Download as pdf or txt
Download as pdf or txt
You are on page 1of 29

Coordination Chemistry Reviews xxx (2018) xxx–xxx

Contents lists available at ScienceDirect

Coordination Chemistry Reviews


journal homepage: www.elsevier.com/locate/ccr

Review

Conformation versatility of ligands in coordination polymers: From


structural diversity to properties and applications
Na Li a,c, Rui Feng b,c, Jian Zhu a, Ze Chang a,c,⇑, Xian-He Bu a,b,c,⇑
a
School of Materials Science and Engineering, National Institute for Advanced Materials, Tianjin Key Laboratory of Metal and Molecule-Based Material Chemistry, Nankai
University, Tianjin 300350, China
b
State Key Laboratory of Elemento-Organic Chemistry, College of Chemistry, Nankai University, Tianjin 300071, China
c
Collaborative Innovation Center of Chemical Science and Engineering (Tianjin), Tianjin 300072, China

a r t i c l e i n f o a b s t r a c t

Article history: Coordination polymers (CPs, including metal–organic frameworks, MOFs) have become an active topic in
Received 1 October 2017 chemical and materials science. The construction of CPs depends heavily on the organic ligands as well as
Received in revised form 17 May 2018 the metal centers, and the versatility of ligands contributes to the structural and property diversities of
Accepted 17 May 2018
CPs. Here, the structural diversity in CPs achieved by the conformation versatility of ligand is systemat-
Available online xxxx
ically reviewed. In addition, the effect of ligand conformation on the properties of CPs is highlighted with
discussions of potential applications of CPs in areas of storage, separation, and luminescence.
Keywords:
Ó 2018 Elsevier B.V. All rights reserved.
Coordination polymers
Conformation of ligand
Structural diversity
Storage and separation
Luminescence

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
2. Structural diversity of CPs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
2.1. Structural diversity of CPs based on flexible ligands . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
2.1.1. CPs based on flexible ligands bearing –(CH2)n– Spacers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
2.1.2. CPs based on flexible ligands bearing ACH2AOA Spacers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
2.1.3. CPs based on flexible ligands bearing ACH2ANA Spacers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
2.1.4. CPs based on flexible ligands bearing ACH2AS– Spacers. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
2.2. Structural diversity of CPs based on non-flexible ligands. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
2.2.1. CPs based on semi-rigid ligands . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
2.2.2. CPs based on rigid ligands . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
2.3. Impact of ligand flexibility on the structures of CPs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
3. Properties and applications of CPs with conformation-changing ligands . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
3.1. Storage and separation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
3.2. Luminescence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
4. Summary and outlook . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00

⇑ Corresponding authors at: School of Materials Science and Engineering, National Institute for Advanced Materials, Tianjin Key Laboratory of Metal and Molecule-Based
Material Chemistry, Nankai University, Tianjin 300350, China.
E-mail addresses: changze@nankai.edu.cn (Z. Chang), buxh@nankai.edu.cn (X.-H. Bu).

https://doi.org/10.1016/j.ccr.2018.05.016
0010-8545/Ó 2018 Elsevier B.V. All rights reserved.

Please cite this article in press as: N. Li et al., Conformation versatility of ligands in coordination polymers: From structural diversity to properties and
applications, Coord. Chem. Rev. (2018), https://doi.org/10.1016/j.ccr.2018.05.016
2 N. Li et al. / Coordination Chemistry Reviews xxx (2018) xxx–xxx

1. Introduction dynamic behavior of CPs and their potential applications as stor-


age/separation materials and luminophores.
Coordination polymers (CPs, including metal–organic frame- The continuous exploration of chemists has led to considerable
works, MOFs) have become an active topic in chemical and mate- progress to comprehend the coordination behaviors of ligands in
rials science [1–4]. Since the late 1950s when the concept of CPs the assembly process [23], and some methods have been developed
arose, a large amount of CPs have been reported, and the related to achieve multiple conformations of the ligands and the structural
research has covered a lot of fields [5–8]. The wide scientific atten- diversity of CPs [24–26]. However, the structural diversity as well
tion on CPs not only arises from their intriguing topologies and as the properties and applications of CPs based on the ligands with
diverse structures [9–12], but is also driven largely by their high conformation versatility have been rarely summarized. Here, we
potential applications in gas adsorption and separation [13], lumi- focus on the structural diversity of CPs constructed from ligands
nescent materials [14], and catalysis [8], etc [15,16]. The CPs are with versatile conformations (Scheme 1) to illustrate the key role
constructed from inorganic vertices (metal ions/clusters) and of these ligands in tuning structures and functions. Considering
organic linkers via coordination bonds and usually feature uniform the extensive reports, some representative examples and several
infinite crystalline networks. The diverse structures and properties common rules or design strategies for ligand conformation
originating from the inorganic-organic hybrid nature of CPs con- targeted constructions, such as ‘‘pillaring strategy”, ‘‘positional
tribute a lot to their versatility, which is beyond that of many other effect”, and ‘‘confined conformation”, have been summarized. In
inorganic materials. Therefore, much efforts have been made in the addition to the presentation of structural diversity of CPs, recent
investigation of new CPs with unique structures and specific func- advances of CPs in terms of their potential applications due to
tions, and various strategies and methods have been developed the structural diversity and conformation versatility of the ligands
accordingly [17–19]. It should be noted that the contribution of are thoroughly reviewed to demonstrate their critical roles. On the
metal ions to the diversity of CPs is limited in comparison with basis of all the points mentioned above, this review is expected to
organic ligands. This fact is largely due to the limited number of provide useful instructions for structure and function targeted con-
metal-center species and their relatively restricted coordination struction of CPs.
modes. In contrast, benefiting from the almost unlimited potential
of organic ligands in species and conformations, more effort has
2. Structural diversity of CPs
been dedicated to the design and application of unique organic
ligands for the construction of CPs, leading to the blooming of
CPs are attractive for their diverse structures and applications,
CPs research [20,21].
which were mainly determined by the characteristics of inorganic
Among the various organic ligands, multi-donor organic ligands
and organic building units. Compared with the inorganic building
with versatile conformations have attracted much attention [22].
units with limited number of species and coordination modes,
The different conformations of the ligands often give rise to a vari-
the organic ligands exhibit diverse configurations and conforma-
ety of symmetries and connection modes. It should be noted that
tions, and are supposed to contribute greatly on the construction
since the construction of CPs is mainly based on the self-
of CPs with multiple structures, topologies, and distinct properties
assembly of metal centers and organic ligands, the conformations
(All the network topology symbol are presented according to that
of the ligands are greatly determined by the physical factors (tem-
collected and summarized by the Reticular Chemistry Structure
perature, pressure, etc) and chemical factors (concentration, pH,
Resource (RCSR) [27–29]). To date, although an overwhelming
solvent, etc) of reaction conditions. Therefore, it is challenging to
number of ligands have been developed through convenient syn-
precisely acquire and regulate the structures of the CPs based on
thesis, the definite prediction of the structures of CPs from organic
this kind of ligand. On the other hand, the ligand in the framework
ligands is still a challenge due to the ungovernable freedom of the
of CPs may respond to external stimuli (temperature, pressure,
conformation of ligands. Hence it is quite advisable to discover
guest molecules, etc) to change their conformations, which could
some common rules to understand the influence of ligands with
trigger the structure transformation and dynamic behaviors of
conformation versatility on the structural diversity of CPs, which
the CPs. As a consequence, much effort has been devoted to making
could further benefit the rational design and synthesis of CPs. In
use of conformation versatility to realize the structural diversity of
this section, we will emphasize the impact of organic ligands (flex-
CPs.
ible and non-flexible) featuring conformation versatility on struc-
According to the origin of the conformation versatility, the
tural and topological diversification of CPs. Some strategies
organic ligands could be mainly classified into two classes: (i)
including ‘‘pillaring strategy” and ‘‘positional effect” are introduced
ligands with ‘‘flexible” spacers (alkyl chains and ether chains,
to reveal that how the configuration of the ligand could be regu-
etc.) showing tunable relative distance and orientation between
lated in the CPs to promote the rational design and construction.
the coordination sites. (ii) ligands with ‘‘non-flexible” spacers
Many factors impacting the conformation of ligands, such as tem-
showing potential multiple conformations contributed by the
perature, solvent, and template in the assembly process, and the
bending and rotation of chemical bonds between the coordination
external stimuli that could affect ligands in the CPs have also been
sites. Typically, ligands with flexible spacers are more varied in the
introduced. Notably, since the definition of ‘‘flexible ligand” and
coordination assemblies and tend to generate versatile and dis-
‘‘non-flexible ligand” is somewhat crossed considering different
torted framework due to their structural features. In contrast, the
moieties of the ligand, some ligands with flexible spacers (usually
conformational change of non-flexible linkers, originating mainly
containing C, O, N, S, etc hybrid atoms and –(CH2)n– moieties) are
from the rotation of coordination sites around single bonds and
categorized as flexible ligands according to their observed confor-
bending of ligand backbones, is relatively inconspicuous but can
mations. The construction of CPs on the basis of mixed-ligands are
also contribute to the diversity of CPs. Aiming at the structural
also included in this section.
diversity of CPs, many flexible organic moieties, such as alkyl,
amide, sulfur ether, azo, etc, have been applied as versatile spacers
of ligands. Beyond the contribution to the diversity of CPs’ struc- 2.1. Structural diversity of CPs based on flexible ligands
tures, the versatile conformation of organic ligands may also
endow the according CPs with variable pore geometry and ligand In general, the spacers of flexible ligands contain rotatable cova-
originated emissions, which will benefit the achievement of lent bonds, which stem from at least one sp3 hybrid atom (usually

Please cite this article in press as: N. Li et al., Conformation versatility of ligands in coordination polymers: From structural diversity to properties and
applications, Coord. Chem. Rev. (2018), https://doi.org/10.1016/j.ccr.2018.05.016
N. Li et al. / Coordination Chemistry Reviews xxx (2018) xxx–xxx 3

Scheme 1. The structural diversity of CPs constructed with organic ligands showing conformation versatility (the figures were reproduced with permissions from the
copyright holders).

C, O, N, S, etc.) or their combinations to induce conformation topologies [32–37], including srs, ThSi, dia, pcu, sqc, zst, etc. Gener-
changes. Typically, hybrid atoms in the flexible ligands do not par- ally speaking, the diversification of flexible ligands could be real-
ticipate in the coordination in the self-assembled process except ized by the –(CH2)n– moieties coupled with different carboxylate
for some functionalized moieties, such as dithioether and disulfox- and N-donor groups (usually pyridine, imidazole, triazole, tetrazole,
ide ligands. Instead, the flexible ligands adopt distinct symmetries etc.). The conformational freedom of these linkers is usually derived
as a consequence of mechanical compressing, stretching or twist- from the bending or twisting of the alkane chains, which commenly
ing, generating diverse structures of CPs. Generally, by deliberately show linear, helix, or V-shape configurations when coordinating to
controlling the nature of spacers between the functional groups of the metal centers. It can be concluded that the conformational free-
the flexible ligands, the resulting structures and properties of CPs dom of linkers has significant influence on the final topology and
can be systematically and regularly adjusted. The emphasis in this dimensionality of CPs. For example, a series of typical flexible dito-
part will be on the construction of CPs from flexible ligands includ- pic linkers (HOOCA(CH2)nACOOH) with varying alkane chains have
ing C, N, O, and S hybrid atoms or groups, showing how structural been widely investigated. The different conformations of linkers
diversity of CPs is gained from ligands with conformation versatil- can affect the relative disposition of secondary building unit
ity and illuminating the impact of the ligands on the resulting (SBU) and network topologies of CPs (Scheme 2) [38]. Three differ-
structures, some design principles such as ‘‘pillaring strategy” ent aliphatic dicarboxylic acids L1 (succinic acid, H2suc), L2 (adipic
and ‘‘spacer effect” are also introduced. acid, H2adip), and L3 (malonic acids, H2mal) have been selected to
monitor the effect of ligand flexibility on the self-assembly of CPs
2.1.1. CPs based on flexible ligands bearing –(CH2)n– Spacers in a mixed ligand system with PTA (1,3,5-triaza-7-phosphaadaman
As a class of typical flexible moiety showing various conforma- tane) as a prime building block (Fig. 2). Assembly of the different
tion, –(CH2)n– (n = 1,2. . .) moieties featuring high degree of flexi- carboxylate ligands, PTA, and Ag(I) yields three distinct structures
bility have attracted widespread attention in the construction of (1–3): Complex 1 features uninodal 3-c SP 1-periodic net with the
CPs. So far, a number of flexible ligands bearing –(CH2)n– moieties linear suc2- ligands; Complex 2 shows a uninodal 4-connected layer
have been applied for the construction of CPs to give charming net- with the skl topology containing crooked adip2- ligands; Complex 3
works. Various ligands with coordination sites from ditopic to hex- presents a uninodal 4-c dia topology with the mal2- ligand [39]. The
atopic [30], multiple geometries [31] and different length have structural difference of the complexes 1–3 should be mainly attrib-
been reported (Fig. 1). uted to the different A(CH2)nA chain length.
The ligands with –(CH2)n– skeleton show different geometry due The impact of the flexible ligands on the structure of CPs can be
to the bending and rotation of the flexible backbone, thus resulting further illustrated by two (3,4)-connected zinc complexes, {[Zn
in distinct connection modes and corresponding CPs with diverse (TIPA)(L3)1/2](NO3)3H2On (4) and {[Zn(TIPA)(L4)1/2](NO3)H2O}n

Please cite this article in press as: N. Li et al., Conformation versatility of ligands in coordination polymers: From structural diversity to properties and
applications, Coord. Chem. Rev. (2018), https://doi.org/10.1016/j.ccr.2018.05.016
4 N. Li et al. / Coordination Chemistry Reviews xxx (2018) xxx–xxx

Fig. 1. The referred ligands with A(CH2)nA spacers in the Section 2.

(5) (TIPA = tris(4-(1H-imidazol-1-yl)phenyl)amine), which have to 9.75 Å and the decrease of the dihedral angle of two adjacent
been reported by Zheng’s group by combining aliphatic dicar- inorganic nodes from 85.16° to 43.67°, leading to a converted
boxylic acids and two Zn-TIPA subnets [32]. Complex 4 with structure with a different chiral character and topology (Fig. 3).
shorter L3 (malonic acid) ligands reveals an interpenetrated net- In comparison to complex 5, the chiral subnet of complex 4 shows
work with utp topology. By replacing L3 with L4 (glutaric acid) fea- a strong second harmonic generation (SHG) response, which indi-
turing longer alkyl chain, a chiral porous complex with srs cates its potential application in nonlinear optics.
topology was obtained. In this case, the introduction of longer As another example of CPs based on flexible dicarboxylic acid
ligand results in an increase of the Zn  Zn distance from 6.75 Å with the A(CH2)nA skeletons, the assembly of Mn(II) with flexible

Please cite this article in press as: N. Li et al., Conformation versatility of ligands in coordination polymers: From structural diversity to properties and
applications, Coord. Chem. Rev. (2018), https://doi.org/10.1016/j.ccr.2018.05.016
N. Li et al. / Coordination Chemistry Reviews xxx (2018) xxx–xxx 5

Scheme 2. Structural diversity of CPs based on ditopic linkers (HOOCA(CH2)n-


ACOOH) showing different flexibility (the scheme was reproduced from Ref. [38],
with permission from the copyright holder).

Fig. 4. Distinct 3D networks and the L5 ligands with different conformation of 6 (a),
7 (b), and 8 (c) (the figures were reproduced from Ref. [40], with permission from
the copyright holder).

Fig. 2. The architectural diversity of three new bioactive silver-organic networks


(the figure was reproduced from Ref. [39], with permission from the copyright
holder).

Fig. 3. The structures of 4 and 5 achieved by tuning the length of flexible aliphatic Fig. 5. Schematic illustration of the distinct networks of 9–11 generated by
dicarboxylic acids (the figure was reproduced from Ref. [32], with permission from different ligands conformations (the figures were reproduced from Ref. [41], with
the copyright holder). permission from the copyright holder).

Please cite this article in press as: N. Li et al., Conformation versatility of ligands in coordination polymers: From structural diversity to properties and
applications, Coord. Chem. Rev. (2018), https://doi.org/10.1016/j.ccr.2018.05.016
6 N. Li et al. / Coordination Chemistry Reviews xxx (2018) xxx–xxx

zwitterionic dicarboxylate ligand L5 gives rise to three CPs with the coordinate groups connected with two flexible methylene car-
different structures (Fig. 4) [40], [Mn4(L5)4(N3)(H2O)4]n bon atoms, which result in drastic structural changes of CPs [43].
(ClO4)7nnCH3OH3nH2O (6), [Mn2.5(L5)(N3)5(H2O)2]n (7), and By using naphthalene as the flexible backbone between the flexible
[Mn2(L5)(N3)4]nnH2O (8) (L5 = 1,4-bis(4-carboxylato-1-pyridi bisimidazol groups, the L8 ligand (1,4-bis(imidazol-1-ylmethyl)na
nium)butane). In complex 6, L5 ligand shows three different kinds phthalene) is coordinated with different metal ions and fumaric
of conformations (TTG, TTT or TGT, T = transoid, G = gauche), in acid to form eight new metal complexes [44]. This flexible ligand
which different zigzag-like inorganic one dimensional (1D) chains can adopt varying conformations due to the relative rotation of
generate a three dimensional (3D) network with (456)2 imidazole groups. In addition, other multitopic derivatives
(4651861972) topology. Complex 7 shows a 3D net with a point [36,45–48] are available to construct CPs with structural diversity.
symbol of 610 with the L5 ligand adopts the TTT conformation. For example, benzene-1,3,5-tris(methylenephosphonic acid (L9)
Complex 8 displays a third type of 3D network and the L5 ligands and 2,4,6-trimethylbenzene-1,3,5-tris(methylenephosphonic acid
assume TTT and GTG conformations. It is obvious that the diverse (L10) possess three ‘‘phosphonic-arm” units linked by a CAC bond
conformations of the L5 ligand have led to the various networks of and show flexible conformation due to the rotation of CAC single
6–8 with different topologies. bond [49]. The reaction of the L9 with CuCl22H2O affords a 3D net-
In addition to distinct structures, the flexible conformation of work complex [Cu3(L9)(H2O)3.6]H2O (12), in which the ligand
A(CH2)nA moieties can also benefit the formation of CPs with pen- shows cis, trans, trans-conformation. In contrast, a similar reaction
etrating or interpenetrating structures [41,42]. Bai et al. have between L10 and Cu(NO3)23H2O gives rise to a two dimensional
reported three CPs with penetrating networks based on L6 (pamoic (2D) array of nanocage structure [Cu4(L10)2(H2O)4] (13). The for-
acid) and L7 (1,2-bi(4-pyridyl)ethane) ligands, {[Ni(L6)(L7)(H2O)2] mation of distinct coordination architectures clearly originates
3DMF}n (9), {[Zn(L6)(L7)]3DMF}n (10), and {[Cd2(L6)2(L7)2(H2O)] form the distinct conformations of the ligands: a cage-array struc-
3.5H2O}n (11). Their penetration modes are diverse due to the con- ture and a 3D network arises from the diverse conformation of the
formation change of ligands (Fig. 5) [41]. Compound 9 exhibits a 3- flexible ligands (Scheme 3).
fold interpenetrating CdSO4 net with point symbol (658), 10 shows Apart from the structural diversity achieved during the assem-
a roto-translational [2+2] interpenetrating diamond framework, 11 bly process, the external stimuli are also able to vary the ligand
possesses a 3-fold interpenetrating network with point symbol configuration to trigger a minor change of the structure of corre-
(6748)(6478). The diversity of interpenetration is strongly sponding CPs, especially in the aspect of pore structure of frame-
related to the conformation change of flexible ligands. As shown work. In addition, the flexible ligands with ACH2A group spacers
in Fig. 5, the most remarkable difference of these penetrating most likely show explicit freedom as a consequence of the restric-
framework stems from the nature of flexible ligand. The L6 ligand tion of torsional flexibility, which is more undemanding to regulate
shows symmetrical and asymmetrical conformations, while the L7 and design the anticipated structures. Taking L11 (bis(4-(4-carbox
ligand exhibits different conformations with versatile distances yphenyl)-1H-pyrazolyl)methane) containing both bis-pyrazolyl
and dihedral angles between the two pyridyl rings, which could rings and a flexible ACH2A group as an example, the reaction of
cooperate with metal ions to result in diverse structures (Table 1). L11 with Cu(II) yields a regular 2D grid framework with a 44 net
By attaching the flexible A(CH2)nA moieties to rigid conjugated [Cu(L11)H2O]xS (S = solvent molecules) (14) [50]. In this case,
backbones, numerous ligands (such as L8, L11, and L13, etc., Fig. 1.) the Cu(II) ions are coordinated by two carboxylate groups, one
have been designed to assemble with different metal ions to con- dipyrazolylmethane unit, and one water molecule. The dipyra-
struct CPs with more structural diversity. The conformational flex- zolylmethane unit shows a V-shape chelating configuration. In
ibility of these ligands mainly arises from the relative rotation of addition, the neighboring layers in 14 present strong hydrogen

Table 1
Conformation details of L6 and L7 ligands in compounds 9–11.

Compound Dihedral angles between two M  M distance Dihedral angles between M  M distance L7 Structure
pyridyl rings (deg) separated by L7 two naphthalate separated by L7 conformation
ligand (Å) rings (°) ligand (Å)
9 0 13.5 136.7 15.9 anti 3-Fold interpenetrating CdSO4
net (658)
10 27.8 13.1 94.1 15.1 anti Roto-translational [2+2]
interpenetrating net
11 0 13.7 79.8 15.08 anti 3-Fold interpenetrating (6748)
(6478)

Scheme 3. Schematic representations of the assembly of 3D network and 2D cage array by adjusting the conformations of ligands (the figure was reproduced from Ref. [49],
with permission from the copyright holder).

Please cite this article in press as: N. Li et al., Conformation versatility of ligands in coordination polymers: From structural diversity to properties and
applications, Coord. Chem. Rev. (2018), https://doi.org/10.1016/j.ccr.2018.05.016
N. Li et al. / Coordination Chemistry Reviews xxx (2018) xxx–xxx 7

bonding between the hydrogen atoms from water ligands and the of alkyl chains. When the solvent guests in the framework were
oxygen atoms from carbonyl ligands. By heating the activated sam- removed, the distance between the layers was decreased due to
ple, a 3D structure [Cu(L11)H2O]-ac (14a) with constricted pore is the shortening of flexible ligands used for padding the empty
obtained (Fig. 6). The 2D-to-3D transformations of framework space. In addition, either elongating and shortening or rotating of
could be considered as a result of changes of the coordination envi- A(CH2)nA group spacers may be applied to regulate the pore struc-
ronment of the metal ions from a five-coordinate structure to a dis- tures and sizes of CPs purposefully. For example, bpe (L7) can be
torted six-coordinate octahedron. This change of the structure is regarded as a flexible ligand for preparing CPs, as L7 possesses
also accompanied by a noticeably extended angle between the the flexible A(CH2)2A spacers [52]. Assemblies of bpa as pillars
bis-pyrazolyl rings. These structural changes give rise to the con- with epda2- (H2epda = 5-ethyl-pyridine-2,3-dicarboxylic acid) and
striction of the limiting pore diameter (3.6 Å) by the bending of Co (II) ions afford a pillared-layer CP {[Co2(epda)2(L7)(H2O)2]
L11 ligand (Fig. 6a), resulting in the exceptional CO2/N2 selectivity 3H2O (15). Similar to 15, CP 15a is obtained by desolvating 15
of the 3D framework. under vacuum. In complex 15a, the rotation of the CAC single
For the regulation of the structure of CPs based on flexible bonds decreases the interlayer distance, which is caused by the
ligands, the ‘‘pillaring strategy” has been developed as an effective removal of guest molecules. Comparing to that of 15, the pore size
method. With this method, isoreticular CPs can be designed and of 15a shrinks from 5.68  4.83 Å2 to 4.84  3.80 Å2 (Fig. 7).
built on the basis of the certain connection modes between the As a typical flexible ligand, bpe has been widely used for the
layer and pillar moieties. As a result, the conformation of pillar construction of CPs. Therefore, a large number of conformations
ligands bearing flexible A(CH2)nA spacers is relatively regulated. of bpe have been determined, which make it suitable for the con-
For example, Alberti and co-workers have successfully synthesized clusion and illustration of the principles of the relationship
a suite of pillar-layer CPs by altering the lengths of alka- between the conformations of the ligand and the structures of
nediphonate ligands [51]. Notably, the interlayer distance of the CPs. A list of the conformation details of bpe ligands and the
pillar-layered CPs can be easily regulated by changing the length structures of the corresponding CPs is given in Table 2. From the
list, it is obvious that the orientation of the pyridine ring can vary
in a wide range from ca. 9.657 to 5.725 Å, which originates from
the flexible backbone of the ligand. On the other hand, the dihedral
angles between two pyridyl rings also can vary from 0 to 83.968°.
Then the combination of different orientations and angles give rise
to the various conformations of the ligand. In the perspective of the
structures of CPs, it is obvious that most 3D framework and 2D lay-
ers are obtained with the ligands showing expanding trans confor-
mations, while ligands with cis conformations result in 1D chains.
These principles are applicable to other flexible ligands showing
similar conformations.

2.1.2. CPs based on flexible ligands bearing ACH2AOA Spacers


The distortion or twist of ACH2AOA spacers through the rota-
tion of single bonds endow the flexible ligands with great flexibil-
ity to give different configurations, which may provide the
diversity of the resulting CPs.
A pcu-CP can be constructed by using a flexible ligand
(1,1-bis-[3,5-bis(carboxy)phenoxy]methane) (L12), [Cu6(L12)3
(DMF)(H2O)5]n(DMF)x (16) [69]. In this complex, all flexible
ligands L12 adopt one kind of conformation to assemble the Oh
symmetric small rhombihexahedron, which is different from the
similar pcu net based on the flexible tetracarboxylate ligands with
two crystallographically independent orientations. Zaworotko and
co-workers utilized conceptual ‘‘bottom-up” design principle to
obtain a nanoscale faceted polyhedra with a diameter of 2.73 nm,
[Cu24(L13)12(H2O)16(DMSO)8]n (L13 = 1,3-bis(5-methoxy-1,3-ben
zene dicarboxylic acid)benzene) (17). In 17, the L13 shows two
kinds of conformations: one adopts a syn-conformation to form a
cylinder, while the other adopts an anti-conformation to result in
a cylinder with different size [70].
The flexible ACH2AOA moiety can offer helical structures in
CPs, demonstrating additional conformational diversity [71–76].
Utilizing L14 (5-(3,5-dicarboxybenzyloxy)isophthalic acid), Zhang
and co-workers have successfully obtained a series of novel helical
lanthanide CPs {[Ln2(L14)2](H2O)3(Me2NH2)2}n (Ln = Tb (III) in
example 18), which possess an isomorphic structure with (6,6)
topological net [77]. In this structure, the two phenyl rings of all
L14 ligands are nonplanar and their dihedral angle is approxi-
mately 90°. Such architectonic L14 ligand linked by Tb(III) ions
Fig. 6. Representations of the coordination environment (a), pore structures (b),
and structures (c) of 14 (left column) and 14a (right column). Color: Cu, green; C,
gives rise to two kinds of right-handed (Tb–L14)n helical chains.
Black; N, Blue; O, Red; H, Grey (the figure was reproduced from Ref. [50], with In addition, another helical chain along the c direction is obtained
permission from the copyright holder). by connecting the [Tb2O16] fragments and oxygen atoms from the

Please cite this article in press as: N. Li et al., Conformation versatility of ligands in coordination polymers: From structural diversity to properties and
applications, Coord. Chem. Rev. (2018), https://doi.org/10.1016/j.ccr.2018.05.016
8 N. Li et al. / Coordination Chemistry Reviews xxx (2018) xxx–xxx

Fig. 7. (a) Reversible transformations of 15 and 15a due to the rotation of CAC single bond of L7 pillar, (b) The channels showing reversible contraction and expansion of the
pore (the figure was reproduced from Ref. [52], with permission from the copyright holder).

carboxyl groups of the L14 ligand. All of the chains were alternately tion of the CPs from a 3D framework with a pore size of 6 Å  6 Å
arranged to exhibit a 3D homochiral framework (Fig. 8). (20) to a 2D layer network with a pore size of 15 Å  8 Å (21) then
L15 (5,50 -methylene-bis(oxy)diisophthalic acid) and L16 to a 3D nonporous framework (22). There results also indicate that
(5,50 -(ethane-1,2-diyl)-bis(oxy)diisophthalic acid) ligands have the conformation of the ligand can be regulated by the reaction
been investigated as fascinating ligands for constructing CPs, as conditions, thus influencing resultant structure variations of CPs
the introduction of the flexible ACH2AOA spacer brings more (Fig. 9).
structural diversity [78]. In this regard, reaction of L15/L16, differ- Besides, a series of CPs based on typical flexible tritopic ligand
ent d-metal ions, and additional bipyridyl ligands affords a series of with ACH2AOA spaces have been designed and synthesized. It
distinct structures. In these complexes, the structural diversity of should be noted that the central benzene backbone of most C3-
CPs is facilitated by the conformational flexibility of the ligands. symmetry ligands is often perpendicular to the ACH2AOA bridged
The different shape of the ligands is mainly caused by the rotation dangling rings rather than coplanar to them, to adapt the coordina-
of the two isophthalic acid motifs around the central aliphatic tion of metal ions, which can benefit the formation of highly con-
spacers, which could be quantified by the dihedral angle d between nected networks. For example, through the combination of
the two phenyl rings and the torsion angle of s (CphenylAO/ hexanuclear {Cd6} building units and L19 (2,4,6-tris[1-(3-carboxyl
OACphenyl in L15, and OACH2ACH2AO in L16) shown in Table 3. phenoxy)ylmethyl]mesitylene, a 3D metal–organic framework
The flexible trigonal ligands with relatively more conforma- [NH2(CH3)2][Cd6(L19)4(DMF)6(HCOO)] (23) has been synthesized
tional freedom make corresponding CPs show various fascinating by Bu and co-workers [81]. Complex 23 shows a unique cages based
structures and topologies. For example, L17 (5,50 ,500 -(2,4,6-tri framework, in which the assembly of cage is triggered by the ligands
meth-ylbenzene-1,3,5-triyl)trismethylene-trisoxy-triisophthalic featuring cis,cis,cis-conformation while the linkage of cages and
acid) is a typical flexible tritopic ligand bearing three 1,3- expanding of network is achieved with the ligands with cis,trans,-
benzenedicarboxylate (bdc2-) units, the flexibility of this ligand is trans-conformation. Bu and co-worker further employed the L19
derived from the three ACH2AOA spacers. Utilizing the flexible ligand in the design and synthesis of a luminescent 3D CP
L17 ligand, Hong and co-workers synthesized a 3D porous polyhe- [NH2(CH3)2]2[Cd17(L19)12(l3-H2O)4(DMF)2(H2O)2]solvent (24) [82].
dral metal–organic framework, [Zn7L172(OH)2(H2O)9]12.25H2O In this complex, hexanuclear and trinuclear Cd(II) building units are
(19), with Zn2(COO)4 and Zn2(COO)3 SBUs [79]. Through the com- bridged by L19 ligands showing cis,cis,cis-conformation to give a 3D
bination of hexacarboxylate ligands L17 and Zn-based SBUs as framework with a unique (3, 6, 12)-connected topological net.
nodes, a porous polyhedral CP with four types of polyhedral cages On the basis of the systematic investigation of this type of
and novel (3,4,6)-connected net was obtained. ligands, an effective strategy for the design and synthesis of
Another example of conformation-induced structural diversity multi-walled cage-based porous CPs was demonstrated by Bu
was reported by Hou and coworkers [80]. The authors employed and co-workers utilizing a mixed molecular building block (MBB)
the flexible L18 (5,50 -(pentane-1,2-diyl)-bis(oxy)diisophthalic acid) method. By means of this strategy, two double-walled octahedral
ligand and Co(II) ions to construct three porous complexes cage-based MOFs with cage-in-cage structure, [M3L192(TPT)2xG]n
{[Co2(L18)(CH3CN)(H2O)3]CH3OHH2O}n (20), {[Co2(L18)(CH3CN) (M = Ni(II) for 25 and Co(II) for 26; TPT = 2,4,6-tris(4-pyridyl)-1,3,
(H2O)3]}n (21), and {[Co5(L18)2(l3-OH)2(H2O)6]2H2O}n (22) under 5-triazine, G = guest molecules), were obtained with isomorphic
distinct reaction conditions. By increasing the concentration of structure (Fig. 10) [83]. The critical point of this strategy is the
1,4-dioxane in reaction conditions, the conformation of the L18 size-matching of mixed MBB, therefore the flexible L19 ligands
ligand transforms from slight distortion (20.0°) to planarity (0.3°) and rigid TPT ligands are selected to construct double-walled
then to a distorted form (82.8°), along with the structural distinc- caged-based MOFs for the following reasons: (i) flexible L19

Please cite this article in press as: N. Li et al., Conformation versatility of ligands in coordination polymers: From structural diversity to properties and
applications, Coord. Chem. Rev. (2018), https://doi.org/10.1016/j.ccr.2018.05.016
N. Li et al. / Coordination Chemistry Reviews xxx (2018) xxx–xxx 9

Table 2
The diverse conformations of bpe (L7), and structural diversity of CPs based on bpe ligand.

Compound Ligand conformation Dihedral angles of two Ligand Structure of CP Ref.


pyridyl rings (°) distance (Å)
10 [Ag2(bpe)5](BF4)2 11.805 9.622 2D layer [53]

20 [Co5(bpe)9(H2O)8(SO4)4](SO4)14H2O 63.314 9.287 3D chiral framework [54]

21.867 9.322

30 [Co(bpe)2(H2O)2](ClO4)2(H2O)2 44.8 6.268 1D chain [55]

40 [Cd(bpe)(SO4)] 0 9.657 3D framework [56]


50 [Cu(bpe)(SO4)(H2O)]2H2O 44.876 6.58 2D (4,4)-layer [56]

60 [Co2(bpe)3(SO4)2(MeOH)2]xSolvent 0 9.340 3D framework with catenated 6-membered [56]


rings
82.220 9.340

70 [Ni6(bpe)10(H2O)16](SO4)6xH2O 34.314 9.281 3D framework with entangled array [56]

45.422 9.422

0
8 [Cu2Br2(bpe)2] 36.857 9.330 2D framework [57]

90 [Cu2(CN)2(bpe)] 83.968 9.322 3D framework with twofold [57]


interpenetration

100 [ZntBu2(bpe)] 69.274 9.347 2D layer [58]

110 {[{(g5-C5Me5)MoS3Cu3}2(m-bpe)3Br4] 35.593 9.234 1D zigzag chain [59]


DMSO3MeCN}n

0 9.206

120 {[{(g5-C5Me5)MoS3Cu3}2(m-bpe)3.5Br4] 22.471 9.280 1D ‘‘Great Wall”-like chain [59]


MeCN}n
46.777 5.915

130 {[{(g5-C5Me5)MoS3Cu3}2(m-bpe)3(bpe) 12.000 9.246 1D double-stranded chains [59]


Br4]0.35DMF}n
12.775 9.333

81.000 9.331

45.491 9.301

140 {[{(g5-C5Me5)MoS3Cu3}2(m-bpe)2(m-Br) 7.243 9.238 1D zigzag chain [59]


(m3-Br)Br2]DMFMeCN}n
0
15 {[Cu3(CO3)2(bpe)3]2ClO4}n 1.062 9.301 3D framework with 2D kagomé layer [60]
0
16 [Co(NCO)2(bpe)2]n 74.157 6.786 1D chain [61]

170 [Mn(dca)2(bpe)] 60.060 5.725 1D chain [62]

180 [Ni(N(CN))2 (bpe)2](N(CN)2)(5H2O) 85.545 9.316 Twofold interpenetrated 3D framework [63]

(continued on next page)

Please cite this article in press as: N. Li et al., Conformation versatility of ligands in coordination polymers: From structural diversity to properties and
applications, Coord. Chem. Rev. (2018), https://doi.org/10.1016/j.ccr.2018.05.016
10 N. Li et al. / Coordination Chemistry Reviews xxx (2018) xxx–xxx

Table 2 (continued)

Compound Ligand conformation Dihedral angles of two Ligand Structure of CP Ref.


pyridyl rings (°) distance (Å)
190 [Cu(N3)2(bpe)] 79.185 9.345 2D layer [64]

200 [{Ni3(H2O)3(bpe)4}(V6O18)]8H2O 59.280 9.371 3D self-catenated framework [65]

44.208 5.726

210 {[Co(bpe)2(N(CN)2)]N(CN)24H2O}n 83.679 9.343 Twofold interpenetrated 3D framework with [66]


a-Polonium type topology

220 {[(WS4Cu4)I2(dpe)2](DMF)4}n 38.113 6.314 2D cluster polymer [67]

230 [Cu3(bpe)3(CO3)2](ClO4)2H2O 0 9.578 3D open framework [68]

Fig. 8. (a) Configurations of the ligands and coordination environments of the Tb(III) ion in 18; (b) The homochiral helical chains; (c) The 3D framework; (d) Schematic
representation of the network of 18 with a Schläfli symbol of {4867} (the figure was reproduced from Ref. [77], with permission from the copyright holder).

ligands and rigid TPT ligands possess similar trigonal configura- (m3-OH)2]Co(H2O)6xG}n (27) with the same ligands [84]. Com-
tions; (ii) C3-symmetric ligands could serve as building blocks for pared with 25 and 26, complex 27 contains different Co3O(COO)6
the construction of octahedral and tetrahedral cage upon coordina- SBU, which is connected by two flexible L19 ligands and one rigid
tion; (iii) flexible L19 ligands have the potential to form a cage face TPT ligand to form L19-TPT-L19 composite structure. For two L19
through the rotation of single bonds of ACH2AOA spacers. In the ligands, the central benzene ring is nearly perpendicular with the
structure of 23 and 24, the outer M12L194 octahedral cage is peripheral flexible phenyl carboxyl groups in syn-syn-syn configu-
obtained by the connection of six paddle-wheel units with four ration, while they are parallelly located at the two sides of one TPT
flexible L19 ligands, while inner six M(II) ions are further con- ligand to give ‘‘sandwich-like” triple layers. These layers are cross-
nected via four TPT ligands that are parallel to the outer L19 linked with Co3 SBU and thus form a chiral open 3D framework. It
ligands, resulting in a double-walled octahedral cage. The cages is worth noting that Co2(CO2)4 serving as the SBU in double-walled
are stacked to form a 3D cage-based framework by sharing the cage-based porous CPs connects the two hetero-double-layered
SBU as vertexs of the cages. building blocks L19-TPT, whereas each trinuclear Co3O(COO)6
To further illustrate the universality of the strategy, the authors SBU links three L19-TPT-L19 pairs in hetero-triple-walled CPs
reported the first 3D hetero-triple-walled CP {[Co6L194(TPT)2 (Fig. 11). By comparing the structures of 26 and 27, it is noticed

Please cite this article in press as: N. Li et al., Conformation versatility of ligands in coordination polymers: From structural diversity to properties and
applications, Coord. Chem. Rev. (2018), https://doi.org/10.1016/j.ccr.2018.05.016
N. Li et al. / Coordination Chemistry Reviews xxx (2018) xxx–xxx 11

Table 3
The conformations of L15 and L16, and structural diversity based on these ligands (the table was reproduced from Ref. [78], with permission from the copyright holder).

The metal type d0 /d00 /° s0 /s00 /° MetalAO/N bonds Nuclearity of Metal-metal distance(s)/Å Coordination Ligand conformation
of different range/Å metal cluster dimensionality
compounds
Zn 39.5(1) 142.3(3) 1.976(2)–2.151(2) 1 – 0D Flattened V-
conformation of L15
Zn 52.6(1) 132.1(3) 1.979(2)–2.015(3), 1 – 1D Flattened V-
1.942(2)–2.039(3) conformation of L15
Mn 64.0(1) 127.9(4) 2.078(2)–2.270(3), 3 3.941(1) 3D V-Shaped twisted
2.447(3) conformation of L15
Co 73.4(1) 141.0(5) 2.051(2)–2.206(3) 4 3.149(1), 3.370(1) 3D V-Shaped twisted
conformation of L15
Ag 12.6(2) 86.7(4) 2.181(2)–2.276(2) 1 2.8588(5), 2.7929(5), 3D U-Shaped
3.1778(4) conformation of L15
Cd 84.3(5) 80.2(4) 2.274(2)–2.628(2) 1 – 3Da Tetrahedral
conformation of L16
Zn 78.4(1) 66.3(3) 1.963(2)–2.109(2) 2 3.723(1) 3D V-Shaped bent
conformation of L16
Ni 84.2(1) 69.8(2) 1.987(1)–2.164(1) 4 3.069(1), 3.366(1) 3D Tetrahedral
conformation of L16
Ag 25.8(1), 64.8(4), 2.179(3)–2.255(3), 1 2.8303(5)–2.8919(6), 3D U-Shaped flat
28.9(1) 67.8(4) 2.448(2)–2.510(3) 3.224(4), 3.2693(4) conformation of L16
Zn 47.2(2) 131.6(2) 2.033(2)-2.432(2) 1 – 3 Da V-Shaped twisted
conformation of L15
Ag 21.7(2) 95.2(4) 2.098(2) 4 2.9–3.1b 3D U-Shape conformation
2.455(5) of L15
Co 88.9(1) 79.7(4) 1.984(2)–2.036(2) 1 – 2Dc Tetrahedral
conformation of L16
Co 83.2(1) 73.5(4) 2.012(3)–2.132(3) 1 – 3D Tetrahedral
conformation of L16
a
The 3D diamondoid arrays doubly interpenetrate in the crystal lattice.
b
The silver ions in this structure reveal twofold positional disorder.
c
The 2D coordination networks are sixfold interpenetrated.

that the coordination of Co2(CO2)4 SBU with L19 and TPT MBBs
tends to form closed polyhedral structures, while for Co3O(COO)6
SBU, a open network structure is preferred for Co3O(COO)6 SBU
(Fig. 11).
On the other hand, a modular pillaring strategy has been
developed by employing the flexible A(CH2)AOA based ligand
to design and construct a series of higher dimensional CPs with
distinguishing topologies, including sql, tbo, sql, kgm, rht, pcu
[26,70,85,86]. Eddaoudi et al. developed a ligand-to-axial strategy
to design and synthesized 3D CPs with expected topology. For
example, a trigonal bifunctional ligands L20 (5-(4-pyridinylme
thoxy)-isophthalic acid), containing isophthalic acid cores and
triazolyl-type moieties that are separated by ACH2AOA spacers,
are designed and utilized to yield a anticipated (3,6)-connected
3D MOFs.
So far, various flexible ligands with ACA((CH2)AO)4A spacers
have been reported (L21, L22, L23, Scheme 4) and have been
applied for the construction of CPs, and several reviews have been
made to demonstrate the influence of different configuration on
the resulting structures [22,23,87–90]. Despite the high symmetry
of the ligands, the produced CPs may crystallize in low symmetric
space groups due to the rotation of ACH2AOA moieties, which
benefits the construction of CPs with ferroelectric and non-linear
optical properties [87]. Anyway, most of them show high symmet-
ric structures, resulting in the centrosymmetric frameworks
[24,25,89,91]. In addition, since these ligands possess four
flexible CACH2AOAR arms, the adjustable connecting angles of
CRACcoreACR of arms are responsive to the diverse ligand configu-
rations. Typically, the ligands exhibit nearly a flat or tetrahedron
structure. The distinct configurations profoundly affect the final
Fig. 9. View of the conformations of the L18 ligand in corresponding structures of structure of CPs. Moreover, other multitopic flexible ligands con-
20–22 (a–c) (the figure was reproduced from Ref. [80], with permission from the taining ACH2AOA spacers were also designed for the fabrication
copyright holder). of CPs with fascinating network structures and properties [92,93].

Please cite this article in press as: N. Li et al., Conformation versatility of ligands in coordination polymers: From structural diversity to properties and
applications, Coord. Chem. Rev. (2018), https://doi.org/10.1016/j.ccr.2018.05.016
12 N. Li et al. / Coordination Chemistry Reviews xxx (2018) xxx–xxx

N-containing groups endow CPs with attractive properties and


potential applications in many fields, such as gas adsorption and
separation, luminescence, chirality, etc. So far, the N-containing
spacers could be classified as (i) imino groups; (ii) amide moieties
with alkyl/aryl spacers; (iii) amide groups; (vi) acylamide groups.
Other more complicated ones have also been designed to enhance
the structural versatility or improve their functional performance.
It is worth mentioning that the ligands with amide or acylamide
groups will be discussed in the following section since some of
them are recognized as semi-rigid ligands in reported articles.
By inserting a longer spacer with amide moieties and alkyl/aryl
groups, a series of flexible ligands can be available to construct
novel CPs [94–97]. Except for helical structures, interpenetrating
networks can be obtained by utilizing these ligands. For example,
the assembly of two flexible bis(pyridylurea) ligands and transition
metal sulfates afford threefold parallel interpenetrated CPs [96].
Notably, ligands containing amide moieties and alkyl groups show
higher degrees of freedom and unpredictable coordination behav-
ior [96,98]. The bis-(3-pyridyl)ethanediamide (L24) is taken as a
representative instance to illustrate how the flexibility of ligands
emerges through their versatile conformations [98]. The L24 tends
to give rise to diverse ligating topologies due to the following rea-
sons: (i) the rotation of the meta-positioning pyridyl rings depend-
ing on the relative orientation of the amide and pyridyl moieties;
(ii) the conformational change of the alkyl spacers gives rise to
diverse geometries of the ligands. The assembly of Cu(II) ions
and L24 under solvothermal condition yields a 1D-looped chain
complex 28. In this structure, the ethyl chain takes anti conforma-
tion with dihedral angle of 86° between the two amide planes, the
pyridyl rings point to the uniform direction with 80.36° dihedral
angle. In addition, amide plane and the pyridyl plane make two
angles of 28.13° and 23.91° with each other, respectively. More-
over, the conformation of the ligand L24 may have altered from
anti to several others, which promote its potential in constructing
CPs with various structures (Scheme 5).
By tuning the spacer lengths of such ligands with different alkyl
chains, the flexibility of ligands and corresponding CPs can be var-
ied. Reaction of L25 with Ag(I) results in ten CPs with distinct
structures (29a–29j) [99]. It also should be noted that L25 can
show anti-anti-anti (AAA), anti-anti-gauche (AAG), anti-gauche-anti
(AGA), anti-gauche-gauche (AGG), gauche-anti-gauche (GAG), and
gauche-gauche-gauche (GGG) conformations. The cis or trans
arrangement of the ligand can be classified by the relative orienta-
tion of the C@O (or NAH) groups (Fig. 12). The flexible ligand (L25)
adopts AAA trans syn-anti conformation in 1D concavo-convex
chain complex 29a, the 1D zigzag chain of complexes 29b–29h
possess the AAA trans syn-syn conformation ligand. The ligand in
complexes 29i–29j show the AAA trans anti-anti, AGG cis anti-anti,
and AGA cis anti-anti conformations, respectively. The different
ligand conformations arise from the changes of the counter anions
and solvents of the reaction conditions (Fig. 12).
On the other hand, ligands containing flexible Schiff-base spac-
ers have also been used to construct CPs, which reveal multiple
structures. For example, L26 (a,a0 -diamino-p-xylene and pyridine-
2-carboxaldehyde) has been used to react with Ag(I) to give three
Fig. 10. The structures of TPT and L19 (a), the unique paddlewheel unit (b), the new CPs (Fig. 13). In these compounds, the Ag(I) ions display sim-
triangular face (c), the outer M12L194 octahedral cage (d), a doublewalled ilar pseudotetrahedral N4 coordination geometry. Two parallel 1D
octahedral cage (e), and a cube of eight double-walled octahedral cages (f) in 25
polymeric chains are found in 30 with one triangular wave chain
or 26. M turquoise, C black, N blue, O red (the figure was reproduced from Ref. [83],
with permission from the copyright holder). based on the anti-conformation ligand and a trapezoidal wave
chain containing the alternate gauche ligand. Compound 31 con-
2.1.3. CPs based on flexible ligands bearing ACH2ANA Spacers tains a trapezoidal wave polymeric chain with two different anti-
Apart from the ligands mentioned above, the flexible conformation ligands. But the triangular wave chain of 30 only
N-containing spacers (simplified as A(CH2)ANA for uniform for- consists of anti-isomer of the ligand [100]. It has been noted that
matting) ligands have received much attention owing to their con- the solvent of the reaction plays an important role in the packing
figurational variability in addition to the fact that some flexible of chains in the CPs.

Please cite this article in press as: N. Li et al., Conformation versatility of ligands in coordination polymers: From structural diversity to properties and
applications, Coord. Chem. Rev. (2018), https://doi.org/10.1016/j.ccr.2018.05.016
N. Li et al. / Coordination Chemistry Reviews xxx (2018) xxx–xxx 13

Fig. 11. The rational construction of double-layered and triple-layered composite building blocks (the figure was reproduced from Ref. [84], with permission from the
copyright holder).

2.1.4. CPs based on flexible ligands bearing ACH2AS– Spacers


A family of flexible ligands bearing various S-donor spacers
(denoted as A(CH2)nASA) can be used to design structures of CPs
[101–103]. Typical moieties include thioether, sulfoxide,
dithioether, disulfoxide, and other S-donor bridging groups. Bu
and co-workers have focused on the study of structural diversity
and modulation of CP architectures based on flexible ditioether
and disulfoxide ligands (Fig. 14), and a comprehensive review on
the structural diversity of corresponding CPs has been made
[104]. As a representative example, three structurally related disul-
foxide ligands have been utilized to control the cavity sizes of CPs
Scheme 4. Illustration of flexible ligands with ACA((CH2)AO)4A spacer and
by varying the spacer length of ligands [105]. Three novel Cu(II)
different coordination groups.

Scheme 5. Conformations of ligand L24: notice that gauche conformation of the ethyl spacer results in increasing the bending of the ligand (the scheme was reproduced from
Ref. [98], with permission from the copyright holder).

Please cite this article in press as: N. Li et al., Conformation versatility of ligands in coordination polymers: From structural diversity to properties and
applications, Coord. Chem. Rev. (2018), https://doi.org/10.1016/j.ccr.2018.05.016
14 N. Li et al. / Coordination Chemistry Reviews xxx (2018) xxx–xxx

Fig. 12. The diverse structures of 29a–29j, three possible orientations for the pyridyl nitrogen atoms of the L25 ligand, and L25 ligand conformations and corresponding
angles for complexes 29a–29j (the figure was reproduced from Ref. [99], with permission from the copyright holder).

complexes, namely {[CuL272](ClO4)2}2n, (33) {[CuL283](ClO4)2}2n (34) which is favorable for constructing helical structures in some cases
and {[CuL293](ClO4)2}3n (35) have been prepared by employing differ- [111].
ent flexible meso-bis(sulfinyl) ligands (Fig. 15). The result shows that
both 33 and 34 are two-dimensional lamellar networks, while Cu(II) 2.2. Structural diversity of CPs based on non-flexible ligands
ions of the former coordinate to four O atoms from sulfoxides and
adjacent Cu(II) ions link ligands in two orientations forming rhom- For a long period, scientists have devoted to the control of the
bus cavity with dimensions of 8.05 Å  9.16 Å. In the latter case, assembly processes of CPs for architecturally targeted construction
the Cu(II) ions are linked by two distinct ligand conformations to by deliberately selecting inorganic SBUs and organic linkers. A
eventually obtain a 2D structure with the square cavity with dimen- large number of reports indicated that the structure of frameworks
sions of 9.44 Å  10.96 Å. With the increase of the spacer length, may be predicted when non-flexible ligands are utilized. In other
complex 35 with a novel 3D structure containing three types of words, in comparison with flexible ligands, the less freedom of
channels in three directions was obtained, yielding octahedral cavi- conformations in the restricted space of ordered frameworks make
ties with dimensions of 11.2 Å  11.8 Å  11.0 Å. Thereafter, a num- non-flexible ligands more promising for structure targeted con-
ber of structures based on flexible ligands bearing thioether, struction. Therefore, the influence of the relative stereochemistry
sulfoxide, dithioether, disulfoxide spaces have been reported [106– variation and conformation transformation of the non-flexible
110]. It is worth noting that some common rules can be summarized organic ligands on the structures and their eventually physical
in related systems: (i) the ligands based on thioether or dithioether and chemical properties are rarely summarized over recent dec-
spacers often take part in the coordination with different metal ions, ades. In this section, some representative non-flexible ligands are
and the adjustment of ligand conformations may promote the reviewed to highlight the influence of their conformation versatil-
change of coordination number of metal centers; (ii) whether S or ity on the structural diversity of CPs.
O atoms of sulfoxide or disulfoxide spacers in flexible ligands would
be involved in the coordination of metal centers depends on the nat- 2.2.1. CPs based on semi-rigid ligands
ure of metal ions; (iii) these flexible ligands often provide a variety of To realize the design of the desired structure, some flexible
conformations due to the change of flexibility and conformation, groups (such as imide, amide, and others) were inserted into the

Please cite this article in press as: N. Li et al., Conformation versatility of ligands in coordination polymers: From structural diversity to properties and
applications, Coord. Chem. Rev. (2018), https://doi.org/10.1016/j.ccr.2018.05.016
N. Li et al. / Coordination Chemistry Reviews xxx (2018) xxx–xxx 15

Fig. 13. Different conformations of L26 with different torsion angles across the phenyl ring (NAC  CAN) in 30, 31 and 32 (the figure was reproduced from Ref. [100], with
permission from the copyright holder).

backbone of non-flexible ligands to obtain semi-flexible ligands or V-shape geometries, thus serving as pillaring ligands with minor
toward structural diversity of CPs and to improve the properties twists [112–115]. For example, the reactions of divalent nickel or
of CPs. cobalt nitrates and 4,40 -oxybis(benzoic acid) (L30) give [[M(L30)
This kind of flexible ligands shows lower degree of flexibility (L31)]H2O] (M = Ni (II) for 36; M = Co (II) for 37, L31 = 4,40 -dipyridy
compared to the flexible ligands, a part of which can exhibit linear lamine), which has intriguing self-catenated two-dimensional

Please cite this article in press as: N. Li et al., Conformation versatility of ligands in coordination polymers: From structural diversity to properties and
applications, Coord. Chem. Rev. (2018), https://doi.org/10.1016/j.ccr.2018.05.016
16 N. Li et al. / Coordination Chemistry Reviews xxx (2018) xxx–xxx

Fig. 14. Structural diversity of CPs based on flexible ligands containing distinct dithioether, disulfoxide spacers (the figures were reproduced from Ref. [104], with permission
from the copyright holder).

Fig. 15. The structures of L27-L29 ligands and the cavities in 33–35 (H atoms and ClO
4 ions are omitted for clarity, the figure was reproduced from Ref. [105], with permission
from the copyright holder).

layered motifs [116]. The reactions of H2oba with cadmium and boxydiphenylamine) [119]. Notably, L32 shows minor steric inter-
potassium ions afford a heterometallic 3D porous framework, ference/hindrance owing to the fact that two carboxyl groups of
[K2Cd3(L30)4(H2O)2(DMF)2](H2O)2 (38), which shows high ligands often locate in different orientations with about 120° angle
adsorption capacity for polar molecules because of its polar partitioned by the imino moiety, leading to the formation of helices
CcarboxylAOcoreACcarboxyl moiety. Zhang and co-workers reported in CPs. In this CP, six Co2(COO)4 units connect six L32 ligands to
four new CPs based on a V-shaped L30 ligand, namely [M2(L30)2 form a twisted chair-form metallocyclic ring due to the flexible
(DMF)2]2DMF (39, M = Zn or Cu, DMF = N,N-dimethylformamide), V-shaped conformation of ligands, which was further intercon-
[Zn2(L30)2(bpy)]DMF (40, bpy = 4,40 -bipyridine), and [Cu2(L30)2 nected to obtain a diamond structure containing infinite left-
(bpy)0.5(DMF)]2DMF (41), which exhibit distinct topologies based handed helices (Fig. 17). This diamond structure is connected to
on M2(COO)4 SBUs and L30 ligands [117]. Recently, L30 was applied each other to form a 3D framework. By inserting the DPNDI ligand
as a flexible carboxylate ligand to construct a 3D lanthanide-based in the diamond net, an interpenetrating 3D network is obtained
CP (LnCP). The two phenyl rings of one L30 ligand could freely twist with a rare uninodal six-connected jsm topology.
along the AOA groups to adapt the coordination geometries of lan- Du et al. reported two Co(II)-based CPs based on the flexible
thanide ions to give {[Eu4(L30)6(H2O)9](H2O)} (42) (Fig. 16) [118]. 4,40 ,40 ’-(benzene-1,3,5-triyl-tris(oxy))tribenzoic acid (L33) ligand,
Beside the O atom, N atom is another widely used linking group [Co3(L33)2(H2O)2](DMF)4.5(H2O)4 (44) and [Co3(L33)2(H2O)2]
to construct semi-rigid ligand. For example, the reaction of DPNDI (DMF)3(H2O)4(dioxane)2.5 (45) [120]. L33 shows variable confor-
(DPNDI = N,N0 -di-(4-pyridyl)-1,4,5,8-naphthalenediimide) with Co mation due to three rotatable AOA moieties. For 44 and 45, the dif-
(II) in the presence of achiral flexible V-shaped ligand results in a ferent dihedral angles between three benzene arms and the central
unique homochiral CP [Co(L32)(DPNDI)0.5]n (43) (L32 = 4,40 -dicar benzene result in the formation of distinct 2-D (3,6)-connected

Please cite this article in press as: N. Li et al., Conformation versatility of ligands in coordination polymers: From structural diversity to properties and
applications, Coord. Chem. Rev. (2018), https://doi.org/10.1016/j.ccr.2018.05.016
N. Li et al. / Coordination Chemistry Reviews xxx (2018) xxx–xxx 17

Fig. 16. The coordination environments of (a) L30 ligand and (b) Eu(III) in 42; (c) The 2-D infinite layer of 42 with the 1-D and highlighted by blue rectangle (the figure was
reproduced from Ref. [118], with permission from the copyright holder).

construct mesoporous CPs. Notably, L34 with amino groups is


more elongated and diverse than other trigonal planar ligands,
resulting in CPs with various structures. Cage-based mesoporous
CPs could be obtained with L34 as the linker, while nonplanar con-
formation is mainly obtained. The conformation reveals a minor
average dihedral angle between the triazine and peripheral rings
due to the relatively flexible geometry [121–123]. Commonly, the
variation of frameworks structure is difficult to be achieved with
this conformation. In contrast, highly distorted Y-like conforma-
tion has been observed with L34, which could induce the formation
of CPs with nanosize 1D channels. Assembly of C3-symmetric L34
ligands and Zn(II) ions affords a unique CP [Zn2(ad)(L34)O1/4]
(Me2NH2)1/2(DMF)6(H2O)4 (46) (ad = adeninate) with trigonal
microporous of 0.8 nm and 3.0 nm hexagonal 1D channels [124].
In this structure, a [Zn8(ad)4] octahedral cage interacts with four
other cages through L34 ligands with Y-like conformation to form
a distorted cage-connector tetrahedron. The connection of tetrahe-
dron by sharing ligands affords the 3D porous framework (Fig. 19).
Because of the augmentability of L34, phase transition could be
Fig. 17. Four left-handed cogenerated helical chains constructed by the L32 ligand
in 43 (the figure was reproduced from Ref. [119], with permission from the achieved with CPs based on this ligand, resulting in different pore
copyright holder). sizes in response to temperature changes. For example, solvother-
mal reaction of Tb(III) and L34 affords a 3D lanthanide MOF,
bilayer networks. Complex 44 shows hexagonal pore structure {[Tb2(L34)2]4H2O6DMF}n (47) [125]. Significantly, single-crystal
with sizes of 13.0  13.0 Å2, while 45 represents distorted quad- X-ray diffraction (XRD) analysis at 293 K reveals that 47 owns a
rangular meshes with pore of 14.0  2.0 Å2. Notably, Co(II) coordi- large 1D rhombic channel A (ca. 27  23 Å2) and four small
nation geometry and ligand binding is similar in compounds 44 olivary-like channels B. Moreover, phase transition is observed at
and 45, the structural difference of 44 and 45 only arises from 121 K to obtain 47a with the same space group as 47. Notably,
the conformation of the L33 ligands (Fig. 18). the integral framework of 47a tends to squash with different pore
Similar tritopic carboxylate ligands with functional imine sizes of 29  20 Å2 due to the augmentation of dihedral angle
groups, such as 4,40 ,400 -(benzene-1,3,5-triyltris(azanediyl))tribenzo between the benzene and triazine rings of the ligand. Besides,
ate) (H3BTATB, L34) are considered as a typical semi-rigid ligand to 47a can be transformed into 47 at room temperature, realizing
illustrate the effect of semi-rigid backbone of ligands on the the reversibly dynamic transformation due to the bending of
structures of CPs. The H3BTATB has been designed and applied to semi-rigid ligands.

Please cite this article in press as: N. Li et al., Conformation versatility of ligands in coordination polymers: From structural diversity to properties and
applications, Coord. Chem. Rev. (2018), https://doi.org/10.1016/j.ccr.2018.05.016
18 N. Li et al. / Coordination Chemistry Reviews xxx (2018) xxx–xxx

Fig. 18. Crystal structures for 44 (a–e) and 45 (f–j) (the figure was reproduced from Ref. [120], with permission from the copyright holder).

Fig. 19. C3-symmetric H3BTATB ligand, distorted cage-connector tetrahedra, and space-filling model of 46 containing two types of 1D channels along the c axis (the figure
was reproduced from Ref. [124], with permission from the copyright holder).

The numerous ligands with amide spacers have been designed


to construct MOFs, resulting in structural diversity with interpen-
etrating [126–128], cage-based [129–132], and pillar-layered [133]
structures and various topologies including agw [134], rht [135],
asc [136], nbo [137], etc. In addition, amide-functionalized groups
are usually regarded as promising spacers for the enhancement of
CO2 adsorption performance of CPs. However, in this section, we
emphasize on studying the impact of amide modified ligands with
different conformation on structures of CPs. Dastidar and co-
workers functionalized the bispyridyl ligands to study the effect
of conformational flexibility of ligands in the self-assembly pro-
cess. Two complexes were obtained by using the N,N’-bis(3-
pyridyl)urea ligand (L35) [138]. The NACaromatic bond rotation of
L35 gives rise to various ligating topologies, which alter the result-
ing MOF architecture. For example, the assembly of L35 ligands
and Cu(II) ions yields two distinct structures: the syn-anti ligands
result in 1D zigzag chain structure, whereas syn-syn or anti-anti
ligands tend to form a metal-organic macrocycle or tubular struc-
ture (Scheme 6).

2.2.2. CPs based on rigid ligands


Compared with the above-mentioned CPs based on flexible and
semi-rigid ligands, rigid ligands are more attractive owing to their
potential to construct CPs with robust framework. However, as
mentioned in the introduction part, conformation versatility could Scheme 6. The structural diversity of CPs based on different conformations of L35
also be achieved with rigid ligands considering the change of the (the scheme was reproduced from Ref. [138], with permission from the copyright
orientation of coordination sites originated from the rotation of holder).

Please cite this article in press as: N. Li et al., Conformation versatility of ligands in coordination polymers: From structural diversity to properties and
applications, Coord. Chem. Rev. (2018), https://doi.org/10.1016/j.ccr.2018.05.016
N. Li et al. / Coordination Chemistry Reviews xxx (2018) xxx–xxx 19

single bond (including CAN or CAC) single between aromatic rings, and phenyl rings in L36 ligand. Compared with 48, all the ligands
thus the rigid ligands could achieve conformational variation adopt the same configuration, leading to the chirality of frame-
because of the rotations around the CAN or CAC bonds. On the work. In 50, the coordination mode III and IV can be observed with
other hand, for some moieties commonly serving as non-flexible dihedral angles of 37.69° and 33.25° between the pyridyl and the
backbone of ligands, such as AN@NA, AC@CA [139], and dithieny- phenyl rings. Different from 48 and 49, 50 has a three-
lethene, configuration transformation of the moieties (reversible dimensional (3,8)-connected (43)2(4661884)-tfz-d net (Fig. 20).
syn-anti conformation change and cyclization, etc.) can even occur Besides, the conformational change of rigid ligand can promote
under certain stimuli to bring in conformation versatility of the the formation of supramolecular isomers with diverse structures.
ligands. This characteristic of ligands can effectively contribute to Two zinc supramolecular isomers, a-[Zn(cis-L37)(H2O)]n (51) and
the structural diversity of CPs. Furthermore, CPs exhibiting stimuli b-[Zn(trans-L37)(H2O)]n (52), have been synthesized by using
responsive dynamic behavior could be readily obtained with 5-(3-carboxy-phenyl)-pyridine-2-carboxylic acid (L37) [144]. The
ligands bearing stimuli responsive moieties, which broaden the L37 ligand displays cis and trans conformation in 51 and 52, respec-
potential applications of CPs in a wide range of fields. To date, tively, leading to a unique 2D puckered layer in compound 51 and a
numerous CPs with dynamic behavior have been reported and 3D open framework in 52. In some cases, the chiral framework can
some excellent reviews have been given [140,141]. However, the be obtained by selecting the rigid ligand due to the free rotation of
structural diversity of CPs constructed with this kind of ligands single bonds [145–147].
have been rarely summarized from the viewpoint of geometric ‘‘Positional isomeric effect” can be considered as an effective
variation of rigid ligands. Herein, we will emphasize the impact strategy to regulate the structures of CPs. The rigid angular dipyr-
configuration transformation of rigid ligands on the structural idyl ligands bearing oxadiazole spacers have attracted much atten-
diversity of CPs. tion for their various potential conformations. With this in mind,
With the above consideration, the asymmetric rigid ligands Bu and co-workers have made much effort in the construction of
with possible syn- and anti- conformations can be considered as CPs with this kind of ligands [148–150]. Here, 2,5-bis(4-pyridyl)-
excellent precursors to construct diverse coordination architec- 1,3,4-oxadiazole (L38) and 2,5-bis(3-pyridyl)-1,3,4-oxadiazole
tures [142]. For example, L36 (5-(40 -carboxylphenyl) nicotinic acid, ligands (L39) were taken as two representative ligands. These
H2cona) has been applied to be coordinated with Ni(II) ions to form ligands show various geometries because of the change of the dihe-
multiple CPs. The L36 as a rigid ligand is promising to exhibit syn dral angles between the oxadiazole ring and two 4-pyridyl rings as
and anti conformations, and the nature of dicarboxylate groups well as between two pyridyl rings located in periphery, coupled
makes it possess a variety of coordination modes. More impor- with versatile bonding modes. With respect to L38, the rotation
tantly, dihedral angles between the pyridyl and phenyl rings of of two peripheral 4-dipyridyl motif around the center oxadiazole
L36 can vary through the rotation of CAC single bonds, which groups results in different dihedral angles of different rings, which
may bring axial chirality. Reaction of H2cona and Ni(II) leads to has little effect on their structures. In contrast, L39 ligands with 4-
the formation of three different CPs Ni(L36)(H2O)2 (48), Ni2(L36)2 dipyridyl motifs exhibit three trans or cis conformations on account
(H2O)30.5H2ODMF (49), and Ni3(L36)2(L36)2(H2O)42H2O (50) of rotation of dipyridyl skeletons, which have a remarkable impact
[143]. Complex 48 features a 2D (3,6)-connected (43)2(466683)- on the structures of the CPs. Based on these features of the ligands,
kgd topology with dinuclear SBUs act as 3-connected and these rigid ligands have been reported to be expeditious for the
6-connected nodes. The ligands show coordination mode I and preparation of versatile CPs (Scheme 7).
the dihedral angles between pyridyl and phenyl rings are 29.31°. Besides, three interpenetrating CPs isomers, namely, Zn(L40)
Compound 49 shows an unprecedented (3,6)-connected chiral (bpdc)(H2O)0.5 (53), Zn(L40)(bpdc)(DMF) (54), and Zn(L40)
(426)2(44628810)-anh topology, in which the ligands adopt coor- (bpdc)(H2O)2(DMF) (55), were prepared by the assembly of Zn
dination type II with dihedral angles of 37.73° between the pyridyl (II), biphenyl-4,40 -dicarboxylic acid (H2bpdc), and N,N’-bis(3-pyri

Fig. 20. Views of the coordination modes of L36 and structures of complexes 48–50 (the figure was reproduced from Ref. [143], with permission from the copyright holder).

Please cite this article in press as: N. Li et al., Conformation versatility of ligands in coordination polymers: From structural diversity to properties and
applications, Coord. Chem. Rev. (2018), https://doi.org/10.1016/j.ccr.2018.05.016
20 N. Li et al. / Coordination Chemistry Reviews xxx (2018) xxx–xxx

Scheme 7. The structural diversity of CPs based on L38 and L39 ligands (the scheme was reproduced from Ref. [148–150], with permissions from the copyright holders).

dyl)benzene-1,3-dicarboxamide (L40) at different reaction temper- However, the crystallinity of 56 remain almost unchanged from
atures [128]. With regard to 53, both the bpdc2- and the L40 link Zn 10 days (56a) to one month (56b). Their structures can be deter-
(II) ions into a ninefold interpenatrated network at 120 °C. In com- mined by single crystal X-ray diffractometer. The structural shrink
plex 53, the dihedral angle between the benzene rings of the bpdc2- for three complexes is reflected in alterative axial length and cell
ligands is approximately 47.78° due to the integral twist. In com- volume as well as the bending of rigid ligands toward opposite ori-
parison to 53, 54 synthesized at 100 °C shows a threefold interpe- entations, resulting in the change of structure from open to closed
natrated structure with the sql topology, although the framework forms (Fig. 21).
formula of 53 and 54 are the same. The differences of the struc- Recently, an exceptional guest-responsive thermal-expansion
tures should be attributed to the distorted configuration of L40 property in a flexible ultramicroporous CP originating from the
ligands with a dihedral angel of 34.98° between the benzene rings. twist conformational transformation of non-flexible ligand has
At 80 °C, another isomer 55 was formed. Complex 55 shows the been reported by Chen and co-workers [158]. Firstly, the authors
same overall sql net with a threefold interpenetration similar to selected a 3-(pyridin-4-yl)benzoic acid (L42) with a bent backbone
54 and a different combination fashion of Zn(II) and L40 ligands,
of which the corresponding dihedral angle between two benzene
rings of the L40 ligand has been changed to approximately 38.68°.
On the basis of the unremitting exploration of researchers for
CPs with diversified structures, some novel and attractive physical
properties, such as thermoresponsiveness, photosensitiveness, and
others, have been achieved [140]. The CPs featuring related proper-
ties usually have a reversible phase variation with a variety of lat-
tice parameters, which are usually triggered by external stimuli.
The responsive nature of CPs stems from the change of organic
and inorganic components upon external stimuli, such as the coor-
dination environment of SBUs as well as the versatility of organic
linkers such as twisting, rotating, bending, etc. By the change of
ligand conformation, various properties of CPs and the flexibility
of the framework with dynamic behaviors have attracted much
attention. However, the prediction of targeted and dynamic proper-
ties of CPs is still a great challenge. Up to now, numerous studies
focusing on the effect of metal clusters on the various flexibility of
framework have been made [151–155]. It should be noted that some
critical reviews have been given by Murdock and other researchers,
which put forward the effective approaches for constructing breath-
ing CPs using the flexibility of ligands [18,140,156]. In the following
part, we will concentrate on some representative cases to elaborate
the impact of non-flexible ligands with variable conformations on
the structures of CPs in order to avoid repetition.
The bending of non-flexible ligand backbones has been shown
to affect the structures of CPs. For example, the assembly of 1,4-
1H-1,2,3-triazole functionalized ligands, namely benzenedi(1H-
1,2,3-triazole) (L41), and Zn(II) yields a 3D CP, Cu(L41)(DMF)
1.2H2O (56), which shows breathing behaviors because of ligand
bending [157]. For 56, Cu(II) centers link the triazolate groups
and DMF molecules to form 1D chain. The adjacent chains are Fig. 21. The structural parameters of complexes 56, 56a, 56b, and structural
linked by L41 ligands to afford a 3D structure with 1D channels. deformations of the 1D channel upon the bending of ligand (the figure was
Phase transformation of 56 can be achieved upon exposure to air. reproduced from Ref. [157], with permission from the copyright holder).

Please cite this article in press as: N. Li et al., Conformation versatility of ligands in coordination polymers: From structural diversity to properties and
applications, Coord. Chem. Rev. (2018), https://doi.org/10.1016/j.ccr.2018.05.016
N. Li et al. / Coordination Chemistry Reviews xxx (2018) xxx–xxx 21

as the linker for constructing CPs and used the density functional 2.3. Impact of ligand flexibility on the structures of CPs
theory (DFT) method to calculate their potential energy profile
for different conformations (Fig. 22a). The result shows that the On the basis of the results mentioned above, it can be found that
conformation of ligands could be adjusted by external stimuli, ligands with similar structures can show distinct conformations
especially temperature changes. Such a conformation reversion due to their flexibility, which further affect the structures of corre-
provides an opportunity to construct temperature–guest–respon sponding CPs. The flexible feature of the ligands also makes them
sive frameworks. Through the assembly of L42 and Mn2+ ions, a easy to adapt various coordination modes of inorganic compo-
3D network with ant topology, [Mn(L42)2] (57), is obtained by nents, which may benefit the assembly of CPs in comparison with
L42 ligand linking trinuclear Mn2+ clusters. Powder XRD reveals that of rigid ligands. For example, Sun and co-workers have
that the crystallinity of 57 remains almost unchanged after heating selected two flexible ligands for the construction of CPs with CuI.
at 673 K for about 2 h. As expected, complex 57 can selectively By controlling the stoichiometry of reactants, six CPs were
responds to N,N-dimethylformamide, resulting in reversible con- obtained, namely [Cu2I2(L46)]n (63), [Cu4I4(L46)]n (64),
formation change of ligand and transformation of whole crystal [Cu6I6(L46)3]n (65), [Cu4I4(L47)]n (66), [Cu2I2(L47)2]n (67), and
lattice (Fig. 22b). This result indicates the potential of non- [Cu4I4(L47)3]n (68), (L46 = 1,4-di(2-methyl-imidazol-1-yl)butane,
flexible ligands in constructing CPs with versatile frameworks. L47 = 1,3-di(imidazol-1-yl)benzene) [162]. For L46 (dmimb) with
Distinctive stimuli-responsive functional units, such as AC@CA flexible C4 backbone, the corresponding CPs 63–65 reveal double-
and AN@NA bonds, can be incorporated into rigid ligands for the chain based 1D ribbon to 2D layers to triple-chain-based 1D ribbon
construction of responsive CPs. A series of photoreactive CPs have structures through the tuning of the CuI/dmimb ratio. While for
been reported by using ligands bearing AC@CA spacers complexes 66–68 constructed from rigid N-donor ligand L47, their
[139,159,160]. Four supramolecular isomers have been con- structures are 2D layers based on the Cu10I10 aggregate, binuclear
structed, namely, [Zn2(L43)(L44)2] (58–61). These isomers are metal–organic macrocycle, and 2D layers containing 1D infinite
obtained from the assembly of Zn(II) ions, 1,4-bis[2-(40 -pyridyl)et [Cu6I6]n columns and [Cu2I2]n ladders, respectively (Fig. 24). By
henyl]benzene (L43) and 4,40 -sulfonyldibenzoate (L44) by varying comparing the structures of the CPs, it is found that the L46 adopts
the reaction solvents [159]. These complexes possess similar 2D various conformations to adapt the diverse structures and arrange-
layer structures, which are constructed by the 1D chains based ments of the inorganic [CumIm] aggregates, while the relatively
on L44 and [Zn2(OOC)4] paddle-wheels linked by L43 as pillars. fixed conformation of L47 results in either 2D layers or discrete
Interestingly, complex 58 reveals a non-interpenetrating 2D layer structures.
structure, while 59 is a parallel interpenetrating network; Com- To further illustrate the influence of rigid and flexible ligands on
plexes 60 and 61 are obtained by the polyrotaxane of neighboring the structure of CPs, Li and co-workers have used biphenyl-3,30 ,5,
2D layers stemming from the conformational differences of L43: 50 -tetracarboxylic acid (L48) and different flexible N-donor ligands
the former shows trans–cis–trans conformation, whereas the L43 [L49 = 4,40 -bipyridine, L50 or L7 = 1,2-bis(4-pyridyl)ethane, L51 =
of latter exhibits trans–trans–trans conformation (Fig. 23a–d). The N,N0 -bis-(4-pyridyl-methyl) piperazine and L52 = 1,10 -(1,4-butane
L43 in 61 could undergo [2+2] cycloaddition reaction under UV diyl)bis(imidazole), Fig. 25a) to construct four Zn(II) coordination
irradiation due to the out-of-phase arrangement of the adjacent polymers, formulated as {[Zn2(L48)(L49)(H2O)2]2H2O}n (69), {[Zn
L43 ligands, which is congruent for photochemical dimerization. (L48)(L50)]2H2O}n (70), {[Zn2(L48)(L52)2]2H2O}n (71) and
Such a process affords a new 2D polyrotaxane complex 62 {[Zn2(L48)(L51)2]H2O}n (72) [163]. These N-donor ligands with
(Fig. 23e). This unique system represents a typical example of diverse flexibility could potentially direct the formation of various
the impact of cycloaddition reaction of the AC@CA moiety on topological networks. For 69, the L49 as a rigid linker shows a
the structural diversity of CPs. nearly coplanar conformation with a distance of 7.0875 Å between
As another example of this mothed, Zhao and co-workers have two N atoms, which are connected by L48 and Zn(II) ions to give a
reported the successful construction of a 3D self-penetrating 3D supramolecular polymer based on rigid layers (Fig. 25b). Ligand
framework using pillar ligand 3,30 -bis[2-(4-pyridyl)ethenyl]azo L50, which have a length of 9.2489 Å and consist of two methylene
benzene (L45) with AC@CA and AN@NA groups. Compared to groups and pyridyl rings, has been applied as relatively flexible
ligands bearing AC@CA spacers, L45 shows conformation diversity molecules to combine with L48 and Zn(II) ions to build a 44-sql
under UV light due to the synergetic conformation transformation sheets with two types of windows. The layers are far enough to
of AC@CA and AN@NA bonds. The dynamic photo-switching in allow the L50 to get the structure, leading to the twofold 2D-3D
the obtained CP gives rise to slight deformation of the structure framework of complex 70 (Fig. 25c). The L52 with the length of
[161]. 12.4809 Å features similar flexibility to L50, but the piperazine

Fig. 22. (a) Potential energy profile of L42 upon the pyridyl-ring rotation; (b) Reversible structure transformation between different phases for 57 (the pyridyl ring (green),
DMF molecule with space-filling mode is highlighted, and the figure was reproduced from Ref. [158], with permission from the copyright holder).

Please cite this article in press as: N. Li et al., Conformation versatility of ligands in coordination polymers: From structural diversity to properties and
applications, Coord. Chem. Rev. (2018), https://doi.org/10.1016/j.ccr.2018.05.016
22 N. Li et al. / Coordination Chemistry Reviews xxx (2018) xxx–xxx

Fig. 23. Representation of the rotaxane units for 60 (a) and for 61 (c) as well as polyrotaxanes structures for 60 (b) and for 61 (d); and the SC-SC transformation from
polyrotaxane 61 to polyrotaxane 62 under UV irradiation condition (the figures were reproduced from Ref. [19], with permission from the copyright holder).

Fig. 24. The diverse structures and corresponding reaction conditions for 63–68 (the figure was reproduced from Ref. [162], with permission from the copyright holder).

Please cite this article in press as: N. Li et al., Conformation versatility of ligands in coordination polymers: From structural diversity to properties and
applications, Coord. Chem. Rev. (2018), https://doi.org/10.1016/j.ccr.2018.05.016
N. Li et al. / Coordination Chemistry Reviews xxx (2018) xxx–xxx 23

fixed, which benefits the prediction of framework structure and


the targeted construction of CPs. Despite this fact, any subtle con-
formation of the non-flexible ligands can contribute to the diver-
sity of the corresponding CPs, even bring the CPs with unique
dynamic behavior and properties. Therefore, it can be concluded
that the versatile conformations of organic ligands are essential
to contribute to the diverse structural and functional characters
of CPs.

3. Properties and applications of CPs with conformation-


changing ligands

As mentioned in the introduction, the conformation versatility


of ligands is associated with the structural diversity of CPs, thus
contributing to their properties and applications. Originating from
the conformation transformation of ligands, structural transforma-
tion or dynamic behaviors could be introduced into CPs, which
may vary the properties of CPs dramatically and make them
promising for certain applications. In the following sections,
selected examples are presented to highlight the advantages of this
kind of CPs, mainly focusing on the potential applications in areas
of storage and separation as well as luminescence.

3.1. Storage and separation

The tunability and modularity of CPs stemming from a variety


of metal ions and organic linkers have resulted in numerous fasci-
nating structures with distinct properties. For the improvement of

Fig. 25. The comparison of length of the four different N-donor ligands and various
topological architectures constructed from these ligands (the figures were repro-
duced from Ref. [163], with permission from the copyright holder).

groups of L52 are too big to across the aperture of 2D layer which is
constructed by L48 and Zn(II) based cluster (Fig. 25d). So the com-
pound 71 exhibits a 3D self-penetrating pillar-layered framework
with two orientations of L52 ligands acting as pillars (Fig. 25e).
In 72, the more flexible L51 with a longer carbon chain (10.4454
Å) is employed to coordinate with L48 ligands and Zn(II) ions to
form a 3D framework with hexagonal channels (Fig. 25f), which
allow the symmetry-related frameworks to penetrate the hexago-
nal channels resulting in a threefold interpenetrated networks
(Fig. 25g).
By comparing the results mentioned above, it is obvious that
the flexible ligands are easier for the construction of CPs with var-
ious structures owing to their relatively more variable conforma-
tions. However, the conformations of these ligands might be
responsive toward various factors (reaction conditions, reagents,
etc.), making the prediction and control of the structure of corre-
sponding CPs difficult. Typically, the conformations of flexible
Fig. 26. (a) Dynamic photo-switching in 73 gives rise to instantly reversible CO2
ligands can be regulated by controlling the reaction conditions or adsorption; (b) CO2 adsorption isotherms of 73 at 303 K with light (red) or without
by introducing the ligands into a well-defined coordination scaf- light (black), and unfiltered light switching environment (blue) (the figure was
fold. As to the non-flexible ligands, the conformation is relatively reproduced from Ref. [166], with permission from the copyright holder).

Please cite this article in press as: N. Li et al., Conformation versatility of ligands in coordination polymers: From structural diversity to properties and
applications, Coord. Chem. Rev. (2018), https://doi.org/10.1016/j.ccr.2018.05.016
24 N. Li et al. / Coordination Chemistry Reviews xxx (2018) xxx–xxx

specified properties of CPs, functional groups can be introduced swing adsorption (LISA) to present a dual stimuli-responsive CP
through the rational design and construction of ligands, such as (mPCNs), in which the Fe3O4 MNPs have been integrated with
the enhancement of CO2 sorption performances of CPs through PCN-250 containing light-responsive 3,30 ,5,50 -azo-benzene
the introduction of polar groups [164,165]. As an alternating tetracarboxylic acid (ABTC, L54) ligands [167]. The CO2 adsorption
approach, modulation of the properties can be achieved through measurements show that all mPCNs have dynamic switching of gas
structural optimization of CPs based on the conformational trans- uptake and release in LISA, MISA and the combination of MISA and
formation of ligands. As mentioned in the second part, the variable LISA (denoted as MaLISA). Particularly, in the MaLISA, each of the
conformation of ligand could affect the pore structure and pore light-induced bending of ABTC and magnetic-induced localized
surface chemistry of CPs, which can be utilized to tune the storage heat is effective for CO2 release. Moreover, the cooperative trigger
and separation performance of the CPs. However, related works is more available in comparison with LISA and MISA alone (Fig. 27).
have rarely been reviewed. Hence, we focus on the distinguished In addition, Guo and co-workers reported a photoswitchable
storage and separation properties stemming from the diverse con- DMOF (diarylethene MOF), Zn(L55)(bpdc)solvents (74) (L55 = dia
formations of ligands. rylethene derivative and H2bpdc = biphenyl-4,40 -dicarboxylic acid)
Currently, the huge energy costs prevent widespread viability of [168]. In complex 74, the Zn(II) center is four-coordinated by two
post-combustion CO2 capture technology. 40% of the power plant bpdc2- and two nitrogen atoms from L55 ligand to form a tetrahe-
capacity has been consumed for the liberation of CO2 from capture dral geometry. The topology of the framework is four-connected
media. Hence, it is a very significant research to develop new dia network with fivefold interpenetration and the accessible vol-
releasing triggers using renewable energy sources. Hill and co- ume of 74 is 36.9% of the crystal cell volume after the removal of
workers reported a low-energy photoresponsive MOF, Zn(AzDC) solvent molecules. Complex 74 exhibits similar photochromism
(L53)0.5 (73), in which the azobenzene derivative dicarboxylate behavior to the L55 ligand with the color change of CP from buff
(AzDC) and trans-1,2-bis(4-pyridyl)ethylene (4,40 -BPE, L53) have to blue, which is directly related to the structural transformation
been utilized as organic ligands [166]. As shown in Fig. 26, complex of the diarylethene unit from the open-ring state to the closed-
73 reveals dynamic behavior for CO2 adsorption, which is attribu- ring state under UV irradiation. After UV irradiation, the CO2
ted to the reversible framework structure transformation under adsorption capacity of 74 reach 20.1 cm3 g1 (298 K), while that
light irradiation or heating. Such behavior originates from the without UV irradiation is 5.0 cm3 g1. What’s more, the complex
trans-cis isomerization of the AzDC ligand accompanied by the con- 74 has a record CO2 desorption capacity of 75% under static irradi-
formation change of 4,40 -BPE. Both the pore volume and pore sur- ation and 76% under dynamic irradiation. Then they studied the
face energy can be modified through the structural transformation. C2H2/C2H4 separation of 74 [169]. It shows high C2H2/C2H4 separa-
Since the trans-cis isomerization and structural change can be tion of 47.1 at 195 K and 100 kPa. When the sample is irradiated
revealed by X-ray diffraction, authors claimed that the dynamic under UV, the adsorption selectivity of C2H2 over C2H4 is 3.0 under
behavior is localized. The photoresponsive approach shows a the same conditions, which makes it quite potential for separation
potential route for low energy-cost CO2 capture and release. applications (Scheme 8).
As the magnetic nanoparticles (MNPs) can serve as ‘nanohea- In recent years, some flexible CPs with gated adsorption effect
ters’ to generate heat locally under an alternating magnetic field, have been discovered, which promoted the exploration of gate
the magnetic induction swing adsorption (MISA) has been con- opening and gate closing process on the gas adsorption and separa-
firmed to be effective for the release of adsorption gas from MOF tion applications of CPs. However, a portion of CPs possess similar
adsorbents. In order to pursue the low-energy approach to regen- gate opening pressures for distinct gases [170], which give rise to
erate CPs, Hill and co-workers combine MISA and light induction an enormous challenge for gas separation with this kind of CPs,

Fig. 27. Schematic illustration of CO2 release from dual stimuliresponsive mPCNs collaboratively driven by both UV light and an alternating magnetic field in a MaLISA
process (the figure was reproduced from Ref. [167], with permission from the copyright holder).

Please cite this article in press as: N. Li et al., Conformation versatility of ligands in coordination polymers: From structural diversity to properties and
applications, Coord. Chem. Rev. (2018), https://doi.org/10.1016/j.ccr.2018.05.016
N. Li et al. / Coordination Chemistry Reviews xxx (2018) xxx–xxx 25

Scheme 8. Representation of separation switch upon UV and visible light for 74


(the scheme was reproduced from Ref. [169], with permission from the copyright
holder).

especially for CO2 and C2H2 with analogical adsorption parameters,


sizes, and shapes. As a representative example for selective adsorp-
tion of CO2 over C2H2 using the gate effect of CPs, Kitagawa and co-
workers reported the design and synthesis of a flexible porous CP
embodied by [Mn(bdc)(L56)]DMF (75) (H2bdc = 1,4-benzenedicar
boxylic acid, L56 = 1,2-di(4-pyridyl)ethylene) [171]. Complex 75
has a twofold interpenetrated box motif with 0-D pores occupying
by DMF molecule (Fig. 28a). The desolvated 75 (75a) is obtained
by heating under vacuum. With regard to 75a, it shows gate opening
type adsorption behavior with sudden uptake for C2H2 over CO2 at
low temperature (195 K) accompanied by structural distortion
through slant of the L56 pillars and warp of the phenylene rings of Fig. 29. (a) Photochromic reaction of the diarylethene derivative (L58) under UV
bdc2- ligands (Fig. 28b). Notably, CO2 is located in end-on orientation and visible light; (b) Photocontrolled tunable luminescence performance of 77 (the
figure was reproduced from Ref. [168], with permission from the copyright holder).
along the channel in 195 K. In contrast, C2H2 has repulsive interac-
tion with the channel. For 75a, the gate opening pressure and the
selectivity of separation for C2H2/CO2 are increased when the tem- addition, in order to verify the mechanism of selective gate opening
perature is raised (Fig. 28d), which leads to a preferable selective behavior, [2+2] photodimerization on 75a can be realized to obtain a
adsorption of CO2 over C2H2 at near ambient temperature. In novel photodimerized CP [Mn2(bdc)2(L57)] (76) (L57 = regio-cis,

Fig. 28. The box structural motif of 75 (a), 75a (b), and 76 (c) with twofold interpenetrated structure (DMF: purple spheres), adsorption isotherms of C2H2 (bottom) for 75a at
220, 240, 246, and 268 K, and gas sorption isotherms for CO2 and C2H2 for 75a (blue) and 76 (red) at 273 K (the figures were reproduced from Ref. [171], with permission from
the copyright holder).

Please cite this article in press as: N. Li et al., Conformation versatility of ligands in coordination polymers: From structural diversity to properties and
applications, Coord. Chem. Rev. (2018), https://doi.org/10.1016/j.ccr.2018.05.016
26 N. Li et al. / Coordination Chemistry Reviews xxx (2018) xxx–xxx

trans, trans-tetrakis(4-pyridyl)cyclobutane, Fig. 28c). For 76, CO2 Investigations about ligand induced fluorescence of CPs, particu-
molecules cannot interact with phenylene rings efficiently due to larly without ligand conformation, have been widely studied and
the tilting of phenylene rings, which makes 76 exhibit almost no systematically reviewed [14]. In contrast, targeted modulation of
gas selectivity for the separation of C2H2 and CO2 (Fig. 28e). ligand conformation for photophysical properties, has been
reported very recently. The CPs, whose photophysical properties
3.2. Luminescence are responsive to stimuli due to the transformation of ligand con-
formation, can be applied as optical sensors and switching devices,
The luminescence of CPs has been widely investigated owing to etc. As mentioned before, the diarylethene ligands with photo-
its potential applications in the fields of sensing and lighting [172]. induced cyclization and olefinic ligands with [2+2] cycloaddition
are the representative photo-switching motifs. In addition to the
impact on the structure of CPs based on the conformation transfor-
mation of these motifs, photophysical properties of the CPs could
also be tuned based on the distinct states of the ligand. These
unique characteristics make the related CPs promising candidates
for optoelectronic materials.
Diarylethene (1,3,5-hexatriene) and cyclohexadiene (L58) were
found to be transformed mutually under the irradiation of UV and
visible lights, which revealed distinct emission properties, respec-
tively. Guo and coworkers reported a photo-switching diary-
lethene CP Zn(L58)(bpdc)solvents (77) constructed by a
diarylethene derivative, which can convert between closed-ring
and open-ring isomers induced by photo-stimulation [168]. Com-
plex 77 exhibits similar variable emission properties to that of
the ligand (Fig. 29a): The fluorescence emission of framework
occurs bathochromic shift under the stimulus of UV lights, and
during irradiation of visible lights, hypsochromic shift of emission
wavelength can be observed. Under UV or visible lights, the color of
crystal is switched between yellow and dark blue. The emission
Scheme 9. The structure of L59 in the open and closed forms (top); (middle) the intensity of the closed-ring form is lower, but can be recovered
schematic presentation of 78 in the on and off states; (bottom) simplified diagram to original intensity by visible light (Fig. 29b). These properties
of the transitions responsible for fluorescence and energy transfer in 78 (the scheme make the complex a potential candidate for light induced
was reproduced from Ref. [173], with permission from the copyright holder).
switching.

Fig. 30. (a) View of the reversible photo/thermoswitchable cyclo-addition for 79; (b) The structure of 79 with binodal {628310}{658}2 net; (c) The CIE diagram of UV
irradiation time-dependence PL properties for 79 (the figure was reproduced from Ref. [174], with permission from the copyright holder).

Please cite this article in press as: N. Li et al., Conformation versatility of ligands in coordination polymers: From structural diversity to properties and
applications, Coord. Chem. Rev. (2018), https://doi.org/10.1016/j.ccr.2018.05.016
N. Li et al. / Coordination Chemistry Reviews xxx (2018) xxx–xxx 27

Another representative framework with similar ligands the conformation versatility of ligands induced structural diversity
[Zn2(L60)(L59)0.85(DEF)1.15](DEF)5.15(H2O)7.25 (78) (DEF = N,N-die in CPs according to the analysis of enormous reported examples,
thylformamide), was reported by Shustova and co-workers [173]. and illustrate the influence of the versatility of ligands on the
L59 (bis(5-pyridyl-2-methyl-3-thienyl)-cyclopentene) was immo- structures of CPs from the aspects of ligands conformation versatil-
bilized into a two-dimensional porphyrin framework constructed ity. By reviewing the achievements in this field, it is noticed that
by Zn2(ZnL60) (L60 = tetrakis(4-carboxyphenyl)-porphyrin) 2D ligands with versatile conformations do contribute to the diversity
framework, and the z-axis of as-synthesized framework was of CPs in the perspective of framework structure and topology.
extended compared to the original Zn2(ZnL60), resulting in the These examples also benefit the summary of the construction prin-
2D-3D framework transformation caused by L59 insertion. The ciples of CPs. More importantly, the research interest of ligands
3D framework shows red-color emission, and the irradiation of with versatile conformations does not fade along with the in-
the pillared framework can be quenched or enhanced with incident depth research of properties and applications of CPs. This type of
lights. The quenching process can be attributed with energy trans- ligands leads to the discovery of CPs with promising applications
fer governed by photo-switchable spectral overlap between in various fields. Accordingly, they deserve more attention in fur-
cyclized L59 and metal-porphyrin layers (Scheme 9). ther research.
On the other hand, studies about [2+2] cycloaddition between Furthermore, it should be noted that there are still many chal-
alkenes have been investigated in depth over the last decades, lenges in this field. From the reviewed examples, it can be found
and CPs is capable for [2+2] photo-triggered cycloaddition with that the conformation control of most flexible ligands during the
the elaborately designed conformation as well as limit motion of self-assembly process of CPs is difficult, which may make the pre-
reaction constituents. However, reports of fluorescence properties cisely targeted construction of CPs difficult. For conformation con-
of CPs with [2+2] cycloaddition are relatively limited. Luo and co- trol of flexible ligands, some unique methods from organic
workers reported a dual-color green-to-blue fluorescence frame- synthesis or crystal engineering will be useful. For example, the
work [Zn(COO)(L61)0.5(L62)0.5] (79, L61 = 1,4-bis[2-(pyridin-4-yl) conformation of a flexible ligand might be fixed through the rever-
ethenyl]benzene, L62 = 2-aminoterephthalic acid) [174]. The as- sible chemical bonding or weak interactions with a rigid template,
synthesized structure can transform into cycloaddition phase with which is commonly used in organic synthesis and crystal engineer-
UV irradiation and the original structure can be recovered by heat- ing. Then CPs with desired conformation of ligands can be obtained
ing the latter phase of framework. The original phase exhibits through a post-synthesis removal of template after the assembly
green emission while the UV-irradiated phase exhibits blue process of CPs. This strategy could also be effective for the modu-
emission, which is also stronger than former fluorescence lation of ligand conformation in CPs to obtain the desired struc-
(Fig. 30). Notably, the turn-on and turn-off luminescence can be tures and properties. This method may result in CPs with high
reversible in a photo/thermo-manner. selectivity toward the used templates, extending the applications
Apart from structural fluorescence changes, the [2+2] cycload- of the corresponding CPs into useful sensors.
dition conversion in framework can also change guest discrimina- In conclusion, a brighter future filled with more interesting CPs
tion properties. Lee and Vittal et al. have reported a series of 2D CPs can be expected with more research effort. We believe this review
[Zn2(L63)(L64)2] (L63 = 1,4-bis[2-(40 -pyridyl)ethenyl]benzene, L64 highlights the elements contributing to the diversity of CPs, and
= 4,40 -sulfonyldibenzoate) [159]. Four types of frameworks were provides a useful reference in the future.
obtained on the basis of different synthesis conditions. Interest-
ingly, two complexes with similar entangled frameworks but dif- Acknowledgments
ferent ligand conformations (trans-cis-trans and trans-trans-trans,
respectively) reveal distinct properties: L63 in the trans-cis-trans This work was supported by the 973 Program of China
species undergoes [2+2] cycloaddition and a new polyrotaxane (2014CB845601) and the National Natural Science Foundation of
framework can be obtained, which is photoinert in trans–trans– China (21421001, 21531005, 91622111, and 21671112).
trans species. The emission of the trans-cis-trans species as well
as the transformed framework can be selectively quenched by
(2,4-dinitrophenylhydrazine), and the former species is more sen- References
sitive. Above all, the fluorescence properties of CPs can be con-
[1] M. Rubio-Martinez, C. Avci-Camur, A.W. Thornton, I. Imaz, D. Maspoch, M.R.
trolled by transformable ligands, and these studies have provided Hill, Chem. Soc. Rev. 46 (2017) 3453–3480.
instruction for synthesizing a variety of photo-stimulated func- [2] Q. Yang, Q. Xu, H.-L. Jiang, Chem. Soc. Rev. 46 (2017) 4774–4808.
tional materials. [3] S. Zhang, Q. Yang, X. Liu, X. Qu, Q. Wei, G. Xie, S. Chen, S. Gao, Coord. Chem.
Rev. 307 (2016) 292–312.
[4] L. Wang, Y. Han, X. Feng, J. Zhou, P. Qi, B. Wang, Coord. Chem. Rev. 307 (2016)
361–381.
4. Summary and outlook [5] S.-N. Zhao, X.-Z. Song, S.-Y. Song, H.-J. Zhang, Coord. Chem. Rev. 337 (2017)
80–96.
[6] H.-Y. Wang, L. Cui, J.-Z. Xie, C.F. Leong, D.M. D’Alessandro, J.-L. Zuo, Coord.
As an interesting field showing rapid and expanded develop- Chem. Rev. 345 (2017) 342–361.
ment, the investigation of coordination polymers has changed its [7] Z.-P. Ni, J.-L. Liu, M.N. Hoque, W. Liu, J.-Y. Li, Y.-C. Chen, M.-L. Tong, Coord.
Chem. Rev. 335 (2017) 28–43.
focus from the exploration and discovery of unique structures to [8] Y. Liu, A.J. Howarth, N.A. Vermeulen, S.-Y. Moon, J.T. Hupp, O.K. Farha, Coord.
the performance targeted construction and application of the Chem. Rev. 346 (2017) 101–111.
materials. Though the attention of research may have been chan- [9] T.-L. Easun, F. Moreau, Y. Yan, S. Yang, M. Schröder, Chem. Soc. Rev. 46 (2017)
239–274.
ged along with the development of this field, the illustration and [10] M. O’Keeffe, O.M. Yaghi, Chem. Rev. 112 (2012) 675–702.
utilization of the principles for desired CP structures and properties [11] D. Zhao, D.J. Timmons, D. Yuan, H.-C. Zhou, Acc. Chem. Res. 44 (2011) 123–
are still one of the critical challenges. Despite the numerous previ- 133.
[12] S. Qiu, G. Zhu, Coord. Chem. Rev. 253 (2009) 2891–2911.
ous reviews aiming at the illustration of principles for structural
[13] R.A. Agarwal, N.K. Gupta, Coord. Chem. Rev. 332 (2017) 100–121.
modulation of CPs based on specific class of organic ligands, our [14] W.P. Lustig, S. Mukherjee, N.D. Rudd, A.V. Desai, J. Li, S.K. Ghosh, Chem. Soc.
contribution herein focuses on the structural diversity of CPs based Rev. 46 (2017) 3242–3285.
on conformation versatility of organic ligands, which has rarely [15] P. Falcaro, R. Ricco, A. Yazdi, I. Imaz, S. Furukawa, D. Maspoch, R. Ameloot, J.D.
Evans, C.J. Doonan, Coord. Chem. Rev. 307 (2016) 237–254.
been systematically summarized. Bearing this in mind, our aims [16] M. Zhang, Z.-Y. Gu, M. Bosch, Z. Perry, H.-C. Zhou, Coord. Chem. Rev. 293–294
in this review are to present a systematic summary to generalize (2015) 327–356.

Please cite this article in press as: N. Li et al., Conformation versatility of ligands in coordination polymers: From structural diversity to properties and
applications, Coord. Chem. Rev. (2018), https://doi.org/10.1016/j.ccr.2018.05.016
28 N. Li et al. / Coordination Chemistry Reviews xxx (2018) xxx–xxx

[17] B. Seoane, S. Castellanos, A. Dikhtiarenko, F. Kapteijn, J. Gascon, Coord. Chem. [67] M.-D. Zhang, L. Qin, X.-Q. Yao, Q.-X. Yang, Z.-J. Guo, H.-G. Zheng,
Rev. 307 (2016) 147–187. CrystEngComm 15 (2013) 7354–7359.
[18] C.R. Murdock, B.C. Hughes, Z. Lu, D.M. Jenkins, Coord. Chem. Rev. 258–259 [68] T.D. Keene, M.J. Murphy, J.R. Price, N.F. Sciortino, P.D. Southon, C.J. Kepert,
(2014) 119–136. Dalton Trans. 43 (2014) 14766–14771.
[19] W.-G. Lu, Z.-W. Wei, Z.-Y. Gu, T.-F. Liu, J. Park, J. Park, J. Tian, M.-W. Zhang, Q. [69] C. Li, W. Qiu, W. Shi, H. Song, G. Bai, H. He, J. Li, M.J. Zaworotko,
Zhang, T. Gentle, M. Bosch, H.-C. Zhou, Chem. Soc. Rev. 43 (2014) 5561–5593. CrystEngComm 14 (2012) 1929–1932.
[20] Y. He, B. Li, M. O’Keeffe, B. Chen, Chem. Soc. Rev. 43 (2014) 5618–5656. [70] V.C. Perry, G.J. Kravtsov, M.J. McManus, Zaworotko, J. Am. Chem. Soc. 129
[21] Y. Bai, Y. Dou, L.-H. Xie, W. Rutledge, J.-R. Li, H.-C. Zhou, Chem. Soc. Rev. 45 (2007) 10076–10077.
(2016) 2327–2367. [71] S. Zang, Y. Su, Y. Li, Z. Ni, Q. Meng, Inorg. Chem. 45 (2006) 174–180.
[22] M. Du, C.-P. Li, C.-S. Liu, S.-M. Fang, Coord. Chem. Rev. 257 (2013) 1282–1305. [72] S. Zang, Y. Su, Y. Li, H. Zhu, Q. Meng, Inorg. Chem. 45 (2006) 2972–2978.
[23] Z.-J. Lin, J. Lu, M. Hong, R. Cao, Chem. Soc. Rev. 43 (2014) 5867–5895. [73] Y.-B. Dong, Y.-Y. Jiang, J. Li, J.-P. Ma, F.-L. Liu, B. Tang, R.-Q. Huang, S.R. Batten,
[24] Y.-Q. Lan, H.-L. Jiang, S.-L. Li, Q. Xu, Adv. Mater. 23 (2011) 5015–5020. J. Am. Chem. Soc. 129 (2007) 4520–4521.
[25] Z.-J. Lin, T.-F. Liu, B. Xu, L.-W. Han, Y.-B. Huang, R. Cao, CrystEngComm 13 [74] Y. Qi, Y. Che, F. Luo, S.R. Batten, Y. Liu, J. Zheng, Cryst. Growth Des. 8 (2008)
(2011) 3321–3324. 1654–1662.
[26] J.F. Eubank, L. Wojtas, M.R. Hight, T. Bousquet, V.C. Kravtsov, M. Eddaoudi, J. [75] L.-P. Zhang, J. Yang, J.-F. Ma, Z.-F. Jia, Y.-P. Xie, G.-H. Wei, CrystEngComm 10
Am. Chem. Soc. 133 (2011) 17532–17535. (2008) 1410–1420.
[27] N.W. Ockwig, O. Delgado-Friedrichs, M. O’Keeffe, O.M. Yaghi, Acc. Chem. Res. [76] L.-H. Cao, H.-Y. Li, S.-Q. Zang, H.-W. Hou, T.C.W. Mak, Cryst. Growth Des. 12
38 (2005) 176–182. (2012) 4299–4301.
[28] A. Phan, C.J. Doonan, F.J. Uribe-Romo, C.B. Knobler, M. O’Keeffe, O.M. Yaghi, [77] M.-L. Ma, C. Ji, S.-Q. Zang, Dalton Trans. 42 (2013) 10579–10586.
Acc. Chem. Res. 43 (2010) 58–67. [78] A. Karmakar, I. Goldberg, CrystEngComm 13 (2011) 350–366.
[29] M. O’Keeffe, M.A. Peskov, S.J. Ramsden, O.M. Yaghi, Acc. Chem. Res. 41 (2008) [79] M. Wu, F. Jiang, W. Wei, Q. Gao, Y. Huang, L. Chen, M. Hong, Cryst. Growth
1782–1789. Des. 9 (2009) 2559–2561.
[30] J.-L. Du, T.-L. Hu, S.-M. Zhang, Y.-F. Zeng, X.-H. Bu, CrystEngComm 10 (2008) [80] R. Ding, C. Huang, J. Lu, J. Wang, C. Song, J. Wu, H. Hou, Y. Fan, Inorg. Chem. 54
1866–1874. (2015) 1405–1413.
[31] T.K. Kim, K.J. Lee, M. Choi, N. Park, D. Moon, H.R. Moon, New J. Chem. 37 [81] D. Tian, R.-Y. Chen, J. Xu, Y.-W. Li, X.-H. Bu, APL Mater. 2 (2014) 124111.
(2013) 4130–4139. [82] D. Tian, Y. Li, R.-Y. Chen, Z. Chang, G. Wang, X.-H. Bu, J. Mater. Chem. A 2
[32] X.-Q. Yao, M.-D. Zhang, J.-S. Hu, Y.-Z. Li, Z.-J. Guo, H.-G. Zheng, Cryst. Growth (2014) 1465–1470.
Des. 11 (2011) 3039–3044. [83] D. Tian, Q. Chen, Y. Li, Y.-H. Zhang, Z. Chang, X.-H. Bu, Angew. Chem. Int. Ed.
[33] H.-L. Tsai, C.-I. Yang, W. Wernsdorfer, S.-H. Huang, S.-Y. Jhan, M.-H. Liu, G.-H. 53 (2014) 837–841.
Lee, Inorg. Chem. 51 (2012) 13171–13180. [84] D. Tian, J. Xu, Z.-J. Xie, Z.-Q. Yao, D.-L. Fu, Z. Zhou, X.-H. Bu, Adv. Sci. 3 (2015)
[34] X.-X. Wang, B. Yu, K. Van Hecke, G.-H. Cui, RSC Adv. 4 (2014) 61281–61289. 1500283–1500289.
[35] J.-X. Yang, Y.-Y. Qin, J.-K. Cheng, Y.-G. Yao, Cryst. Growth Des. 14 (2014) [85] J.F. Eubank, H. Mouttaki, A.J. Cairns, Y. Belmabkhout, L. Wojtas, R. Luebke, M.
1047–1056. Alkordi, M. Eddaoudi, J. Am. Chem. Soc. 133 (2011) 14204–14207.
[36] P.P. Bag, X.-S. Wang, R. Cao, Dalton Trans. 44 (2015) (1962) 11954–11961. [86] J.F. Eubank, F. Nouar, R. Luebke, A.J. Cairns, L. Wojtas, M. Alkordi, T. Bousquet,
[37] X.-D. Zhu, Y. Li, W.-X. Zhou, R.-M. Liu, Y.-J. Ding, J. Lu, D.M. Proserpio, M.R. Hight, J. Eckert, J.P. Embs, P.A. Georgiev, M. Eddaoudi, Angew. Chem., Int.
CrystEngComm 18 (2016) 4530–4537. Ed. 51 (2012) 10099–10103.
[38] T. Basu, H.A. Sparkes, M.K. Bhunia, R. Mondal, Cryst. Growth Des. 9 (2009) [87] Z. Guo, R. Cao, X. Wang, H. Li, W. Yuan, G. Wang, H. Wu, J. Li, J. Am. Chem. Soc.
3488–3496. 131 (2009) 6894–16805.
[39] S.W. Jaros, M.F.C. Guedes da Silva, M. Florek, M.C. Oliveira, P. Smolenski, A.J.L. [88] L.-L. Liang, J. Zhang, S.-B. Ren, G.-W. Ge, Y.-Z. Li, H.-B. Du, X.-Z. You,
Pombeiro, A.M. Kirillov, Cryst. Growth Des. 14 (2014) 5408–5417. CrystEngComm 12 (2010) 2008–2010.
[40] K. Wang, X.-C. Yi, X. Wang, X.-B. Li, E.-Q. Gao, Dalton Trans. 42 (2013) 8748– [89] T.-F. Liu, J. Lue, X. Lin, R. Cao, Chem. Commun. 46 (2010) 8439–8441.
8760. [90] Y.-Q. Lan, S.-L. Li, H.-L. Jiang, Q. Xu, Chem. Eur. J. 18 (2012) 8076–8083.
[41] S. Wang, R. Yun, Y. Peng, Q. Zhang, J. Lu, J. Dou, J. Bai, D. Li, D. Wang, Cryst. [91] L. Zhou, Y.-S. Xue, J. Zhang, H.-B. Du, X.-Z. You, CrystEngComm 15 (2013)
Growth Des. 12 (2012) 79–92. 6199–6206.
[42] T. Cao, Y. Peng, T. Liu, S. Wang, J. Dou, Y. Li, C. Zhou, D. Li, J. Bai, [92] F.-Y. Yi, S. Dang, W. Yang, Z.-M. Sun, CrystEngComm 15 (2013) 8320–8329.
CrystEngComm 16 (2014) 10658–10673. [93] Y. Ling, J. Jiao, M. Zhang, H. Liu, D. Bai, Y. Feng, Y. He, CrystEngComm 18
[43] G.-C. Xu, Q. Hua, T.-A. Okamura, Z.-S. Bai, Y.-J. Ding, Y.-Q. Huang, G.-X. Liu, W.- (2016) 6254–6261.
Y. Sun, N. Ueyama, CrystEngComm 11 (2009) 261–270. [94] Z.-P. Deng, L.-N. Zhu, S. Gao, L.-H. Huo, S.W. Ng, Cryst. Growth Des. 8 (2008)
[44] C.-Y. Li, C.-S. Liu, J.-R. Li, X.-H. Bu, Cryst. Growth Des. 7 (2007) 286–295. 3277–3284.
[45] W. Zhao, Y. Song, T. Okamura, J. Fan, W.-Y. Sun, N. Ueyama, Inorg. Chem. 44 [95] Z.-P. Deng, H.L. Qi, L.-H. Huo, S.W. Ng, H. Zhao, S. Gao, Dalton Trans. 39 (2010)
(2005) 3330–3336. 10038–10050.
[46] X. Wang, L. Zhang, J. Yang, F. Liu, F. Dai, R. Wang, D. Sun, J. Mater. Chem. A 3 [96] B. Wu, J. Liang, Y. Zhao, M. Li, S. Li, Y. Liu, Y. Zhang, X.-J. Yang, CrystEngComm
(2015) 12777–12785. 12 (2010) 2129–2134.
[47] Q.-Y. Yang, K. Li, J. Luo, M. Pan, C.-Y. Su, Chem. Commun. 47 (2011) 4234–4236. [97] Z.-P. Deng, L.-H. Huo, H.-L. Qi, L.-N. Zhu, H. Zhao, S. Gao, CrystEngComm 13
[48] R.-M. Wen, S.-D. Han, G.-J. Ren, Z. Chang, Y.-W. Li, X.-H. Bu, Dalton Trans. 44 (2011) 4218–4227.
(2015) 10914–10917. [98] K. Suman, Rajnikant, V.K. Gupta, M. Sarkar, Dalton Trans. 42 (2013) 8492–
[49] C.-I. Yang, Y.-T. Song, Y.-J. Yeh, Y.-H. Liu, T.-W. Tseng, K.-L. Lu, CrystEngComm 8497.
13 (2011) 2678–2686. [99] P.-C. Cheng, C.-W. Yeh, W. Hsu, T.-R. Chen, H.-W. Wang, J.-D. Chen, J.-C. Wang,
[50] W.M. Bloch, R. Babarao, M.R. Hill, C.J. Doonan, C.J. Sumby, J. Am. Chem. Soc. Cryst. Growth Des. 12 (2012) 943–953.
135 (2013) 10441–10448. [100] B. Chakraborty, P. Halder, T.K. Paine, Dalton Trans. 40 (2011) 3647–3654.
[51] B. Manna, A.V. Desai, S.K. Ghosh, Dalton Trans. 45 (2016) 4060–4072. [101] E. Lee, J.-Y. Kim, S.S. Lee, K.-M. Park, Chem. Eur. J. 19 (2013) 13638–13645.
[52] X.-L. Li, G.-Z. Liu, L.-Y. Xin, L.-Y. Wang, CrystEngComm 14 (2012) 5757–5760. [102] S.-N. Wang, R. Sun, X.-S. Wang, Y.-Z. Li, Y. Pan, J. Bai, M. Scheer, X.-Z. You,
[53] L. Carlucci, G. Ciani, D.M. Proserpio, Chem. Commun. (1999) 449–450. CrystEngComm 9 (2007) 1051–1061.
[54] L. Carlucci, G. Ciani, D.M. Proserpio, S. Rizzato, Chem. Commun. (2000) 1319– [103] Q. Zhang, H. Zhang, S. Zeng, D. Sun, C. Zhang, Chem. Asian J. 8 (2013) 1985–
1320. 1989.
[55] E. Suresh, M.M. Bhadbhade, CrystEngComm 13 (2001) 1–3. [104] J.-R. Li, X.-H. Bu, Eur. J. Inorg. Chem. (2008) 27–40.
[56] L. Carlucci, G. Ciani, D.M. Proserpio, S. Rizzato, CrystEngComm 5 (2003) 190– [105] X.-H. Bu, W. Chen, S.-L. Lu, R.-H. Zhang, D.-Z. Liao, W.-M. Bu, M. Shionoya, F.
199. Brisse, J. Ribas, Angew. Chem. Int. Ed. 40 (2001) 3201–3203.
[57] S. Hu, A.-J. Zhou, Y.-H. Zhang, S. Ding, M.-L. Tong, Cryst. Growth Des. 6 (2006) [106] X.-H. Bu, W. Chen, M. Du, K. Biradha, W.-Z. Wang, R.-H. Zhang, Inorg. Chem.
2543–2550. 41 (2002) 437–439.
[58] J. Lewinski, M. Dranka, W. Bury, W. Sliwinski, I. Justyniak, J. Lipkowski, J. Am. [107] X.-H. Bu, W. Weng, M. Du, W. Chen, J.-R. Li, R.-H. Zhang, L.-J. Zhao, Inorg.
Chem. Soc. 129 (2007) 3096–3098. Chem. 41 (2002) 1007–1010.
[59] W.-H. Zhang, Y.-L. Song, Z.-H. Wei, L.-L. Li, Y.-J. Huang, Y. Zhang, J.-P. Lang, [108] T.-L. Hu, J.-R. Li, Y.-B. Xie, X.-H. Bu, Cryst. Growth Des. 6 (2006) 648–655.
Inorg. Chem. 47 (2008) 5332–5346. [109] J.-R. Li, X.-H. Bu, R.-H. Zhang, Inorg. Chem. 43 (2004) 237–244.
[60] P. Kanoo, C. Madhu, G. Mostafa, T.K. Maji, A. Sundaresan, S.K. Pati, C.N.R. Rao, [110] J.-R. Li, W.-T. Chen, M.-L. Tong, G.-C. Guo, Y. Tao, Q. Yu, W.-C. Song, X.-H. Bu,
Dalton Trans. (2009) 5062–5064. Cryst. Growth Des. 8 (2008) 2780–2792.
[61] N. de la Pinta, G. Madariaga, L. Lezama, M.L. Fidalgo, R. Cortes, Eur. J. Inorg. [111] Y.-Q. Sun, J.-C. Zhong, L. Ding, Y.-P. Chen, Dalton Trans. 44 (2015) (1859)
Chem. (2010) 3491–3497. 11852–11861.
[62] N. de la Pinta, S. Martin, M.K. Urtiaga, M.G. Barandika, M.I. Arriortua, L. [112] Y.-Q. Lan, S.-L. Li, K.-Z. Shao, X.-L. Wang, D.-Y. Du, Z.-M. Su, D.-J. Wang, Cryst.
Lezama, G. Madariaga, R. Cortes, Inorg. Chem. 49 (2010) 10445–10454. Growth Des. 8 (2008) 3490–3492.
[63] T.K. Maji, R. Matsuda, S. Kitagawa, Nat. Mater. 6 (2007) 142–148. [113] X. He, X.-P. Lu, Y.-Y. Tian, M.-X. Li, S. Zhu, F. Xing, R.E. Morris, CrystEngComm
[64] N. De la Pinta, M.L. Fidalgo, L. Lezama, G. Madariaga, L. Callejo, R. Cortes, 15 (2013) 9437–9443.
Cryst. Growth Des. 11 (2011) 1310–1317. [114] Z.-Q. Jiang, G.-Y. Jiang, D.-C. Hou, F. Wang, Z. Zhao, J. Zhang, CrystEngComm
[65] R. Fernandez de Luis, J.L. Mesa, M.K. Urtiaga, E.S. Larrea, T. Rojo, M.I. Arriortua, 15 (2013) 315–323.
Inorg. Chem. 51 (2012) 2130–2139. [115] P. Mahata, C. Mellot-Draznieks, P. Roy, S. Natarajan, Cryst. Growth Des. 13
[66] R. Haldar, T.K. Maji, CrystEngComm 14 (2012) 684–690. (2013) 155–168.

Please cite this article in press as: N. Li et al., Conformation versatility of ligands in coordination polymers: From structural diversity to properties and
applications, Coord. Chem. Rev. (2018), https://doi.org/10.1016/j.ccr.2018.05.016
N. Li et al. / Coordination Chemistry Reviews xxx (2018) xxx–xxx 29

[116] D.P. Martin, R.M. Supkowski, R.L. LaDuca, Inorg. Chem. 46 (2007) 7917–7922. [144] J.-F. Song, J.-J. Luo, Y.-Y. Jia, L.-D. Xin, Z.-Z. Lin, R.-S. Zhou, RSC Adv. 7 (2017)
[117] Y.-X. Tan, Y.-P. He, J. Zhang, Inorg. Chem. 51 (2012) 9649–9654. 36575–36584.
[118] D. Yang, Y. Tian, W. Xu, X. Cao, Q. Ju, W. Huang, Z. Fang, S. Zheng, Inorg. Chem. [145] L. Wang, W. You, W. Huang, C. Wang, X.-Z. You, Inorg. Chem. 48 (2009) 4295–
56 (2017) 2345–2353. 4305.
[119] Q. Yang, L. Huang, M. Zhang, Y. Li, H. Zheng, Q. Lu, Cryst. Growth Des. 13 [146] C.-Y. Chen, P.-Y. Cheng, H.-H. Wu, H.M. Lee, Inorg. Chem. 46 (2007) 5691–5699.
(2013) 440–445. [147] X. Zhao, H. He, F. Dai, D. Sun, Y. Ke, Inorg. Chem. 49 (2010) 8650–8652.
[120] M. Chen, H. Zhao, C.-S. Liu, X. Wang, H.-Z. Shi, M. Du, Chem. Commun. 51 [148] Z. Huang, H.-B. Song, M. Du, S.-T. Chen, X.-H. Bu, J. Ribas, Inorg. Chem. 43
(2015) 6014–6017. (2004) 931–944.
[121] Q.-R. Fang, D.-Q. Yuan, J. Sculley, J.-R. Li, Z.-B. Han, H.-C. Zhou, Inorg. Chem. 49 [149] M. Du, S.-T. Chen, X.-H. Bu, Cryst. Growth Des. 2 (2002) 625–629.
(2010) 11637–11642. [150] M. Du, X.-H. Bu, Bull. Chem. Soc. Jpn. 82 (2009) 539–554.
[122] M. Du, M. Chen, X. Wang, J. Wen, X.-G. Yang, S.-M. Fang, C.-S. Liu, Inorg. Chem. [151] S. Parshamoni, J. Telangae, S. Sanda, S. Konar, Chem. Asian J. 11 (2016) 540–
53 (2014) 7074–7076. 547.
[123] Y.-H. Han, Z.-Y. Zhou, C.-B. Tian, S.-W. Du, Green Chem. 18 (2016) 4086–4091. [152] Y. Peng, G. Li, J. Hua, Z. Shi, S. Feng, CrystEngComm 17 (2015) 3162–3170.
[124] M. Du, X. Wang, M. Chen, C.-P. Li, J.-Y. Tian, Z.-W. Wang, C.-S. Liu, Chem. Eur. J. [153] S. Surble, C. Serre, C. Mellot-Draznieks, F. Millange, G. Ferey, Chem. Commun.
21 (2015) 9713–9719. (2006) 284–286.
[125] H. Zhang, D. Chen, H. Ma, P. Cheng, Chem. Eur. J. 21 (2015) 15854–15859. [154] F. Millange, C. Serre, G. Férey, Chem. Commun. (2002) 822–823.
[126] M.-S. Chen, M. Chen, S. Takamizawa, T.-A. Okamura, J. Fan, W.-Y. Sun, Chem. [155] J. Seo, C. Bonneau, R. Matsuda, M. Takata, S. Kitagawa, J. Am. Chem. Soc. 133
Commun. 47 (2011) 3787–3789. (2011) 9005–9013.
[127] Y. Gong, Y.-C. Zhou, T.-F. Liu, J. Lue, D.M. Proserpio, R. Cao, Chem. Commun. 47 [156] C.-P. Li, J. Chen, C.-S. Liu, M. Du, Chem. Commun. 51 (2015) 2768–2781.
(2011) 5982–5984. [157] A. Demessence, J.R. Long, Chem. Eur. J 16 (2010) 5902–5908.
[128] X.-L. Wu, F. Luo, G.-M. Sun, A.-M. Zheng, J. Zhang, M.-B. Luo, W.-Y. Xu, Y. Zhu, [158] H.-L. Zhou, R.-B. Lin, C.T. He, Y.-B. Zhang, N. Feng, Q. Wang, F. Deng, J.-P.
X.-M. Zhang, S.-Y. Huang, ChemPhysChem 14 (2013) 3594–3599. Zhang, X.-M. Chen, Nat. Commun 4 (2013) 2534.
[129] B. Zheng, J. Bai, J. Duan, L. Wojtas, M.J. Zaworotko, J. Am. Chem. Soc. 133 [159] I.H. Park, R. Medishetty, J.Y. Kim, S.S. Lee, J.J. Vittal, Angew. Chem. Int. Ed. 53
(2011) 748–751. (2014) 5591–5595.
[130] R. Yun, Z. Lu, Y. Pan, X. You, J. Bai, Angew. Chem. Int. Ed. 52 (2013) 11282– [160] I.-H. Park, C.E. Mulijanto, H.-H. Lee, Y. Kang, E. Lee, A. Chanthapally, S.S. Lee, J.
11285. J. Vittal, Cryst. Growth Des. 16 (2016) 2504–2508.
[131] O. Benson, S.P. Argent, R. Cabot, M.J. Lennox, T. Mitra, T.L. Easun, W. Lewis, A.J. [161] W.-C. Song, X.-Z. Cui, Z.-Y. Liu, E.-C. Yang, X.-J. Zhao, Sci. Rep. 6 (2016) 34870.
Blake, E. Besley, S.I. da, S.F. Parker, P. Manuel, M. Savage, H.G.W. Godfrey, Y. [162] S. Yuan, H. Wang, D.-X. Wang, H.-F. Lu, S.-Y. Feng, D. Sun, CrystEngComm 15
Yan, S. Yang, M. Schroder, T. Mitra, T.L. Easun, J. Am. Chem. Soc. 138 (2016) (2013) 7792–7802.
14828–14831. [163] Q. Li, J. Qian, RSC Adv. 4 (2014) 32391–32397.
[132] Z. Lu, J. Zhang, H. He, L. Du, C. Hang, Inorg. Chem. Front. 4 (2017) 736–740. [164] L.J. Murray, M. Dinca, J.R. Long, Chem. Soc. Rev. 38 (2009) 1294–1314.
[133] Z.-H. Xuan, D.-S. Zhang, Z. Chang, T.-L. Hu, X.-H. Bu, Inorg. Chem. 53 (2014) [165] Z. Bao, G. Chang, H. Xing, R. Krishna, Q. Ren, B. Chen, Energy Environ. Sci. 9
8985–8990. (2016) 3612–3641.
[134] J. Duan, Z. Yang, J. Bai, B. Zheng, Y. Li, S. Li, Chem. Commun. 48 (2012) 3058– [166] R. Lyndon, K. Konstas, B.P. Ladewig, P.D. Southon, P.C. Kepert, M.R. Hill,
3060. Angew. Chem. Int. Ed. 52 (2013) 3695–3698.
[135] B. Zheng, Z. Yang, J. Bai, Y. Li, S. Li, Chem. Commun. 48 (2012) 7025–7027. [167] H. Li, M.M. Sadiq, K. Suzuki, C. Doblin, S. Lim, P. Falcaro, A.J. Hill, M.R. Hill, J.
[136] A. Schoedel, A.J. Cairns, Y. Belmabkhout, L. Wojtas, M. Mohamed, Z. Zhang, D. Mater. Chem. A 4 (2016) 18757–18762.
M. Proserpio, M. Eddaoudi, M.J. Zaworotko, Angew. Chem. Int. Ed. 52 (2013) [168] F. Luo, C.-B. Fan, M.-B. Luo, X.-L. Wu, Y. Zhu, S.-Z. Pu, W.-Y. Xu, G.-C. Guo,
2902–2905. Angew. Chem. Int. Ed. 53 (2014) 9298–9301.
[137] B. Zheng, H. Liu, Z. Wang, X. Yu, P. Yi, J. Bai, CrystEngComm 15 (2013) 3517– [169] C.-B. Fan, L. Le Gong, L. Huang, F. Luo, R. Krishna, X.-F. Yi, A.-M. Zheng, L.
3520. Zhang, S.-Z. Pu, X.-F. Feng, M.-B. Luo, G.-C. Guo, Angew. Chem. Int. Ed. 56
[138] D.K. Kumar, A. Das, P. Dastidar, Cryst. Growth Des. 7 (2007) 205–207. (2017) 7900–7906.
[139] I.H. Park, R. Medishetty, H.H. Lee, C.E. Mulijanto, H.S. Quah, S.S. Lee, J.J. Vittal, [170] J.-P. Zhang, X.-M. Chen, J. Am. Chem. Soc. 131 (2009) 5516–5521.
Angew Chem. Int. Ed. 54 (2015) 7313–7317. [171] M.L. Foo, R. Matsuda, Y. Hijikata, R. Krishna, H. Sato, S. Horike, A. Hori, J. Duan,
[140] A. Schneemann, V. Bon, I. Schwedler, I. Senkovska, S. Kaskel, R.A. Fischer, Y. Sato, Y. Kubota, M. Takata, S. Kitagawa, J. Am. Chem. Soc. 138 (2016) 3022–
Chem. Soc. Rev. 43 (2014) 6062–6096. 3030.
[141] Z. Chang, D.-H. Yang, J. Xu, T.-L. Hu, X.-H. Bu, Adv. Mater. 27 (2015) 5432– [172] M.D. Allendorf, C.A. Bauer, R.K. Bhakta, R.J. Houk, Chem. Soc. Rev. 38 (2009)
5441. 1330–1352.
[142] Q.-Q. Li, W.-Q. Zhang, C.-Y. Ren, Y.-P. Fan, J.-L. Li, P. Liu, Y.-Y. Wang, [173] D.E. Williams, J.A. Rietman, J.M. Maier, R. Tan, A.B. Greytak, M.D. Smith, J.A.
CrystEngComm 18 (2016) 3358–3371. Krause, N.B. Shustova, J. Am. Chem. Soc. 136 (2014) 11886–11891.
[143] J.-J. Hou, R. Zhang, Y.-L. Qin, X.-M. Zhang, Cryst. Growth Des. 13 (2013) 1618– [174] F.-F. Li, L. Zhang, L.-L. Gong, C.-S. Yan, H.-Y. Gao, F. Luo, Dalton Trans. 46
1625. (2017) 338–341.

Please cite this article in press as: N. Li et al., Conformation versatility of ligands in coordination polymers: From structural diversity to properties and
applications, Coord. Chem. Rev. (2018), https://doi.org/10.1016/j.ccr.2018.05.016

You might also like