.
C45H35Cd3NO15
Form. weight
1166.99
1492.51
1333.09
Cryst. system
Trigonal
Monoclinic
Triclinic

Space group
R3
C2/c
P1
a ()
14.070(6)
47.928(4)
15.174(2)
b ()
14.070(6)
13.828(1)
15.185(2)
c ()
19.864(8)
26.187(4)
18.174(2)
()
90
90
81.094(3)
()
90
114.491(2)
84.734(3)
()
120
90
78.952(3)
3406(2)
15794(2)
4052.2(7)
V (3)
Z
3
8
2
1.006
0.927
0.998
GOFa
0.0669
0.0865
0.0780
Rint
0.0547/0.1396
0.0538/0.1338
0.0598/0.1542
R1b/wR 2c [I > 2(I)]
0.1186/0.1548
0.0877/0.1443
0.0814/0.1604
R1b/wR 2c (all data)
P
P
P
P
a
1/2 b
c
2
2 2
2
2 2 P
S = [ wF0 Fc ) /(Nobs Nparam)] . R1 = F0| |Fc/ |F0|. wR2 = [ wF0 Fc ) / wF02]1/2; w = 1/[ 2F02) + (aP)2
(maxF02,0) + 2Fc2)/3 with a = 0.0749 (1), 0.0678 (2), 0.0962 (3), 0.1209 (4), and b = 29.2136 (4).

1284 | CrystEngComm, 2016, 18, 12821294

4
C66H88N8O25Zn4
1655.00
Orthorhombic
Pnnm
17.256(4)
23.227(6)
25.426(6)
90
90
90
10191(4)
4
1.090
0.0575
0.0711/0.1975
0.0747/0.2018
+ bP] where P =

This journal is The Royal Society of Chemistry 2016

CrystEngComm

instrument excitation and emission slits were set at 5 and 5


nm, respectively.

3. Results and discussion


3.1. Comments on the MII/bpdc and MII/bpdc/bpdb systems
The X-ray diffraction analysis of the four crystallised metal
organic architectures allows their classification according to
the title secondary or ternary systems. The structural diversity
found in these systems is explained in terms of the different
factors imposed by the synthetic conditions, which affect the
coordination mode of the bpdc ligands (see Scheme 1). To
start with, the main difference between the two compounds
obtained in the MII/bpdc system (1 and 4), apart from the
metal atom, originates from the formation of the oxide anion
under mild-medium solvothermal conditions of the solution
containing the reagents when long reaction times are
employed.25 In this way, the absence of the latter anion favours the generation of trinuclear secondary building units
(SBUs) that require bpdc ligands to acquire coordination
mode a, which leads to 2D layers of compound 1 given the
water molecules capping the SBUs. In contrast, the occurrence of Zn4O clusters limits the coordination pattern of
bpdc to mode b, thus extending the polymerization of the
backbone along three perpendicular directions giving rise to
the cubic 3D porous framework of 4. To that end, one of the
tetrahedral vertexes is blocked by the coordination of solvent
molecules.
The addition of a bpdb co-ligand to the reaction media,
which has a preferred coordination mode (c in Scheme 1)
and a marked trend to act as a pillar, gives two types of 3D
frameworks shaped by the topology of neutral Zn-bpdc layers
(compounds 2 and 3). The overall architecture of 2 is the consequence of the junction of the layers present in compound 1
owing to the replacement of terminal water molecules by linear bpdb bridges. On the other hand, applying longer reaction times favours the occurrence of more stable
[Zn2O2CR)4] paddle-wheel entities in which the bpdc ligand
displays coordination mode b (see Scheme 1), as a consequence of a mechanism involving dissolution and
recrystallisation processes from compound 2.26 Thus, the 3D
primitive cubic type (-Po) structure of compound 3 is
obtained. This fact seems to indicate that compounds 2 and
3 are the kinetic and thermodynamic products of the Zn/
bpdc/bpdb system, respectively.
On another level, it is remarkable to note that the 3D
structures of compounds 2, 3, and 4 crystallise as doubly
interpenetrated frameworks given the huge size of voids potentially achieved according to the length of the bridging ligands. This fact, although drastically reduces the accessible
surface area of the resulting frameworks, gives the opportunity to control the size of the pores and to gain access to pore
structures with narrower sections. The repetitive 2-fold entanglement, far from being a mere coincidence, is well explained
on the basis of the pore size distribution computationally calculated (see section 3.3).

This journal is The Royal Society of Chemistry 2016

Paper

3.2. Description of the structures


and
Compound 1 crystallises in the trigonal space group R3
results from the pilling of 2D sheets generated by trinuclear
SBUs. These centrosymmetric SBUs lie on a three-fold axis of
rotation and contain two different cadmiumII) atoms (Fig. 1).
Two Cd1 atoms occupy the edges of the unit and render a
distorted capped octahedron (SCOC = 3.71)27 formed by an O7
donor set established by six carboxylate oxygen atoms of
three chelating symmetry related bpdc ligands and the terminal water molecule that prevents the SBUs from further polymerization (Table 2).
Cd2 lies on the centre of the unit and exhibits an almost
ideal octahedral environment (SOC = 0.60) established by an
O6 donor set from six carboxylate oxygen atoms. It must be
noted that three Cd1Ocarb bonds, those established with the
bridging oxygen atoms, are remarkably longer than the
remaining coordination bond distances. Hence, all bpdc ligands exhibit the same coordination mode (a, see Scheme 1)
in such a way that three bpdc bridge each Cd1 atom with the
central Cd2. As a consequence, six dicarboxylic linkers arise
from the SBU in order to join it with six surrounding units.
The topological analysis of the resulting 2D network has been
performed with the TOPOS program,28 which indicates that
the network belongs to the Shubnikov hxl type, which is also
described by the (364653) point symbol (Fig. 2). Within the
sheets, the metal atoms of the trimeric SBUs are spread out
perpendicular to the mean plane whereas the bpdc ligands,
essentially planar, are arranged outwards it (forming an angle
of nearly 65 with regard to the aromatic rings), which affords some corrugation to the sheet. These sheets resemble
those found for the recently published [Cd34-azdc)3DMF)2]n
compound in which the azdc spacer establishes the same topological pattern and the coordination water molecules are
replaced by DMF.29 The sheets are then piled up on top of
one another in an ABC fashion in which the terminal water
molecules of one layer point towards the triangular rings of

Fig. 1 Trinuclear SBU of compound 1 showing the coordination


environments of the two crystallographically independent Cd atoms
(hydrogen atoms have been omitted for clarity). Symmetries: (i) y + 1,
x y, z. (ii) x + y + 1, x, z. (iii) x+ 4/3, y + 2/3, z 4/3. (iv) y + 1/3,
x + y +2/3, z 4/3. (v) x y + 1/3, x 1/3, z 4/3.

CrystEngComm, 2016, 18, 12821294 | 1285

Paper

CrystEngComm

Table 2 Selected bond lengths for compound 1a

Cd1O1A
Cd1O2A

2.263(4)
2.568(5)

Cd2O2A

2.264(4)

Fig. 3 Coordination environment of the metal atoms establishing the


secondary building unit in compound 2 (hydrogen atoms have been
omitted for clarity).

Fig. 2 (a) Junction of the SBUs giving rise to the 2D sheet. (b) Pilling
of the sheets (hydrogen atoms have been omitted for clarity).

the upper and lower sheets, which gives rise to an overall 3D


packing that encloses discrete voids filled with solvent molecules. A detailed analysis with PLATON software estimates
that these voids stand for 22% of the total volume of the cell
(Fig. S1).
The crystal structure of compound 2 consists of a 3D open
framework that can be considered as the consequence of the
junction of the isoreticular 2D sheets of compound 1 through
the bpdb pillaring linkers. The SBUs that build up the structure are also trinuclear units formed by three crystallographically independent zincII) atoms (Fig. 3). Zn1 atom occupies the centre of the building unit and is surrounded by a
slightly distorted octahedral coordination environment (SOC =
0.23) established by an O6 donor set from six carboxylate oxygen atoms belonging to four bpdc ligands. The difference regarding the SBUs of compound 1 comes from Zn2 and Zn3
occupying the edges, which show NO4 donor sets resembling
square pyramidal (SSPY = 1.87) and trigonal bipyramidal
(STBPY = 2.83) coordination polyhedra, respectively (Table 3).
As a result of the coordination number decrease in the
outer metal centres, only two of the six bpdc ligands that

1286 | CrystEngComm, 2016, 18, 12821294

emerge from the trinuclear unit maintain coordination mode


a (A ligand, see Scheme 1) in order to join the neighbouring
units, whereas the remaining four follow coordination mode
b (B ligand). Given the fact that each SBU links to other six
through these bridges, the resulting 2D layer keeps the same
topology of the previous compound. It is worth noting that A
ligands preserve the out-of-plane arrangement with regard to
the mean plane of the layer (ca. 65 between the aromatic
rings and the layer), whereas B ligands spread completely
perpendicular to it (Fig. 3). Hereafter, twisted bpdb pillaring
ligands (ca. 29 between their aromatic rings) pierce adjacent
layers through the triangular windows and link alternative
layers to one another leading to a doubly interpenetrated 3D
structure in which each net can be referred to as a hex hexagonal primitive network with the (36418536) point symbol
(Fig. 4). In spite of the drastic volume drop caused by the
interpenetration, the overall framework preserves high porosity consisting of 2D void systems that represent almost 49%
of the unit cell volume (Fig. S2).
The X-ray crystal structure analysis revealed that the secondary building unit of compound 3 is a paddle-wheel entity
(Fig. 5) in which two crystallographically independent zincII)
atoms are coordinated to the carboxylate oxygen atoms of
four bpdc ligands (coordination mode b in Scheme 1),
establishing an intradimeric ZnZn distance of around 2.9 .
The apical positions of the paddle-wheel entity are completed by the coordination of the pyridine nitrogen atoms of
the pillaring -bpdb linkers, rendering square pyramidal NO4
donor sets (SSPY of 0.24 and 0.23 for Zn1 and Zn2) (Table 4).
Each paddle-wheel unit is connected to four neighbouring
ones by means of the 4-bpdc linkers, which are slightly
twisted (angle between aromatic rings of ca. 34) giving rise
to 2D neutral [Zn24-bpdc)2] layers that show a distorted
squared grid of approximate dimensions of 15.1 15.1

This journal is The Royal Society of Chemistry 2016

CrystEngComm

Paper

Table 3 Selected bond lengths for compound 2a

Zn1O1A
Zn1O3A
Zn1O1B
Zn1O4Bi)
Zn1O2C(i)
Zn1O4C(ii)
a

2.044(2)
2.067(2)
2.197(2)
2.118(2)
2.045(2)
2.062(2)

Zn2O2A
Zn2O4A
Zn2O1B
Zn2O2B
Zn2N1D

1.947(2)
1.926(2)
2.101(2)
2.268(2)
2.037(3)

Zn3O3B
Zn3O4B
Zn3O1C
Zn3O3C(iii)
Zn3N14Div)

2.103(3)
2.223(2)
1.928(2)
1.943(2)
2.075(3)

Symmetries: (i) x, y + 1, z. (ii) x, y 1, z 1/2. (iii) x, y 2, z 1/2. (iv) x 1/2, y 1/2, z 1/2.

assuming the centroids of the SBUs. The layers are then assembled together by the coordination bpdb pillaring spacers
that arise from the building units almost perpendicularly,
thus leading to a highly open 3D pcu network with the (412
63) point symbol.
However, compound 3 avoids extremely large void space
by constructing a twofold interpenetrated framework. Even
so, this compound still exhibits somewhat large cavities that
are interconnected through a 3D system that is occupied by
solvent molecules and stands for 65.2% of the total unit cell
volume (Fig. 6 and Fig. S3 in the ESI). Although this sort of
backbone has been observed for other pillaring ligands,30
most of them do not exhibit such a high pore volume/cell volume ratio even if interpenetration of the framework is
prevented.
The crystal structure of compound 4 consists of a 3D open
framework built up from the linkage of Zn4O SBUs by means
of bpdc linkers. The asymmetric unit contains half a Zn4O
cage, two bpdc ligands, a disordered coordination DMF molecule, and a coordination water molecule in addition to disordered crystallisation DMF molecules, one of which has been

Fig. 4 (a) Molecular and schematic view of a single 3D network built


by the junction of layers. (b) Interpenetration of the two subnets in
compound 2.

This journal is The Royal Society of Chemistry 2016

located. Two of the three crystallographically independent


ZnII) atoms (Zn1 and Zn2) are tetrahedrally coordinated by
three carboxylate oxygen atoms and the central oxide anion
(ST = 0.28 and 0.34), whereas Zn3 shows a severely distorted
octahedral O6 coordination environment (SOC = 3.72) completed by the water and DMF molecules (Table 5). It is worth
mentioning that the fourth metal of the cluster is generated
by the mirror plane reflecting the Zn3 atom and that the
DMF molecule is affected by such symmetry operation, which
makes it to be disordered. Each Zn4O cluster is connected to
six 4-bpdc (coordination mode b in Scheme 1) spacers that
join it with six neighbouring clusters along a 3D pcu framework (Fig. 7).
The relatively high rotational freedom of the aromatic
rings of the spacers is reflected by their disordered dispositions, which places them rotated one another at equivalent
dispositions. As observed for preceding compounds (Fig. 8),
this network also allows the interpenetration of a second network, a fact that reduces the accessible volume in the overall
crystal structure although it has been estimated to account
for 66% of the unit cell (Fig. S4). The framework of 4 resembles that of previously published {[Zn4Obpdc)3DEF)2]7DMF
4H2O}n (IRMOF-9) compound,5a although the herein described compound contains different coordination solvent
molecules given the solvent mixture employed in the synthesis that is beneficial from the economic point of view.

Fig. 5 3D backbone of compound 3 showing the coordination


polyhedra (hydrogen atoms have been omitted for clarity).

CrystEngComm, 2016, 18, 12821294 | 1287

Paper

CrystEngComm

Table 4 Bond lengths of the coordination environments in compound


3a

Zn1O2A
Zn1O4A(i)
Zn1O2B
Zn1O4B
Zn1N1C
a

2.043(2)
2.046(3)
2.021(3)
2.035(3)
2.028(3)

Zn2O1A
Zn2O3A(i)
Zn2O1B
Zn2O3B
Zn2N14C

2.028(2)
2.043(3)
2.062(3)
2.042(3)
2.033(3)

Symmetry: (i) x, y + 1, z.

Fig. 7 Fragment of the framework showing the connection of the


Zn4O cluster to the neighbouring ones through the bpdc ligands and
coordination environments. Atoms drawn in lighter colours
correspond to disordered parts of the structure.

3.3. Pore structure analysis

Fig. 6 View of the 3D packing of compound 3 showing: (a) entangled


three-dimensional subnets and (b) 3D interconnected void system.

As previously observed, all crystal structures contain remarkable void percentages spanning from discrete cavities to
interconnected channels so, at a first glance, some of these
compounds could a priori behave as porous frameworks to
support the adsorption of small gas molecules. In fact, TG/
DTA analyses confirm that all structures remain stable upon
release of the solvent molecules (see Tables S1S4 in the
ESI). However, the pore structures have to fulfil some other
geometrical requirements so that they can permit the diffusion of adsorbate molecules along the system. With the aim
of better characterising the pore features of these compounds, we have analyzed their pore size distribution
through a Monte Carlo procedure implemented in a code developed by Herdes and Sarkisov (Fig. 9).31 This method explores the free volume of the framework with a probe of incremental size that allows the examination of all pores

Table 5 Selected bond lengths for compound 4a

Zn1O1A
Zn1O1A(i)
Zn1O1B
Zn1O1C

1.945(4)
1.945(4)
1.954(6)
1.930(5)

Zn2O3A
Zn2O3A(i)
Zn2O2B
Zn2O1C

1.934(5)
1.934(5)
1.957(7)
1.966(5)

Zn3O2A(i)
Zn3O4A
Zn3O3Bii)
Zn3O1C
Zn3O1D
Zn3O1W

2.002(4)
1.943(7)
1.937(6)
1.952(3)
2.314(12)
2.218(15)

Symmetries: (i) x, y, z. (ii) x 1, y, z.

1288 | CrystEngComm, 2016, 18, 12821294

This journal is The Royal Society of Chemistry 2016

CrystEngComm

Paper
Table 6 Geometrical analysis of the pore systems of compounds

Comp.a

Geom.
vol.b

He
vol.c

Surf.
aread

1
2
2-NoInt
3
3-NoInt
4
4-NoInt

0.188
0.617
1.719
1.048
2.593
1.082
2.463

0.126
0.567
1.673
1.037
2.524
0.997
2.332

178
1077
3876
2644
5139
2570
4153

Pore
diametere
Lim.

Max.

Dimens. f

1.23
4.56
8.95
6.70
11.40
7.87
11.7

5.67
8.71
11.58
9.19
15.10
11.30
17.8

0D
2D
3D
3D
3D
3D
3D

Analyzed compounds, considering the interpenetrated (bare


number) and non-interpenetrated (number-NoInt) frameworks.
b
Geometric volume (cm3 g1). c Accessible volume for He (cm3 g1).
d
Accessible specific surface area (m2 g1). e Limiting and maximum
pore diameters (). f Dimensionality of the pore structure for a N2
probe.

Fig. 8 Packing of compound 4 along the crystallographic c axis


showing the two interpenetrated frameworks.

contributing to the accessible volume, while it also provides


useful information, such as the specific surface area
(Table 6), which has been proven to correlate well with the
experimental data.32 Moreover, non-interpenetrated networks
have been also studied in order to estimate the largest porosity available for each system.

Fig. 9 Pore size distributions for interpenetrated (filled blue circles)


and non-interpenetrated (open red squares) compounds estimated
from X-ray structures.

This journal is The Royal Society of Chemistry 2016

In the light of the results, some interesting conclusions


can be drawn. First, the packing of the 2D layers in compound 1 confirms the segregation of the accessible volume in
the form of voids with dimensions within the 0.10.6 nm
range, which indicates that the dimensions based on van der
Waals radii overestimate those derived from the computational route (see the ESI). This fact prevents this compound
from being used for gas adsorption purposes. On the other
hand, it is worth noting the significant change in the accessible surface area between the empty frameworks of 2 and 3 in
spite of their structural resemblance (maximum pore sizes of
ca. 0.9 nm in both compounds). For comparison purposes,
the specific surface area of 3 is ca. 2.5 times larger than that
of 2. Although surprising at first sight, one has to take into
account that the 3D primitive cubic-type architectures built
up from paddle-wheel SBUs joined together through N-donor
pillars remain as one of the most common classes of
tailorable porous coordination polymers.33 In terms of porosity, intermediate to the previous compounds is the porous
structure of 4 that exhibits the biggest pores of all the
compounds.
Owing to the fact that interpenetration arises from the
presence of large free spaces in a single network, its control
and prediction remains a major challenge although some recent works have pointed out that factors such as synthetic
conditions,34 liquid phase epitaxy,35 the shape of the of the
SBU,36 and especially, the length of the ligand37 play a definite role. The occurrence of 2-fold interpenetrated frameworks of compounds 2, 3, and 4 may be related to a synthetic
and a ligand designed control. On the one hand, the solvent
used in the reaction mixture has demonstrated to be capable
of directing the synthesis towards catenated or non-catenated
isomers. For instance, it well known that the use of DMF or
DEF (diethylformamide) discerns between the catenated or
non-catenated isomers of the paddle-wheel shaped
[LCu2H2O)2] MOF.38 Moreover, Sun et al. successfully demonstrated that employing DMSO (dimethylsulfoxide) and DMA
(dimethylacetamide) renders PCN-6 (catenated) and PCN-6

CrystEngComm, 2016, 18, 12821294 | 1289

Paper

(non-catenated) compounds, respectively.39 Therefore, given


that DMF is a widely employed solvent that has been proven
to allow interpenetration in contrast to similar solvents, it
was chosen to promote interpenetrated structures. Regarding
the ligand design, a previous work with azdc (azobenzene4,4-dicarboxylate)
and
pbptz
(3,6-bis4-pyridyl)-1,2,4,5tetrazine) linkers depicts how the resulting frameworks exhibit 2-, 3-, and 4-fold entanglements.29 That structural
feature was satisfactorily explained on the basis of the correlation found between the maximum pore diameter and the
degree of interpenetration. In particular, it was observed that
the value for a given non-interpenetrated structure roughly
doubles that of the interpenetrated one, in such a way that the
degree of interpenetration acquired by the structure corresponds to that which leaves a pore diameter that better fits the
latter 2 : 1 ratio. Thus, we have made use of shorter bpdc and
bpdb ligands in order to slightly minimise the void volume so
that the degree of interpenetration is decreased and uniform
(as it is the case of compounds 2, 3, and 4). In fact, it can be
observed these doubly interpenetrated frameworks fulfil the
2 : 1 ratio of non-interpenetrated : interpenetrated structures.

3.4. Adsorption properties


As previously stated, compounds 2, 3, and 4 show open 3D
architectures containing relatively large interconnected pore
structures despite their doubly interpenetrated frameworks
(see Table 6), which seems to indicate that these microporous
compounds could support the adsorption of gas molecules.
However, one must take into account that the crystallographic porosity assumes a rigid framework that remains almost invariable throughout outgassing and adsorption/desorption processes, which is not fulfilled for structures in
which dynamic features involving framework transformations
are expected, as it is the case (flexible ligands, interpenetration, and so on). A recent review on flexible MOFs by
Schneemann et al.40 considers four flexural effects that modify the gas adsorption behaviour: i) breathing of the structure
implying a substantial displacement of the atoms and a subsequent change in the unit cell volume, ii) swelling
characterised by an (gradual) enlargement of the unit cell volume without a change in the cell shape and space group, iii)
linker rotation involving a spatial alignment of a ligand by
turning around a rotational axis, and iv) subnetwork displacement in interpenetrated frameworks (Scheme 2).
In principle, the last situation may be considered as the
most influencing one given the absence of strong supramolecular interactions among the subnets, in such a way that
the removal of the solvent molecules allows the two subnets
to drift, relocate, or shift to each other. As a consequence,
the adsorption behaviour is dominated by the gate-opening
effect that makes the porous structure accessible only for certain species under certain conditions.15,41 Accordingly, although several attempts were performed on polycrystalline
samples of compounds 3 and 4, N2 adsorption characterisation at 77 K revealed negligible uptake in both cases. In

1290 | CrystEngComm, 2016, 18, 12821294

CrystEngComm

Scheme 2 Schematic depiction of the gate-opening transformation


due to the displacement of subnets forming compound 2.

contrast, compound 2 shows a less common type V N2 adsorption isotherm in which a flat curve indicative of no adsorption is observed in the lower pressure, followed by an
abrupt increase at a relative pressure (P/P0) close to 0.5 that
reaches a value of ca. 85 cm3STP) g1 at P/P0 1. Interestingly, the desorption curve exhibits a monotonic decrease until low relative pressure (ca. 0.1) where it shows a sharp decay, defining a wide hysteresis loop (Fig. 10).
Despite the fact that the shape of the adsorption curve is
characteristic of mesoporous materials,42 herein the onset
pressure can be related to a gate-opening caused by the increase of the adsorbate pressure. On the contrary, the desorption branch agrees with the expected behaviour of a microporous structure. In this kind of compounds once the
desorption process reaches the onset pressure (below 0.1
0.15), the gate-closing takes place as a result of the displacement of the subnets. This behaviour leads to reproducible
hysteresis loops. With the aim of achieving a better estimate
between the adsorption behaviour and the pore system, the
DFT method was applied on experimental isotherms since it
provides more accurate results for MOFs than other standard
methods.43 The results for the adsorption branch show a
mesoporous pore distribution (ca. 5.8 nm) which is inconsistent with the crystal structure and can be attributed to an artefact caused by the dynamic behaviour of the structure. Nevertheless, the analysis of the desorption curve yields a pore
size distribution well within the microporous region (centred
at 1.3 nm) that agrees fairly well with that expected from the
crystal structure.

3.5. Photoluminescence properties


Metalorganic frameworks based on d10 metal centres and organic ligands containing conjugated systems, such as herein
presented dicarboxylic bpdc and N-donor bpdb, are considered as promising candidates for developing hybrid photoactive materials with potential application as light emitting
diodes (LEDs) or chemical sensors.18 Therefore, the room

This journal is The Royal Society of Chemistry 2016

CrystEngComm

Paper

centred at 418 and 470 nm under excitation at 350 nm,


respectively.44
Interestingly, compounds 13 show a tunable emissive response according to different excitation wavelengths (in particular, 257, 344, and 385 nm) (Fig. 11). When excited at 257
nm, compound 1 shows three main emissions located at 420,
485, and 529 nm, in addition to other smaller peaks. However, all these emissions are almost negligible under excitation at 344 nm, whereas they are again observed, slightly
shifted with regard to the initial spectrum, when the sample
was excited at 385 nm. Based on these results, the emission
of compound 1 can be ascribed to * n or even *
transitions in close packing states, which generates the observed weak red shift.45 Compounds 2 and 3 exhibit closely
related emission spectra in good agreement with their almost
identical composition. Under excitation at 257 nm, both samples emit the same structureless peaks centred at 420, 485,
and 529 nm, although many other smaller emissions also
possess a noticeable intensity, which may correspond to
LMCT of bpdb ligands or metal-perturbed intraligand charge
transfer of the bpdc ligand. The few differences found in the
spectra of both compounds are attributed to the presence of
two coordination modes in compound 2. Regarding the evolution of the emission spectra with the excitation wavelength,
these compounds show the same trend as compound 1, although weak transitions are still observable under excitation
at 344 nm. The emission spectrum of compound 4 contains a
broad peak at 460 nm and two weak shoulder bands at 420
and 485 nm (exc = 257 nm), respectively. All these bands are
consistent with the emission wavelength of the free H2bpdc
ligand. The excitation of the sample at higher wavelengths
provides a slight shift of the main peak while the shoulders
remain unaltered.

Fig. 10 (a) Adsorption and desorption N2 isotherms at 77 K for


compound 2. (b) DFT computed pore size distribution on both
branches.

temperature solid state emission spectra of all as-synthesised


compounds have been recorded. As previously reported, free
H2bpdc and bpdb ligands exhibit broad emission bands

This journal is The Royal Society of Chemistry 2016

Fig. 11 Emission spectra of compounds: (a) 1, (b) 2, (c) 3, and (d) 4 at


different excitation wavelengths.

CrystEngComm, 2016, 18, 12821294 | 1291

Paper

CrystEngComm

well as their unlikely experimental adsorption performance


may be explained assuming that these frameworks are prone
to undergo remarkable transformations that have an important impact on their pore structures. This fact generates compound 2 to show a sort of gate-opening behaviour upon N2
adsorption and compounds 3 and 4 to exhibit a negligible
uptake. With regard to the luminescence properties, these
compounds provide an interesting tunable ligand-centred or
ligand-to-metal charge transfer based emission in response
to variable excitation wavelengths. All in all, coupling of the
luminescence response and potential porosity of the compounds could signify their utility as photofluorescent sensors
for small adsorbates.

Acknowledgements

Fig. 12 Wavelength dependent emission colours of compounds 14


according to the CIE 1931 chromaticity coordinate diagram.

In order to provide deeper insights into the emission colour in response to the excitation wavelength, all spectra have
been transformed into chromaticity coordinates and located
in the CIE 1931 diagram (Fig. 12).46 It can be clearly seen that
while compounds 1 and 4 give a somewhat stable blue emission, compounds 2 and 3 are notoriously influenced by the
excitation wavelength since the colour oscillates from a deep
blue to a clear blue emission. This fact endows these compounds with potential performance as photoluminescent
sensors.

Conclusions
Four new compounds based on d10 metal atoms and
biphenyl-4,4-dicarboxylic (bpdc) and 1,4-bis4-pyridyl)-2,3diaza-1,3-butadiene (bpdb) ligands have been obtained under
solvothermal conditions. The crystal structures range from
2D layers to open 3D frameworks according to the coordination mode established by bpdc, which mainly depend on the
reaction conditions, but also on the addition of a bpdb coligand to the synthesis. In this sense, control of the synthetic
conditions (reaction time of the solvothermal treatment) allows tuning the resulting SBUs which gives access to structural variability. A common structural feature of the open
three-dimensional architectures is their crystallisation as
doubly interpenetrated networks, which gives pore structures
characterised by more uniform distributions and narrower
sections compared to potentially achievable noninterpenetrated structures. The repetitive 2-fold entanglement
obeys the relationship established by computed pore size distributions of non-interpenetrated and interpenetrated frameworks. Physical instability upon outgassing of compounds as

1292 | CrystEngComm, 2016, 18, 12821294

This work was supported by the MEC of Spain (Project


CTQ2011-24478) and the Junta de Andaluca (FQM-1484 and
FQM-195). Javier Cepeda acknowledges Universidad del Pas
Vasco/Euskal Herriko Unibertsitatea for his postdoctoral fellowship and Jesus M. Ugalde for his guidance and support
along the postdoctoral period.

Notes and references


1 (a) B. Chen, S. Ma, F. Zapata, F. R. Fronczek, E. B.
Lobkovsky and H.-C. Zhou, Inorg. Chem., 2007, 46, 1233; (b)
H. B. T. Jeazet, C. Staudt and C. Janiak, Chem. Commun.,
2012, 48, 2140; (c) J.-R. Li, J. Sculley and H.-C. Zhou, Chem.
Rev., 2012, 112, 869.
2 (a) J. Shang, G. Li, R. Singh, Q. Gu, K. M. Nairn, T. J. Bastow,
N. Medhekar, C. M. Doherty, A. J. Hill, J. Z. Liu and P. A.
Webley, J. Am. Chem. Soc., 2012, 134, 19246; (b) S. Bourrelly,
P. L. Llewellyn, C. Serre, F. Millange, T. Loiseau and G.
Frey, J. Am. Chem. Soc., 2005, 127, 13519; (c) Y. Yan, S. H.
Yang, A. J. Blake and M. Schroder, Acc. Chem. Res., 2014, 47,
296; (d) S. Prez-Yez, G. Beobide, O. Castillo, M. Fischer,
F. Hoffman, M. Frba, J. Cepeda and A. Luque, Eur. J. Inorg.
Chem., 2012, 5921; (e) O. K. Farha, I. Eyazici, N. C. Jeong,
B. G. Hauser, C. E. Wilmer, A. A. Sarjeant, R. Q. Snurr, S. T.
Nguyen, A. Yazaydin and J. T. Hupp, J. Am. Chem. Soc.,
2012, 134, 15016; ( f ) J.-R. Li, R. J. Kuppler and H.-C. Zhou,
Chem. Soc. Rev., 2009, 38, 1477.
3 H.-C. Zhou and S. Kitagawa, Chem. Soc. Rev., 2014, 43, 5415.
4 (a) Y. B. Zhang, H. L. Zhou, R. B. Lin, C. Zhang, J. B. Lin,
J. P. Zhang and X. M. Chen, Nat. Commun., 2012, 3, 642; (b)
H. Furukawa, N. Ko, Y. B. Go, N. Aratani, S. B. Choi, E. Choi,
A. Yazaydin, R. Q. Snurr, M. O'Keeffe, J. Kim and O. M.
Yaghi, Science, 2010, 329, 424; (c) O. K. Farha, A. Yazaydin, I.
Eryazici, C. D. Malliakas, B. G. Hauser, M. G. Kanatzidis,
S. T. Nguyen, R. Q. Snurr and J. T. Hupp, Nat. Chem.,
2010, 2, 944.
5 (a) M. Eddaoudi, J. Kim, N. Rosi, D. Vodak, J. Wachter, M.
O'Keeffe and O. M. Yaghi, Science, 2002, 295, 469; (b) J.
Duan, M. Higuchi and S. Kitagawa, Inorg. Chem., 2015, 54,
1645; (c) J. Cepeda, S. Prez-Yez, G. Beobide, O. Castillo,

This journal is The Royal Society of Chemistry 2016

CrystEngComm

10

11

12
13

14

15

16

17

M. Fischer, A. Luque and P. A. Wright, Chem. Eur. J.,


2014, 20, 1554; (d) D. Yuan, D. Zhao, D. Sun and H.-C. Zhou,
Angew. Chem., Int. Ed., 2010, 49, 5357.
(a) F. Nouar, J. F. Eubank, T. Bouquet, L. Wojtas, M. J.
Zaworotko and M. Eddaoudi, J. Am. Chem. Soc., 2008, 130,
1833; (b) Y. Yan, X. Lin, S. H. Yang, A. J. Blake, A. Dailly,
N. R. Champness, P. Hubberstey and M. Schroder, Chem.
Commun., 2009, 1025; (c) B. Zheng, J. Bai, J. Duan, L. Wojtas
and M. Zaworotko, J. Am. Chem. Soc., 2011, 133, 748.
(a) Y. F. Zhu, Y. Fang and S. Kaskel, J. Phys. Chem. C,
2010, 114, 16382; (b) L. Q. Ma, D. J. Mihalcik and W. B. Lin,
J. Am. Chem. Soc., 2009, 131, 4610; (c) M. Hirscher, Angew.
Chem., Int. Ed., 2011, 50, 581.
(a) D. M. D'Alessandro, B. Smit and J. R. Long, Angew.
Chem., Int. Ed., 2010, 49, 6058; (b) Y. Bae and R. Q. Snurr,
Angew. Chem., Int. Ed., 2011, 50, 2.
(a) R. K. Deshpande, J. L. Minnaar and S. G. Telfer, Angew.
Chem., Int. Ed., 2010, 49, 4598; (b) S. Ma, D. Sun, M.
Ambrogio, J. A. Fillinger, S. Parkin and H.-C. Zhou, J. Am.
Chem. Soc., 2007, 129, 1858; (c) R. K. Deshpande, G. I. N.
Waterhouse, G. B. Jameson and S. G. Telfer, Chem.
Commun., 2012, 48, 1574.
(a) J. L. C. Rowsell and O. M. Yaghi, J. Am. Chem. Soc.,
2006, 128, 1304; (b) J. L. C. Rowsell and O. M. Yaghi, Angew.
Chem., Int. Ed., 2005, 44, 4670; (c) S. Bureekaew, H. Sato, R.
Matsuda, Y. Kubota, R. Hirose, J. Kim, K. Kato, M. Takata
and S. Kitagawa, Angew. Chem., Int. Ed., 2010, 49, 7660.
(a) B. Chen, S. Ma, F. Zapata, F. R. Fronczek, E. B.
Lobkovsky and H. Zhou, Inorg. Chem., 2007, 46, 1233; (b) B.
Chen, S. Ma, E. J. Hurtado, E. B. Lobkovsky and H. Zhou,
Inorg. Chem., 2007, 46, 8490.
L. Bastin, P. S. Brcia, E. J. Hurtado, J. A. C. Silva, A. E.
Rodrigues and B. Chen, J. Phys. Chem. C, 2008, 112, 1575.
(a) K. Seki, Phys. Chem. Chem. Phys., 2002, 4, 1968; (b) R.
Kitaura, K. Seki, G. Akiyama and S. Kitagawa, Angew. Chem.,
2003, 115, 444 (Angew. Chem. Int. Ed., 2003, 42, 428).
J. M. Seco, D. Fairen-Jimenez, A. J. Calahorro, L. MndezLin, M. Prez-Mendoza, N. Casati, E. Colacio and A.
Rodrguez-Diguez, Chem. Commun., 2013, 49, 11329.
(a) R. Matsuda, R. Kitaura, S. Kitagawa, Y. Kubota, T. C.
Kobayashi, S. Horike and M. Takata, J. Am. Chem. Soc.,
2004, 126, 14063; (b) D. Bradshaw, J. E. Warren and M. J.
Rosseinsky, Science, 2007, 315, 977; (c) C. Serre, C. MellotDraznieks, S. Surble, N. Audebrand, Y. Filinchuk and G.
Frey, Science, 2007, 315, 1828.
(a) T. K. Maji, R. Matsuda and S. Kitagawa, Nat. Mater.,
2007, 6, 142; (b) J. L. C. Rowsell and O. M. Yaghi, Angew.
Chem., 2005, 117, 4748 (Angew. Chem. Int. Ed., 2005, 44, 4670).
(a) P. K. Yadav, N. Kumari, P. Pachfule, R. Banerjee and L.
Mishra, Cryst. Growth Des., 2012, 12, 5311; (b) J. Yang, J.-F.
Ma, S. R. Batten, S. W. Ng and Y.-Y. Liu, CrystEngComm,
2011, 13, 5296; (c) Z.-F. Chen, Z.-L. Zhang, Y.-H. Tan, Y.-Z.
Tang, H.-K. Fun, Z.-Y. Zhou, B. F. Abrahams and H. Liang,
CrystEngComm, 2008, 10, 217; (d) Y.-L. Liu, K.-F. Yue, B.-H.
Shan, C.-J. Wang and Y.-Y. Wang, Inorg. Chem. Commun.,
2012, 17, 30.

This journal is The Royal Society of Chemistry 2016

Paper

18 Z. H. Hu, B. J. Deibert and J. Li, Chem. Soc. Rev., 2014, 43,


5815.
19 Bruker Apex2, Bruker AXS Inc., Madison, Wisconsin, USA,
2004.
20 G. M. Sheldrick, SADABS, Program for Empirical Adsorption
Correction, Institute for Inorganic Chemistry, University of
Gottingen, Germany, 1996.
21 A. Altomare, M. C. Burla, M. Camilla, G. L. Cascarano, C.
Giacovazzo, A. Guagliardi, A. G. G. Moliterni, G. Polidori and
R. Spagna, J. Appl. Crystallogr., 1999, 32, 115.
22 (a) G. M. Sheldrick, SHELX-2014, Program for Crystal
Structure Refinement, University of Gttingen, Gttingen,
Germany, 2014; (b) L. J. Farrugia, J. Appl. Crystallogr.,
1999, 32, 837.
23 A. L. Spek, J. Appl. Crystallogr., 2003, 36, 7.
24 J. Rodrguez-Carvajal, J. FULLPROF 2000, version 2.5d;
Laboratoire Lon Brillouin (CEA-CNRS), Centre d'tudes de
Saclay, Gif sur Yvette Cedex, France, 2003.
25 (a) W. Lu, Z. Wei, Z.-Y. Gu, T.-F. Liu, J. Park, J. Park, J. Tian,
M. Zhang, Q. Zhang, T. Gentle III, M. Bosch and H.-C. Zhou,
Chem. Soc. Rev., 2014, 43, 5561; (b) Q. Yue, Q. Sun, A.-L.
Cheng and E.-Q. Gao, Cryst. Growth Des., 2010, 10, 44; (c)
D. J. Tranchemontagne, J. R. Hunt and O. M. Yaghi,
Tetrahedron, 2008, 64, 8553.
26 (a) P. Mahata, M. Prabu and S. Natarajan, Inorg. Chem.,
2008, 47, 8451; (b) P. M. Forster, A. R. Burbank, C. Livage, G.
Fery and A. K. Cheetham, Chem. Commun., 2004, 368; (c) J.
Cepeda, R. Balda, G. Beobide, O. Castillo, J. Fernndez, A.
Luque, S. Prez-Yez, P. Romn and D. Vallejo-Snchez,
Inorg. Chem., 2011, 50, 8437.
27 (a) M. Llunell, D. Casanova, J. Cirera, J. M. Bofill, P.
Alemany, S. Alvarez, M. Pinsky and D. Avnir, SHAPE (1.7),
University of Barcelona, Barcelona, 2010; (b) S. Alvarez, D.
Avnir, M. Llunell and M. Pinsky, New J. Chem., 2002, 26, 996;
(c) D. Casanova, P. Alemany, J. M. Bofill and S. Alvarez,
Chem. Eur. J., 2003, 9, 1281.
28 (a) TOPOS, http://www.topos.ssu.samara.ru; (b) V. A. Blatov,
M. O'Keeffe and D. M. Proserpio, CrystEngComm, 2010, 12,
44; (c) V. A. Blatov, A. P. Shevchenko and D. M. Proserpio,
Cryst. Growth Des., 2014, 14, 3576.
29 B. Fernndez, J. M. Seco, J. Cepeda, A. J. Calahorro and A.
Rodriguez-Diguez, CrystEngComm, 2015, 17, 7595.
30 (a) L.-N. Jia, Y. Zhao, L. Hou, L. Cui, H.-H. Wang and Y.-Y.
Wang, J. Solid State Chem., 2014, 210, 251; (b) N. Sikdar, A.
Hazra and T. K. Maji, Inorg. Chem., 2014, 53, 5993; (c) A.
Pichon, C. Mendicute-Fierro, M. Nieuwenhuyzen and S. L.
James, CrystEngComm, 2007, 9, 449; (d) P. V. Dau, M. Kim,
S. J. Garibay, F. H. L. Mnch, C. E. Moore and S. M. Cohen,
Inorg. Chem., 2012, 51, 5671.
31 C. Herdes and L. Sarkisov, Langmuir, 2009, 25, 5352.
32 L. Sarkisov and A. Harrison, Mol. Simul., 2011, 37, 1248.
33 (a) H. Chun, D. N. Dybtsev, H. Kim and K. Kim, Chem. Eur.
J., 2005, 11, 3521; (b) H. J. Park and M. P. Suh, Chem.
Commun., 2010, 46, 610.
34 J. J. Zhang, L. Wojtas, R. W. Larsen, M. Eddaoudi and M. J.
Zaworotko, J. Am. Chem. Soc., 2009, 131, 17040.

CrystEngComm, 2016, 18, 12821294 | 1293

Paper

35 O. Shekhah, H. Wang, M. Paradinas, C. Ocal, B. Schupbach,


A. Terfort, D. Zacher, R. A. Fishcer and C. Woll, Nat. Mater.,
2009, 8, 481.
36 N. L. Rosi, J. Kim, M. Eddaoudi, B. L. Chen, M. O'Keeffe and
O. M. Yaghi, J. Am. Chem. Soc., 2005, 127, 1504.
37 (a) D. Sun, Z.-H. Yan, M. Liu, H. Xie, S. Yuan, H. Lu, S. Feng
and D. Sun, Cryst. Growth Des., 2012, 12, 2902; (b) L. Han,
Z.-H. Li, J.-S. Chen, X.-P. Wang and D. Sun, Cryst. Growth
Des., 2014, 14, 1221; (c) X. Zhao, F. Liu, L. Zhang, D. Sun, R.
Wang, Z. Ju, D. Yuan and D. Sun, Chem. Eur. J., 2014, 20,
649.
38 L. Ma and W. Li, J. Am. Chem. Soc., 2008, 130, 13834.
39 (a) D. Sun, S. Ma, Y. Ke, D. J. Collins and H.-C. Zhou, J. Am.
Chem. Soc., 2006, 128, 3896; (b) S. Ma, D. Sun, M. Ambrogio,
J. A. Fillinger, S. Parkin and H.-C. Zhou, J. Am. Chem. Soc.,
2007, 129, 1858.
40 A. Schneemann, V. Bon, I. Schwedler, I. Senkovska, S. Kaskel
and R. A. Fischer, Chem. Soc. Rev., 2014, 43, 6062.
41 (a) Y.-S. Bae, A. Yazaydin and R. Q. Snurr, Langmuir,
2010, 26, 5475; (b) R. Kitaura, K. Seki, G. Akiyama and S.
Kitagawa, Angew. Chem., Int. Ed., 2013, 42, 428; (c) Y. Sakata,
S. Furukawa, M. Kondo, K. Hirai, N. Horike, Y. Takashima,

1294 | CrystEngComm, 2016, 18, 12821294

CrystEngComm

42
43

44

45
46

H. Uehara, N. Louvain, M. Meilikhov, T. Tsuruoka, S. Isoda,


W. Kosaka, O. Sakata and S. Kitagawa, Science, 2003, 339, 193;
(d) G. Kumari, N. R. Patil, V. S. Bhadram, R. Haldar, S. Bonakala,
T. K. Maji and C. Narayana, J. Raman Spectrosc., 2015, DOI:
10.1002/jrs.4766; (e) R. Babarao, C. J. Coghlan, D. Rankine,
W. M. Bloch, G. K. Gransbury, H. Sato, S. Kitagawa, C. J. Sumby,
M. R. Hill and C. J. Doonan, Chem. Commun., 2014, 50, 3238.
S. Kitagawa, R. Kitaura and S.-I. Noro, Angew. Chem., Int. Ed.,
2004, 43, 2334.
(a) ASiQwin V1.11, October 2010, Quantachrome Instruments;
(b) J. Landers, G. Y. Gor and A. V. Neimark, Colloids Surf., A,
2013, 437, 3; (c) J. Moellmer, E. B. Celer, L. Luebke, A. J.
Cairns, R. Staudt, M. Eddaoudi and M. Thommes,
Microporous Mesoporous Mater., 2010, 129, 345.
(a) K. Chen, Y.-S. Kang, L. Luo, Y. Zhao, P. Wang, Q. Liu, Y.
Lu and W.-Y. Sun, Polyhedron, 2014, 79, 239; (b) S. Q. Zhang,
F.-L. Jiang, M.-Y. Wu, J. Ma, Y. Bu and M.-C. Hong, Cryst.
Growth Des., 2012, 12, 1452.
Q. Zheng, F. Yang, M. Deng, Y. Ling, X. Liu, Z. Chen, Y.
Wang, L. Weng and Y. Zhou, Inorg. Chem., 2013, 52, 10368.
(a) A. D. Broadbent, Color Res. Appl., 2004, 29, 267; (b) T.
Smith and J. Guild, Trans. Opt. Soc., London, 1931, 33, 3.

This journal is The Royal Society of Chemistry 2016

Potrebbero piacerti anche