Download as pdf or txt
Download as pdf or txt
You are on page 1of 21

Coordination Chemistry Reviews 414 (2020) 213300

Contents lists available at ScienceDirect

Coordination Chemistry Reviews


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

Review

Recent progress on supramolecular assembly of organoplatinum(II)


complexes into long-range ordered nanostructures
Yifei Han, Zongchun Gao, Cong Wang, Ruolei Zhong, Feng Wang ⇑
CAS Key Laboratory of Soft Matter Chemistry, Department of Polymer Science and Engineering, University of Science and Technology of China, Hefei, Anhui 230026, PR China

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

Article history: Organoplatinum(II) complexes have received enormous attention because of their intriguing photo-
Received 2 December 2019 physical and chemotherapeutic behaviors. They are also regarded as excellent building blocks for
Received in revised form 2 March 2020 supramolecular assembly due to the square planar geometry and coordination unsaturation characters,
Accepted 14 March 2020
which facilitate to incorporate p–p stacking, Pt(II)–Pt(II) metal–metal and other non-covalent interac-
Available online 2 April 2020
tions together. In this review, we have summarized the state-of-art progress of organoplatinum(II)-
based long-range ordered nanostructures, as catalogued by the diverse assembling modes. In addition,
Keywords:
potential applications of organoplatinum(II)-based supramolecular nanostructures in optoelectronics,
Supramolecular assembly
Organoplatinum(II) complexes
sensing, and biomedical fields are briefly introduced.
Metal–metal interactions Ó 2020 Elsevier B.V. All rights reserved.
Pathway complexity
Supramolecular materials

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
2. Construction of organoplatinum(II)-based supramolecular nanostructures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
2.1. Primary self-assembly of monodentate Pt(II) complexes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
2.2. Primary self-assembly of multidentate Pt(II) complexes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
2.3. Hierarchical assembly of Pt(II) complexes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
2.4. Assembly of Pt(II) complexes into diverse morphologies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
2.5. Pathway complexity in assembling Pt(II) complexes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.6. Multi-component co-assembly of Pt(II) complexes. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.7. Hybrid co-assembly with the presence of Pt(II) complexes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
3. Functionality of organoplatinum(II)-based supramolecular nanostructures. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
3.1. Sensing materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
3.2. Emitting materials. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
3.3. Semi-conductive materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
3.4. Biomedical materials. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.5. Other potential applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
4. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
Conflict of interest . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
Appendix A. Supplementary data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18

⇑ Corresponding author.
E-mail address: drfwang@ustc.edu.cn (F. Wang).

https://doi.org/10.1016/j.ccr.2020.213300
0010-8545/Ó 2020 Elsevier B.V. All rights reserved.
2 Y. Han et al. / Coordination Chemistry Reviews 414 (2020) 213300

1. Introduction region [13–15]. The emergent spectral transition at the biologically


interesting window would benefit for the biomedical applications
Platinum(II) complexes represent an important type of transi- of organoplatinum(II) complexes. Third, the modular incorporation
tion metal complexes with d8 electron configuration and square of various functional groups imparts multifunctionlity and adapt-
planar geometry. When organic ligands are bound to the metal ability to the resulting supramolecular nanostructures [16]. In this
center, a variety of organoplatinum(II) complexes generate with context, supramolecular assembly of organoplatinum(II) com-
mono- and multi-dentate character [1–3]. Spin–orbit coupling plexes has achieved considerable progress over the last two dec-
between the metal and ligand units renders fruitful photo- ades [17–22].
physics to the resulting organoplatinum(II) complexes, including Up to now, both monodentate and multidentate (including bi-,
ligand-centered (LC), metal-to-ligand charge transfer (MLCT), and tri- and tetra-) organoplatinum(II) complexes (Scheme 1) have
ligand-to-ligand charge transfer (LLCT) electronic transitions, etc. been employed to construct well-ordered supramolecular assem-
[4,5]. As a consequence, the absorption/emission wavelength, the blies. For the mononuclear Pt(II) complexes, the intermolecular
emission intensity/lifetime, and the Stokes shifts can be elabo- Pt(II)–Pt(II) binding affinity is rather low (<10 kcal mol1) [14],
rately modulated, which are promising for organic light-emitting which is not strong enough to drive the formation of long-range-
diodes (OLEDs), sensing, and bio-imaging applications [6,7]. ordered assemblies. Accordingly, the elaborate choice of low-
Besides, the organoplatinum(II) complexes have shown bright pro- steric ligands (such as halogen, acetylene, pyridine, etc.), together
spect in chemotherapeutic fields [8], attributing to the clinical suc- with the synergistic incorporation of hydrogen bonds, facilitate
cess of cisplatin [Pt(II)Cl2(NH3)2] for the anti-cancer applications. to enhance intermolecular complexation strength especially in
Bottom-up supramolecular assembly exists ubiquitously in nat- dilute solution. On the other hand, multidentate organoplatinum
ural biological systems such as DNA and proteins. In view of its low (II) complexes possess large p-planar surface. Thanks to the con-
cost and convenient fabrication characters, supramolecular assem- current participation of p–p stacking and Pt(II)–Pt(II) metal–metal
bly has become an effective technique toward long-range ordered interactions, multidentate organoplatinum(II) complexes are
nanostructures with the emergent functions. Up to now, various regarded as excellent building blocks for supramolecular assembly.
artificial supramolecular assemblies have been established on the It should be noted that bottom-up assembling processes are
basis of small molecules, polymers, and inorganic nanoparticles influenced by a number of intrinsic and extrinsic parameters. In
[9–11]. Organoplatinum(II) complexes also represent ideal this regard, unravelling the assembling mechanism and pathway
supramolecular building blocks due to the square planar geometry is necessary to dictate the ordering of supramolecular nanostruc-
and coordination unsaturation characters, which display strong tures. Depending on the energy evolution of supramolecular
tendency to stack with each other. The resulting long-range assembling processes, two different mechanisms can be divided,
ordered nanostructures are endowed with some fascinating prop- namely isodesmic and cooperative mechanism [23–25]. For the
erties. First, supramolecular assembly generates the rigidifying isodesmic mechanism, non-covalent forces involved in each of
effect for organoplatinum(II) complexes. The declince of non- the assembling step are identical to each other. In comparison,
radiation decay rate, together with the shielding of chormophore the cooperative mechanism initiates with the nucleation step, fol-
from the environment, could potentially give rise to the emission lowed by the nonlinear growth of the elongated chain. As com-
enhancement property [12]. Second, supramolecular assembly is pared to the isodesmic mechanism, the cooperative one avoids
often accompanied by Pt(II)–Pt(II) orbital overlapping of the neigh- the formation of small oligomers, and thereby favors for the forma-
bouring organoplatinum(II) complexes. Accordingly, 5d2z orbitals tion of long-range ordered nanostructures with high degree of
are re-organized into low-lying dr and high-lying dr* orbital for polymerization value. In addition, pathway complexity also plays
the Pt atoms, leading to the appearance of metal–metal-to-ligand a crucial role to determine organization and morphology of the
charge transfer (MMLCT) transition at low-energy visible/NIR final nanostructures, by tailoring supramolecular assembly to the

Scheme 1. Typical organoplatinum(II)-based chemical structures for supramolecular assembly.


Y. Han et al. / Coordination Chemistry Reviews 414 (2020) 213300 3

thermodynamic equilibrium state or kinetic non-equilibrium state (Fig. 2a). The critical gelation concentration is approximately
[26–28]. 1 mM in dodecane and hexane for both 1a and 1b. H-aggregation
The aim of this review is to highlight state-of-the-art progress of platinum(II) acetylides occurs during supramolecular assembly
of organoplatinum(II)-based supramolecular assembling systems, process, as reflected by the obvious blue shift (~40 nm) in the
especially those developed in the recent ten years. Several elegant absorption spectra. Moreover, the platinum(II) acetylide com-
reviews have already summarized spectroscopy and functionality plexes are phosphorescent both in the solution and gel states. Tri-
of organoplatinum(II)-based assemblies [14,15,17–22]. Neverthe- plet energy transfer occurs when small amount of 1b (4 mol%) is
less, few of them are focused specifically to long-range ordered doped into supramolecular gels of 1a. It involves exciton diffusion
supramolecular nanostructures. To stimulate further research among 1a chromophores, followed by Dexter exchange energy
activity in this area, herein the emphasis is paid to the molecular transfer to the trapped compound 1b. This work provides the first
design principles, which enable precise control over long-range example of triplet exciton diffusion and trapping in an
ordered supramolecular organization of organoplatinum(II) com- organoplatinum(II)-based supramolecular assembling system.
plexes. For convenience of the detailed discussion, six major sys- Platinum(II) acetylide-based supramolecular systems have been
tems are catalogued by the supramolecular assembling modes also extensively studied by Yang, Xu, and coworkers [30–35]. A
(Fig. 1). In addition, the applications of organoplatinum(II)-based series of linear-, triangular-, and rectangular-shaped monomers
supramolecular nanostructures in sensing, emitting, semi- are designed and synthesized. Intriguingly, various functional
conducting, bio-medical fields are briefly introduced. groups such as iptycene, arene, porphyrin, and azobenzene have
been incorporated to the rigid platinum(II) acetylide units. Most
of the synthetic compounds exhibit high gelation efficiency in non-
2. Construction of organoplatinum(II)-based supramolecular polar hydrocarbon solvents. Moreover, honey comb-patterned
nanostructures microporous films [31] and nanospheres [33] are fabricated for
some of the synthetic compounds under the specific conditions.
2.1. Primary self-assembly of monodentate Pt(II) complexes Despite the progress achieved, in-depth understanding of the
assembling mechanism involved in platinum acetylide-based sys-
Platinum(II) acetylides represent a typical type of monodentate tems have remained unexplored. Considering that natural assem-
organoplatinum(II) complexes. In 2008, Schanze and coworkers bling systems such as microtubules and tobacco mosaic virus
have reported a series of coil-rod-coil monomers (see compounds tend to adopt cooperative nucleation–elongation mechanism, our
1a–b in Fig. 2a) [29], consisting of platinum(II) acetylide as the research group sought to mimic them and construct
rigid rod and two tri(dodecyloxy)phenyl units on the periphery. organoplatinum(II)-based supramolecular assemblies via the same
With the presence of rod–rod p–p stacking and rod–coil phase seg- mechanism. To attain this objective, the key strategy involves the
regation, these monomers aggregate in hydrocarbon solvents to integration of two amide groups into rod-like platinum(II) acety-
form supramolecular nanofibers and gels at room temperature lide unit (see monomers 2a–b in Fig. 2b). During the self-
assembly process, hydrogen bonding and p–p stacking interactions
are participated in a synergistic manner [36–40]. (S)-3,7-
dimethyloctyl groups are introduced into the peripheral side chain,
facilitating to induce a helical bias for the self-assembled struc-
tures (Fig. 2b). Accordingly, the supramolecular assembling pro-
cess can be quantitatively probed in apolar methylcyclohexane,
via the combination of circular dichroism (CD) spectroscopic mea-
surement and the mathematical analysis. Non-sigmoidal melting
curve is obtained (Fig. 2b), proving the involvement of cooperative
nucleation–elongation mechanism to form long-range ordered
supramolecular polymers.
Moreover, subtle change on the molecular structure exerts sev-
ere impact on the supramolecular polymerization behavior. For
example, the less bulky PMe3 and PEt3 substitution on platinum
(II) acetylide complexes are capable of forming single-handed
supramolecular fibers, while no aggregation is detected for the
counterpart complex with the bulky PBu3 substitution [37]. Poten-
tial applications of the resulting supramolecular nanostructures
are also exploited. For instance, the [4+2] endoperoxidation is trig-
gered by irradiating 2a under 460 nm LED, whilst the backward
deoxygenation reaction proceeds upon standing at room tempera-
ture [39]. Fluorescent anti-counterfeiting QR codes are further fab-
ricated for supramolecular polymers of 2a via the inkjet-printing
technique. Thanks to the reversible photo-responsiveness, the
embedded information in QR codes can be reversibly erased and
restored with fast response and ease of operation (Fig. 2b).
As compared to the above platinum(II) phosphine complexes,
platinum(II) pyridine complexes possess low steric hindrance.
Accordingly, they are also employed as the primary building blocks
to construct well-ordered nanostructures [41–43]. In this respect,
Fernández and coworkers have designed and synthesized com-
pound 3a (Fig. 3). It is capable of assembling into nanofibers in
Fig. 1. Organoplatinum(II)-based long-range ordered nanostructures catalogued by methylcyclohexane via the cooperative nucleation–elongation
supramolecular assembling modes. mechanism [44]. Remarkably, 3a adopts slipped arrangement for
4 Y. Han et al. / Coordination Chemistry Reviews 414 (2020) 213300

Fig. 2. (a) Structures of 1a–b, and phosphorescent gel of 1a in dodecane (c = 1 mM). (b) Graphic representation for the cooperative supramolecular polymerization of 2a–b.
Bottom-left, variable-temperature CD spectra of 2a in methylcyclohexane, and the corresponding non-sigmoidal melting curve (inset); bottom-right, photographs of the
fluorescent QR code (size: 4  4 cm on paper) from 2a and its photo-responsiveness. Reproduced with permission of 2008 American Chemistry Society from ref. [29], and
2018 Nature Publishing Group from ref. [39].

the neighboring monomers in the self-assembled state (Fig. 3). It CAH---p forces, along with CAH---O interactions between the
involves p-stacking of four out of six aromatic rings, together with peripheral triethylene glycol chains of two different molecules,
several weak intermolecular interactions such as CAH---p, stabilize the aggregated structure.
C„C---Pt, and CAH---Cl forces (Fig. 3). In terms of the amphiphilic
monomer 3b (Fig. 3), it is prone to self-assemble into deep yellow 2.2. Primary self-assembly of multidentate Pt(II) complexes
gels in alcohols (MeOH, EtOH, n-BuOH) and water [45]. In the
self-assembled structures, 3b also packs in a slipped fashion. As Multidentate Pt(II) complexes, especially platinum(II) polypyri-
evidenced by X-ray diffraction of 3b, multiple CAH---Cl and dine complexes, are capable of generating Pt(II)–Pt(II) metal–metal
interactions upon supramolecular aggregation. The process is com-
monly accompanied by the dramatic spectroscopic changes. Hence,
these complexes have been widely exploited for the fabrication of
long-range ordered supramolecular nanostructures. Yam’s group
has pioneered the self-assembly study of platinum(II) terpyridine
[Pt(terpy)] complexes. In their early study, they have reported
the polymorphism phenomena (dark green color versus red color
of the single crystals) of [Pt(terpy)(C„CAC„CH)]OTf (OTf = tri
fluoromethanesulfonate) crystals [46]. Short intermolecular Pt-Pt
distance of 3.388 Å exists for the dark green crystal form, denoting
the presence of strong Pt(II)–Pt(II) metal–metal interactions.
On this basis, Yam and coworkers sought to construct
supramolecular metallogels, by attaching tris-dodecyl hydrocar-
bon chains on the periphery of alkynylplatinum(II) terpyridine unit
(Fig. 4a) [47–49]. For monomer 4a (Fig. 4a), it is prone to form clear
gel in organic solvents such as dodecance and DMSO (critical gela-
tion concentration: 4.4 mg/mL in DMSO; Tgel = 55 °C). Drastic color
change from purple to orange occurs for 4a during gel-to-sol tran-
sition process. It originates from the presence/absence of Pt(II)–Pt
(II) metal–metal interactions during supramolecular aggregation/
de-aggregation processes. Interestingly, the gelation process can
Fig. 3. Structure of 3a–b, and energy-minimized geometry for the tetramer of 3a
be modulated by structural variation at the molecular level. For
obtained via semi-empirical PM6 calculation. Reproduced with permission of 2016 example, 4b (Fig. 4a) with three tert-butyl substituents on the ter-
Wiley-VCH Verlag GmbH from ref. [44]. pyridine ligand forms less stable metallogels, as reflected by its
Y. Han et al. / Coordination Chemistry Reviews 414 (2020) 213300 5

Fig. 4. (a) Gelation behaviors of 4a–c in DMSO. (b) Structures of 5a–b, and the solvato-chromic behavior of 5a. (c) Unusual emission enhancement phenomenon during the
gel-to-sol phase transition of 6. Reproduced with permission of 2007 Royal Society of Chemistry from ref. [48], 2007 Wiley-VCH Verlag GmbH from ref. [51], and 2009
American Chemistry Society from ref. [53].

higher critical gelation concentration (7.8 mg/mL) and lower sol– supramolecular gels on the basis of monomer 6 (Fig. 4c) [53].
gel transition temperature (Tgel = 45 °C) than those of 4a. The result Intriguingly, 6 shows emission enhancement during the gel-to-
is ascribed to the absence of Pt(II)–Pt(II) metal–metal interactions sol phase transition upon increasing the temperature. This unusual
for the 4b self-assembly process, since the steric bulky tert-butyl behavior is rarely encountered, which is in stark contrast to the
substituents greatly hinder Pt centers from closely proximity. In conventional thermotropic luminescent gels. The unique phe-
addition, 4a and 4c (Fig. 4a) display distinct color for the metallo- nomenon is ascribed to heating-induced increase in the degrees
gel [purple for the OTf salt versus deep red for the PF6 (hexafluo- of freedom and motion, which consequently increase the excimer
rophosphate) salt, Fig. 4a]. Hence, counter-anion variation population. Luminescence quantum yield of the excimeric excited
governs the extent of Pt(II)–Pt(II) and p–p interactions involved state is likely to be much higher than that of the IL excited state,
in the gelation process. In addition, Yi and coworkers have leading to a strongly luminescent sol form than the gel form of 6.
designed cholesterol-appended Pt(terpy) gelators, which form Supramolecular assembly has also been investigated for other
stable metallogels in alcohol, ethyl acetate and toluene upon son- tri- and bi-dentate platinum(II) complexes [54–57]. For example,
ication [50]. our research group has replaced alkynyl Pt(terpy) by the isoelec-
Hydrogen bonds provide the additional stabilization factor for tronic platinum(II) 6-phenyl-2,20 -bipyridine isocyanide complex
Pt(terpy)-based supramolecular assemblies. Ziessel and coworkers [58]. An obvious solvochromic phenomenon occurs during the
have attached Pt(terpy) to a 3,5-diacylamido framework (see reversible assembling process, with yellow-green emission in the
monomers 5a-b in Fig. 4b) [51]. The amide units in 5a-b are monomeric state and red-emission in the aggregated state. Assem-
expected to strengthen intermolecular association via intermolec- bling organoplatinum(II) complexes can be also proceeded in aque-
ular hydrogen bonds. The resulting systems display solvato- ous media [59–61]. For example, De Cola et al. have designed
chromic properties, by reversibly converting among the mono- neutral platinum(II) 1,3-di(2-pyridine)-benzene complexes, which
meric, oligomeric and polymeric states (Fig. 4b). TEM and AFM bear hydrophilic oligo-ethylene glycol chains to increase the water
measurements confirm the presence of linear fibrils with lengths solubility. The designed amphiphilic molecules tend to aggregate
of several hundred nanometers in the dodecane gels (critical gela- to phosphorescent hydrogels, by utilizing host–guest interactions
tion concentration = 7.5 mmol/L). Pt(II)–Pt(II) metal–metal interac- between cyclodextrins and the oligo-ethylene glycol tail of
tions are involved in the gelation state, as reflected by the deep- organoplatinum(II) complexes [59]. Moreover, the resulting
green color and near-infrared emission at 830 nm. For 5b with supramolecular nanostructures exhibit a high cell permeability
the branched alkyl chains on the periphery, it tends to form hexag- and low cytotoxicity, which is promising for long-lived bio-
onal 2D lattice mesophases. Besides, Yam’s group has also reported imaging application [60].
Pt(terpy) complex with the L-valine-modified alkynyl ligand [52]. In addition to the above mononuclear cyclometallated platinum
The compound is capable of forming gel in acetonitrile (critical (II) systems, supramolecular assembly can also take place for the
gelation concentration = 2.5 mg/mL), attributed to the cooperative multinuclear platinum(II) complexes [62–65]. Yam’s group has
participation of Pt(II)–Pt(II), p–p, and hydrogen bonding reported an elegant example, in which dinuclear Pt(terpy) are teth-
interactions. ered by oligo(para-phenylene ethynylene)s unit [63]. Two hydro-
Generally, Pt(terpy) complexes suffer from low emission inten- philic triethylene glycol and two hydrophobic octadecane chains
sity in fluid solution, because of the involvement of non-radiative are attached to the rigid segment, endowing amphiphilic character
d–d deactivation pathway [1]. To address this issue, a feasible to the resulting monomers 7a–b (Fig. 5a). Notably, 7a tends to
strategy is to replace terpyridine by the 2,6-bis(N-alkylbenzimida aggregate into tubular architecture in DMSO, with a uniform diam-
zol-20 -yl)pyridine (bzimpy) analogue. Triplet intra-ligand (IL) p– eter of ~13.0 nm and length of 2.0–5.5 lm (Fig. 5a). In stark con-
p* emission of the bzimpy ligand is lower-lying in energy than that trast, 7b, bearing tert-butyl units on one of the terpyridine unit,
of terpyridine, thus giving rise to the highly emissive character. On is prone to form helical ribbons with the diameter of ~9.5 nm
this account, Yam’s group has designed a novel type of emissive and 0.7–2.1 lm in length (Fig. 5a). Presumably, presence of steric
6 Y. Han et al. / Coordination Chemistry Reviews 414 (2020) 213300

Fig. 5. (a) Supramolecular assembly of 7a and 7b into nanotubes and helical nanoribbons in DMSO, respectively. (b) Distinct supramolecular assembly mechanisms for
monomers 8a–c. Reproduced with permission of 2016 National Academy of Sciences from ref. [63], and 2018 American Chemistry Society from ref. [64].

monomer 8a with the shortest hydrophobic chains, it undergoes


isodesmic mechanism for the self-assembly process (Fig. 5b). Upon
increasing the alkyl lengths from AC6H13 to AC12H25 (monomer
8b), self-assembly follows the cooperative nucleation–elongation
mechanism (Fig. 5b). Nevertheless, it switches back to the isodes-
mic one, when the hydrophobic chains are further prolonged
(monomer 8c, Fig. 5b). Hence, variation of the peripheral alkyl
chains exerts a profound influence on the supramolecular assem-
bling mechanisms.
Additionally, Naota and co-workers have reported a chiral,
clothespin-shaped dinuclear trans-bis(salicylaldiminato) Pt(II)
complex 9 (Fig. 6) [66,67]. The compound is non-emissive in
organic solvents such as cyclohexane. Nevertheless, upon ultra-
sound sonication for a few seconds, hetero-chiral, interpenetrative
dimeric structure generates, further leading to the formation of
supramolecular polymers and phosphorescent gels (Fig. 6). The
gels convert back to the solution state after heating above Tgel.
Emission signal of the supramolecular gels can be further con-
Fig. 6. Ultrasound-induced sol-to-gel phase transition of clothespin-shaped molec-
ular 9. Reproduced with permission of 2011 American Chemical Society from ref. trolled by sonication time, linker length, and optical activity of
[66] and 2019 Wiley-VCH Verlag GmbH from ref. [67]. the complexes. This intriguing example represents a remote and
instant control over the supramolecular gelation processes.
bulky tert-butyl groups on 7b restricts the formation of directional
Pt(II)–Pt(II) metal–metal interactions, and thereby influences the 2.3. Hierarchical assembly of Pt(II) complexes
final morphologies of the self-assembled nanostructures.
In an extensive study, the same group has designed dinuclear Hierarchical assembly is regarded as an important route toward
platinum(II) bzimpy monomers 8a–c (Fig. 5b) [64]. The overall complex supramolecular systems [68]. Generally, the hierarchical
hydrophilicity/hydrophobicity are adjusted, by introducing the process relies on the construction of a pre-assembled
peripherical hydrocarbon chains with the different length. For structural motif, which further assembles through the additional
Y. Han et al. / Coordination Chemistry Reviews 414 (2020) 213300 7

non-covalent forces. Foldamers mimic the secondary structures of the pincer ligands, while the pentameric m-OPE acts as the foldable
proteins and nucleic acids, which are appealing helical scaffolds for linker [71]. Thanks to the presence of four dodecyl chains, it facil-
the hierarchical supramolecular assembly. In this respect, Yam and itates rod-coil phase segregation between the inner single-turn fol-
coworkers have previously prepared a series of dinuclear Pt(terpy) damer and the peripheral flexible chains. Accordingly, high-order
complexes connected by the oligomeric m-phenylene ethynylene hierarchical architectures are prone to form in acetonitrile via
(m-OPE) linker [69]. The synthetic compounds can fold back onto intra- and inter-molecular Pt(II)–Pt(II) metal–metal and p–p stack-
themselves in an intramolecular manner, giving rise to the forma- ing interactions. The hierarchical assembling process of 10 adopts
tion of short helical strands. The folding process is mediated by the cooperative nucleation-elongation mechanism. Depending on
changing the solvent from CHCl3 to CH3CN. When the acid- Meijer–Schenning–van-der-Schoot mathematical model [24], the
sensitive oligo(ethynyl-pyridine) is incorporated as the spacer enthalpy release upon elongation (he) and critical elongation tem-
instead of m-OPE, the reversible switching from transoid to cisoid perature (Te) are calculated to be 117.1 kJ mol1 and 318 K,
forms can be manipulated with multi-stimuli responsiveness respectively. The number-averaged degree of polymerization value
[70]. On this basis, it is fascinating to develop higher-order hierar- for 10 is calculated to be 1500 at room temperature. Ultimately,
chical nanostructures, by employing helical foldamers as the pre- orange-colored metallogel forms, because of the reciprocal associ-
assembled intermediate. ation of multiple helices. It should be mentioned that the m-OPE
To achieve this goal, Yam’s group has further designed com- spacer length is critical for the hierarchical assembly processes.
pound 10 (Fig. 7a). The tridentate bzimpy units on 10 serve as For the structurally similar compound with the tetrameric m-OPE

Fig. 7. (a) Graphic representation for the hierarchical assembly of 10. (b) Graphic representation for the hierarchical assembly of (R)-11 and (S)-11. Right bottom, CD spectral
changes of (R)-11 upon increasing the water content in CH3CN. (c) Dynamic switching of 12 with distinct emission color changes. Reproduced with permission of 2017
American Chemistry Society from ref. [71], 2013 Royal Society of Chemistry from ref. [72], 2019 National Academy of Sciences from ref. [73].
8 Y. Han et al. / Coordination Chemistry Reviews 414 (2020) 213300

linker, intramolecular folding of the two terminal platinum(II) Yang and co-workers have also performed systematic studies
bzimpy moieties is restricted. As a consequence, it fails to gelate on hierarchical self-assembly of organoplatinum(II) moieties [74–
in CH3CN under the same circumstances. 80]. In their design, metal-coordination macrocycles and cages
Later on, Yam and coworkers have designed two enantiomers are initially formed as the pre-assembled motif in CH2Cl2. When
(S)-11 and (R)-11 (Fig. 7b) [72]. The presence of two hydrophilic tri- platinum(II) bzimpy units are attached on the periphery, hierarchi-
ethylene glycol monomethyl ethers facilitates the hierarchical cal self-assembly takes place upon gradual addition of apolar
assembly in aqueous medium. Thanks to the presence of chiral cyclohexane, as driven by metallophilic Pt–Pt, p-p stacking and
binaphthol unit, the folding and hierarchical assembling processes hydrophobic interactions. The resulting metallogels are collapsed
can be directly monitored via CD spectroscopy. Upon increasing the upon addition of electron-rich coronene, because of the presence
water/acetonitrile volume ratio, (R)-11 exhibits an extraordinary of competitive donor–acceptor interactions between coronene
growth of the Cotton effect, reaching a maximum at 70% water in and alkynylplatinum(II) bzimpy units [74]. In another related
the high-energy region (271–340 nm, Fig. 7b). The progressive study, the same group has found that the platinum(II) bzimpy-
increase of CD intensity for (R)-11 reflects the formation of helically based assemblies feature intriguing chromic behaviors upon selec-
stacked array of M-helices. There is a slight increase in the CD signal tive absorption of CH2Cl2 vapor [75].
in the region of the MMLCT transition tail at 430 nm (Fig. 7b), sug-
gesting the participation of Pt(terpy) moieties for the hierarchical
assembly process. In terms of (S)-11, its CD spectrum exhibits a mir- 2.4. Assembly of Pt(II) complexes into diverse morphologies
ror image relationship to that of (R)-11 under same conditions.
Fibril-like assemblies are observed for both (S)-11 and (R)-11, with Supramolecular assembly process is primarily governed by the
the length of 2.0–2.5 lm and width of 130–480 nm. Therefore, hier- combination of various non-covalent driving forces. With the del-
archical assembly through the reciprocal association of multiple icate balance of these non-covalent bonds, it is feasible to obtain
helices leads to the formation of supramolecular columns. diverse morphologies under different conditions. For example,
Very recently, the same group has constructed another type of Yam and co-workers have designed a negatively-charged Pt(bz-
platinum(II)-based hierarchical assemblies, which display drastic impy) complex 13 (Fig. 8a), consisting of two pendant sulfonate
phosphorescent emission color changes. The design principle relies units [81]. Notably, the solvation conditions have a profound
on the molecular hinge 12 (Fig. 7c) [73]. The use of rigid butadiynyl impact on the final morphologies of supramolecular nanostruc-
linker as the rotating axis, together with the incorporation of tures. In 1:1 water/acetone (v/v), the monomeric species domi-
cyclometalated organoplatinum(II) moieties as the flaps, leads to nates, as manifested by the absence of MMLCT bands in both
the stabilization of 12 in the closed form in CH2Cl2. This state is UV–Vis and emission spectra (Fig. 8a). Increase of either acetone
mainly stabilized by intra-molecular p–p stacking interaction or water content leads to the appearance of MMLCT absorption
between the organoplatinum(II) pincers, leading to the deep-red bands (Fig. 8a), suggesting the existence of supramolecular aggre-
emission in CH2Cl2. Upon increasing the amount of apolar hexane, gation. Nevertheless, the assembling scenario is a little different in
the red emission band gradually declines (kmax = 690 nm in both media. Specifically, upon increasing the water content, 13
CH2Cl2), accompanied by the emergence of green emission tends to assemble into bilayered vesicular structures (Fig. 8a). It
(kmax = 516 nm in hexane/CH2Cl2 = 9/1). The green emission orig- is likely that charged sulfonate groups are pointing outward in
inates from triplet intra-ligand excited state. The signal illustrates water, whilst the hydrophobic Pt(bzimpy) moieties pack together
the predominance of the open form of 12, which hierarchically to avoid unfavorable contact with water. On the other hand, in
aggregates via intermolecular stacking. In addition to solvent vari- 9:1 acetone:water (v/v), the sulfonate groups are no longer well
ation, the reversible close-to-open conformational changes can be solvated. As a consequence, the sulfonate ionic heads start to
also regulated upon heating/cooling. aggregate and pull the complexes into close proximity, leading to

Fig. 8. (a) UV–Vis absorption and color changes of 13 upon varying water/acetone volume ratio. TEM images demonstrate the morphological change from vesicle to
nanofibers upon increasing the acetone content. (b) Solution color changes of 14 upon increasing the content of water in THF. TEM images show the change from nanoring to
nanofibers upon increasing the water content. Reproduced with permission of 2011 American Chemistry Society from ref. [81], and 2014 American Chemistry Society from
ref. [82].
Y. Han et al. / Coordination Chemistry Reviews 414 (2020) 213300 9

their aggregation into nanofibers (Fig. 8a). More interestingly, mor- tems [83–89]. In their design principles, organoplatinum(II) com-
phological transformation is accompanied by the variation of opti- plexes are covalently attached to flexible polymers such as
cal signals. The red color in aqueous solution, together with the polystyrene. It is found that solvent polarity exerts crucial impact
blue color in 9:1 acetone:water (v/v), are distinct from that in on the morphologies of the resulting assemblies. Briefly, free-
the monomeric state (yellow color in 1:1 acetone:water, v/v), as standing bilayered sheets are formed in toluene/methanol solvent
directly visualized by the naked eye (Fig. 8a). (v/v = 33/67), as evidenced by TEM and HAADF-STEM experiments.
Later on, the same group has designed an amphiphilic Pt(terpy) When the methanol content further increases to 90 vol%, the bilay-
complex 14 (Fig. 8b) [82], which also assembles into diverse nanos- ered nanosheets convert into core-corona micelles, with the diam-
tructures under different solvation conditions. The key design prin- eter of 25–80 nm [87]. Such morphological transformation is
ciple lies in the presence of hydrophobic polyhedral oligomeric primarily ascribed to the solvophobic effects of hydrophobic poly-
silsesquioxanes (POSS) unit on 14. In pure THF, 14 is in the molec- styrene in methanol.
ularly dissolved state. With the gradual addition of water into THF
(water/THF: 30/70, v/v), 14 aggregates into nano-ring structure 2.5. Pathway complexity in assembling Pt(II) complexes
(Fig. 8b). It is primarily governed by the hydrophobic–hydrophobic
interactions between the neighboring POSS scaffolds. When the Supramolecular nanostructures described in the above sections
water content further increases (water/THF: 70/30, v/v), the mor- are primarily in the thermodynamic equilibrium state. Besides,
phology converts from nano-rings to nano-rods (Fig. 8b). It is pathway complexity can be also involved in the organoplatinum
accompanied by the emergence of low-energy MMLCT spectro- (II)-based supramolecular assembly process. As a consequence,
scopic bands. Unambiguously, the organoplatinum(II) complexes both equilibrium thermodynamic and non-equilibrium kinetic
pack together in a head-to-head manner, with the participation products form under same conditions. For example, Rybtchinski
of directional metallophilic interactions. Therefore, the subtle and co-workers have reported pathway-dependent supramolecu-
interplay of Pt(II)–Pt(II) metal–metal, p–p stacking, and lar assembly via the solvent processing technique [90]. The
hydrophobic-hydrophobic interactions are responsible for the designed amphiphile 14 contains perylene diimide/Pt(terpy) core,
morphological transformation. along with PEG and tripeptide on each of the end (Fig. 9a). At high
Besides, Bu and coworkers have also reported morphological water content (water/THF = 95:5, v/v), 14 assembles into disor-
transformation of organoplatinum(II)-based supramolecular sys- dered curved fibers due to the existence of strong hydrophobic

Fig. 9. (a) Graphic illustration of different morphologies obtained by diluting a water/THF = 80/20 mixture of 14 to 95/5 at different evolution times during self-assembly (0,
45, and 2400 min). (b) Illustration of the landscape for supramolecular assembly of 15, along with the snapshots of time-dependent morphologies. (c) Graphic representation
for precise control over the length of supramolecular nanofibers assembled from 16 via the seeded growth approach. Reproduced with permission of 2011 Wiley-VCH Verlag
GmbH from ref. [90], 2015 Nature Publishing Group from ref. [91], and 2015 Royal Society of Chemistry from ref. [93].
10 Y. Han et al. / Coordination Chemistry Reviews 414 (2020) 213300

interactions. Simultaneously, a weak CD signal appears, which technique, the three different types of supramolecular assemblies
does not evolve over time. It is rationalized that supramolecular can be visualized in real time.
aggregates are in the kinetic-trapped state, which fails to convert On this basis, Manners and coworkers have prepared another
into the thermodynamic state because of the slow assembly/disas- amphiphile 16 (Fig. 9c) [93,94]. The pyridine ancillary ligand with
sembly dynamics. Upon increasing the THF percentage (water/ a long poly(ethylene glycol) (PEG16) chain could not only enhance
THF = 80/20, v/v), CD signal of the as-prepared solution increases the solubility, but prevent inter-chain packing at the fiber core.
with time, leading to the formation of straight fibers after three Accordingly, polydisperse rod-like structures are prepared, in
days. Obviously, the good solvent, THF, attenuates the solvophobic- which the diameter is closer to the diameter for a single
ity, providing the metastable assemblies which can equilibrate to supramolecular polymer chain. They have further explored
the thermodynamic product. On this basis, a water/THF = 80/20 ‘‘seeded growth approach” to achieve accurate control over the
(v/v) solution of 14 is prepared and allowed to evolve for 0 min, length and morphology of the assembling structures. Briefly, small
45 min or 2400 min, followed by dilution with water to 5% THF seed fibers are prepared in chloroform under the sonication condi-
content. Despite the equal solvent composition under the final con- tion. Further addition of 16 into the seeded fibers (number-average
ditions (95:5 water/THF, v/v, c = 1  104 M), distinct assemblies contour length: 60 nm) in acetonitrile leads to the formation of
such as short curved fibers, long fibers and nanotubes are obtained uniform nanofibers with relatively narrow length distributions
(Fig. 9a), which do not interconvert for at least two months. (number-average contour length: 390 nm; length distribution:
Besides, De Cola and co-workers have reported an elegant 1.21). The fibers remain colloidally stable in solution for at least
example, which is capable of evolving from kinetic metastable to 6 months, while no significant changes in the length and length
thermodynamic states over time [91,92]. For the designed triden- distribution are observed over 3 days. Hence, seeded growth
tate platinum(II)-based amphiphile 15 (Fig. 9b), it initially aggre- method suppresses the dynamic exchanging of unimers. Such
gates into a kinetically metastable state (state A, Fig. 9b) in kinetic controlled processes facilitate to provide well-ordered
water/1,4-dioxane. For the resulting nanoparticles, hydrophilic supramolecular nanostructures.
flexible tails are solvated in the highly polar medium, while the Pathway complexity is also encountered for monodentate Pt(II)
hydrophobic platinum(II) complexes pack tightly in the inner core. complexes during the supramolecular assembly processes. In this
The orange phosphorescent emission of state A suggests the forma- respect, Fernández and coworkers have prepared an amphiphilic
tion of tight Pt(II)–Pt(II) metal–metal interactions. Intriguingly, monomer 17 (Fig. 10), which is unsymmetrically substituted with
after standing at room temperature for three weeks, state A con- triethyleneglycol and dodecyloxy side chains [95]. 17 undergoes
verts to the thermodynamically stable state (state C, Fig. 9b) with self-aggregation upon cooling its methylcyclohexane solution from
the blue emitting character. The thermodynamic state exhibits 353 K to 303 K. Weak coupling of the p-scaffold suggests the for-
micrometer long ribbon-like structures. The conversion process mation of slipped arrangement (state A, Fig. 10), which is similar
from the kinetic to thermodynamic states is rather slow, which to those of the aforementioned compounds 3a–b (Fig. 3). At this
can be sped up by increasing the volume of 1,4-dioxane in water. stage, discoidal nanoparticles are observed with the size between
Detailed kinetic studies reveal the presence of a transient state 20 and 60 nm, suggesting the involvement of anti-cooperative
(state B, Fig. 9b) during the A-to-C conversion process. It emits mechanism. Intriguingly, further cooling the methylcyclohexane
green phosphorescence, suggesting the presence of Pt(II)–Pt(II) solution of 17 to 283 K leads to a pronounced spectral change
metal–metal interactions with slightly longer distance than that (8 nm red shift of kmax, Fig. 10). It indicates the formation of
of state A. With the confocal fluorescence microscopic imaging pseudo-parallel arrangement (state B, Fig. 10), with less pro-

Fig. 10. Graphic representation for pathway complexity of compounds 17–18 during the self-assembly processes. Bottom right, variable temperature UV–Vis spectral changes
of 17 upon cooling. Reproduced with permission of 2019 Wiley-VCH Verlag GmbH from ref. [95].
Y. Han et al. / Coordination Chemistry Reviews 414 (2020) 213300 11

nounced translational displacement for the neighboring monomers Co-assembly can also take place between two different types of
than that of state A. For state B, fiber networks of several micro- organoplatinum(II) complexes. In this context, Magnus’ green salt,
meters in length are observed, demonstrating the cooperative [Pt(NH3)4][PtCl4], is regarded as a prototype [98]. In the crystal of
supramolecular polymerization of 17 into long 1D nanostructures. Magnus’ green salt, the oppositely charged platinum(II) complexes
Hence, the buffering of monomers into anti-cooperative assemblies form alternating chains via electrostatic and Pt(II)–Pt(II) metal–
may open up a new strategy for the pathway control of metal interactions. Inspired by this fascinating structure, much
organoplatinum(II)-based nanostructures. In a related study, the efforts have been devoted toward multi-component
same group has found that the two competing states [slipped (A) organoplatinum(II)-based co-assemblies. For example, Che’s group
vs pseudo-parallel (B)] do not interconvert over time in a period has designed two distinct organoplatinum(II) monomers, namely
of at least six months for monomer 18 (Fig. 10) [96]. With the pre- the neutral species 21 and the cationic species 22 (Fig. 12a)
cise controlling of cooling rate and concentration, both states can [99,100]. Nanowires prefer to form upon mixing equimolar amount
be isolated concomitantly in the same solution. The phenomenon of 21 and 22 together in the concentrated acetone, accompanied by
is in close resemblance to polymorphism encountered in the the emergence of a new absorbance in the range of 604–615 nm.
crystals. The newly-formed band is assigned to MMLCT transitions, demon-
strating the involvement of intermolecular Pt(II)–Pt(II) metal–
metal interactions. Upon further replacing chloride on 21 by the
2.6. Multi-component co-assembly of Pt(II) complexes acetonitrile ligand, morphology of the co-assembled structure
undergoes a nanowire-to-nanowheel metamorphism process
As compared to mono-component self-assembly, multi- (Fig. 12a). Higher reaction temperatures (up to 70 °C) and excess
component co-assembly is considered as an efficient protocol amount of NH4PF6 facilitate to the formation of nano-wheel struc-
toward complexed supramolecular systems. In this context, Che tures. Hence, the fascinating co-assembled supramolecular sys-
and co-workers have successfully constructed supramolecular tems feature ligand-triggered morphological evolution character.
cross-linked networks and hydrogels, by co-assembling mono- Our research group has also developed supramolecular alter-
and di-nuclear platinum(II) compounds together (see 19 and 20 nate co-assemblies via the synergistic participation of Pt(II)–Pt(II)
Fig. 11) [97]. Remarkably, the length of flexible oligo(oxyethylene) metal–metal interactions [101,102]. Specifically, positively-
chain on 20 exerts a crucial impact on the gelation capability in charged Pt(terpy) complex 23 is designed as the acceptor monomer
aqueous media. In detail, the long bridging chain (n = 76 for 20a) (Fig. 12b). It self-assembles in apolar methylcyclohexane/1,2-dich
is capable of maintaining unfolded conformation, which facilitates loroethane (95:5, v/v) through intermolecular p–p stacking and
intermolecular connection of supramolecular nanofibers derived van der Waals interactions. As a result, one-dimensional nanofi-
from 19. Hence, adding 2.0 wt% of 20a into the aqueous solution bers form via nucleation–elongation cooperative mechanism. It is
of 19 results in the spontaneous formation of hydrogels with bright worthy to note that, due to the presence of sterically bulky tert-
red emission. On the contrary, the ditopic cross-linker 20b with butyl units, Pt(II)  Pt(II) metal–metal interactions are ruled out
short oligo(oxyethylene) chain prefers to adopt folded conforma- for the 23 mono-component self-assembly process (Fig. 12b).
tion. It fails to form hydrogels with 19, due to the lack of efficient When the neutral monomer 24 (Fig. 12b) is added into 23 in
intermolecular crosslinking. methylcyclohexane/1,2-dichloroethane (95:5, v/v), intermolecular

Fig. 11. Graphic representation for supramolecular co-assembly between 19 and 20 to form hydrogels. Reproduced with permission of 2014 Royal Society of Chemistry from
ref. [97].
12 Y. Han et al. / Coordination Chemistry Reviews 414 (2020) 213300

Fig. 12. (a) Graphic representation for the nanowire-to-nanowheel metamorphism process of co-assembly 21/22 via the ligand-substitution reaction. (b) Graphic
representation for the formation of supramolecular alternate co-assembly 23/24, because of the participation of Pt(II)  Pt(II) metal–metal interactions for the hetero-
complexation process. Reproduced with permission of 2008 Wiley-VCH Verlag GmbH from ref. [99], and 2018 Royal Society of Chemistry from ref. [101].

donor–acceptor interactions are prone to form due to their similar partmentalized agent. Considering the presence of non-emissive
p-surfaces. Meanwhile, the reduced steric hindrance between 23 gold(III) headgroup and hydrophilic polyethylene glycol tail, 26
and 24 gives rise to the formation of Pt(II)–Pt(II) metal–metal tends to assemble into micelles in water, which render synergistic
interactions (Fig. 12b), which prevails the hetero-complexation rigidifying and hydrophobic shielding effects to the platinum(II)
strength over the homo-complexation one. As a consequence, the terpyridine complexes. Accordingly, it gives rise to enhanced phos-
alternate arrangement exists as the thermodynamic stable state phorescent emission of the platinum(II) complexes 25a (or 25b) in
for co-assembly 23/24, with the appearance of MMLCT absorption aqueous environment. More importantly, the 25a-26 (or 25b-26)
band between 550 and 700 nm. Such an ‘‘emergence upon co- hetero-complexation prevails over the homo-complexation ones,
assembly” property is further applied for red-light irradiated oxi- ascribed to the strong p-stacking strength between neutral gold
dation of a-terpinene to ascaridole, which is unattainable for each (III) headgroup on 26 and positively-charged platinum(II) phos-
of the individual species. By utilizing the similar concept, Yam and phors 25a–b. As a result, the two platinum(II) phosphors are den-
coworker have also realized two-component organoplatinum(II) sely packed, yet separately partitioned via the compartmentalized
co-assemblies in aqueous solution [103]. agent 26. Highly efficient light harvesting and energy transfer can
Very recently, our research group has further developed a tern- be realized in aqueous medium (Fig. 13), by taking advantage of
ary co-assembled system in aqueous medium, consisting of triplet exciton stabilization and ordered spatial organization for
positively-charged Pt(terpy) complexes 25a–b and neutral gold D-A phosphors 25a–b. Therefore, the compartmentalization effect
(III) amphiphile 26 (Fig. 13) [104]. The purpose of the ternary co- involved in the ternary co-assembled system strategy provides a
assembled system is to construct artificial light-harvesting systems feasible approach toward color-tunable phosphorescent nanoma-
with high triplet energy transfer efficiency. In general, several chal- terials in aqueous media.
lenges need to be addressed to realize triplet energy transfer in
water. Specifically, the phosphors should be shielded from water 2.7. Hybrid co-assembly with the presence of Pt(II) complexes
and molecular oxygen, which facilitate to maintain intense emis-
sion intensity. Moreover, the donor–acceptor phosphors should As discussed in Section 2.6, supramolecular co-assembly takes
be organized in close proximity, yet simultaneously avoiding direct place between structurally similar organoplatinum(II) compounds.
homo- and hetero-interactions to minimize the potential energy Besides, co-assembly can be also realized for structurally dissimilar
losses. To meet these requirements, 26 is employed as the com- components, by programming directional non-covalent bonds
Y. Han et al. / Coordination Chemistry Reviews 414 (2020) 213300 13

species [106–108]. The PAA block is capable of neutralizing the


cationic Pt(terpy) 27 (Fig. 14). In the meantime, the PEG blocks
not only provide excellent solubility in the mixture of
acetonitrile-methanol–water (1:1:8, v/v/v), but play a crucial role
in inhibiting the growth of nanostructures in transverse direction.
Accordingly, uniform core–shell nanofibers are obtained for
27/PEG45-b-PAA69 co-assembly as the kinetically trapped product
(Fig. 14). Inside the nanofibers, cylindrical core is formed by PAA
blocks and hexagonally packed molecular columns of 27 along
the fiber long axis. Either sonication or thermal annealing disrupts
the core–shell nanofibers, and thereby leads to the formation of
short patchy nanofibers (Fig. 14). The side and end patches direct
transverse and longitudinal self-assembly of the short patchy
nanofibers to form core–shell nanobelts as the thermodynamically
stable state (Fig. 14) [107,108]. On this basis, it is also
capable of constructing segmented heterojunction materials, by
co-assembling different type of organoplatinum(II) complexes
with PEG-b-PAA [106].
Apart from surfactants and polymers, biomacromolecules such
as nucleotides and proteins can also co-assemble with organoplat-
inum(II) complexes [109,110]. Recently, De Cola, Cornelissen and
co-workers have reported a fascinating system comprising plat-
inum(II) amphiphiles (28 or 29) and the coating protein (CP) of
cowpea chlorotic mottle virus (Fig. 15) [110]. The latter one is pos-
itively charged, which tends to complex with the negatively
charged platinum(II) amphiphiles 28–29 via electrostatic interac-
tions at the neutral buffer solution. Interestingly, by tuning the side
Fig. 13. Graphic representation for the triplet energy transfer of ternary chains on 28–29, morphology of the final co-assembled structures
co-assembly 26/25a/25b in aqueous environment. Reproduced with permission of can be modulated. In detail, co-assembly between 28 and CP leads
2019 Nature Publishing Group from ref. [104].
to the presence of spherical particles, with the diameters of
approximately 18 nm (Fig. 15). It is rationalized that protein–
between organoplatinum(II) complexes and the complementary protein interactions dominates for the 28/CP co-assembly process,
agents. For example, electrostatic interactions exist between since the aggregates derived from 28 itself are evidently smaller
positively-charged alkynyl Pt(terpy) and negatively-charged poly- than the inner cavity of icoshahedral protein cage (ca. 21 nm). In
electrolyte such as poly(acrylic acid) carboxylate. Yam’s group stark contrast, co-assembly between 29 and CP generates rod-
has proposed the ‘‘polyelectrolyte-induced assembly” concept, by like nanostructures (length: 100–500 nm, thickness: 18 nm,
taking advantage of multivalent electrostatic forces between Fig. 15). It lies in the formation of long (>50 nm) and thin
organoplatinum(II) complexes and the anionic polymer [84,105]. (~6 nm) rods for the 29 self-aggregated structure. Accordingly,
Simultaneously, MMLCT bands appear due to the presence of Pt the electrostatic interactions between 29 and CP force the forma-
(II)–Pt(II) metal–metal interactions in the co-assembled structure. tion of non-icosahedron morphology.
In recent years, the same group has employed poly(ethylene Besides, hybrid organic–inorganic system can be attained,
glycol)-b-poly(acrylic acid) (PEG45-b-PAA69) as the co-assembled by employing organoplatinum complexes as one of the

Fig. 14. Graphic representation for conversion from the kinetically trapped nanofibers to thermodynamically stable nanobelts upon sonicating co-assembly 27/PEG45-b-
PAA69. Reproduced with permission of 2017 Elsevier Sciencedirect from [107].
14 Y. Han et al. / Coordination Chemistry Reviews 414 (2020) 213300

Fig. 15. Graphic representation for supramolecular co-assembly systems between coating protein of cowpea chlorotic mottle virus and platinum(II) amphiphiles 28 or 29.
Reproduced with permission of 2018 American Chemical Society from [110].

a better passivation with CTAB. As a consequence, 30 prefers to


attach on the ends of GNRs. In an extensive study, side-by-side
assembly of GNRs have also been realized by the same group, by
rational designing of negatively-charged Pt(terpy) complex 31
(Fig. 16) [112]. The sulfonate moieties on 31 feature high cation-
binding affinity, which preferentially bind to the ammonium head
groups of CTAB along the sides of the GNRs. Accordingly, it brings
the neighboring GNRs together in a side-by-side manner, which
can be visualized by electron microscopy (Fig. 16). The study
demonstrates that well-ordered inorganic–organic hybrid assem-
blies can be achieved via the elaborate choice of organoplatinum
(II) building blocks.

3. Functionality of organoplatinum(II)-based supramolecular


nanostructures

3.1. Sensing materials


Fig. 16. Graphic representation for the end-to-end and side-by-side assembly of
gold nanorods via the assistance of platinum(II) complexes 30 and 31, respectively. During supramolecular assembly/disassembly of organoplat-
Reproduced with permission of 2016 American Chemical Society from [111], and inum(II) complexes, the low-energy MMLCT transition signals
2018 Wiley-VCH Verlag GmbH from ref. [112]. emerge and disappear in a reversible manner. Taking advantage
of this property, together with the aforementioned
‘‘polyelectrolyte-induced assembly” concept [105], Yam and
co-assembling building blocks. Yam and co-workers have designed coworkers have exploited the applications of organoplatinum(II)-
a positively-charged Pt(terpy) complex 30, consisting of a thioac- based assemblies in chemo-sensing field. For example, they con-
etate unit on the periphery (Fig. 16) [111]. Upon adding pyrro- struct a hybrid co-assembling system {poly(phenylene ethynylene
lidine, the in situ deprotected form of 30 would preferentially sulfonate) (abbreviated as PPE-SO 3 ) and [Pt(terpy){C„CAC6H4-
attach to the active ends of the gold nanorods (GNRs). As a conse- ACH2N(CH3)3}](OTf)2)}, with the involvement of Förster resonance
quence, GNRs aggregate into chain-like structures via end-to-end energy transfer (FRET) character [113]. When human serum albu-
assembly (Fig. 16), which is mediated by non-covalent Pt(II)–Pt min (HSA) is added, it binds strongly with PPE-SO 3 , and thereby
(II) metal–metal and p–p stacking interactions provided by the leads to the disassembly between Pt(terpy) and the anionic poly-
Pt(terpy) units. The instinct factor for end-to-end assembly lies mer. Accordingly, label-free detection of HSA is realized in a sensi-
in the stabilization layer of GNRs, which is composed of the inter- tive and selective manner, by monitoring the ratiometric spectral
digitated bilayer of cationic cetrimonium bromide (CTAB). Since changes between MMLCT and MLCT/LLCT signals. The label-free
CTAB displays a better binding affinity towards the {1 0 0} and sensing strategy has also been applied by the same group to the
{1 1 0} side facets than the {1 1 1} end facets, the GNRs sides have specific detection of glucose [114], lysozyme [115], thrombin,
Y. Han et al. / Coordination Chemistry Reviews 414 (2020) 213300 15

collapse of organogels. Remarkably, chiral discrimination of (R)-


and (S)-2,20 -bis(diphenylphosphino)-1,10 -binaphthyl (abbreviated
as (R)- and (S)-binap, respectively) enantiomers is achieved,
through a visually enantio-controlled breakup of the supramolecu-
lar gels (Fig. 17). The chiral recognition behaviors are strongly cor-
related with metal-coordination capability of the chiral phosphine
ligands. (R)-Binap phosphine coordinates efficiently to the plat-
inum center (Fig. 17), leading to the breakup of intermolecular Pt
(II)–Pt(II) metal–metal and p–p stacking interactions. On the con-
trary, ligand exchange hardly occurs for (S)-binap phosphine ligand
(Fig. 17), because of its chirality mismatching with the chiral
supramolecular gels.

3.2. Emitting materials

Upon supramolecular assembly of organoplatinum(II) com-


plexes, it decreases non-radiation decay rate and thereby enhances
the emission intensity. Additionally, spin-coating and ink-jet print-
Fig. 17. Graphic representation for visual chiral recognition of chiral binap through ing techniques can be employed to fabricate ordered films. It thus
enantioselective metallogel collapsing. Reproduced with permission of 2011 Wiley- avoids the expensive vacuum deposition technique encountered
VCH Verlag GmbH from ref. [119].
for small molecules, and thereby reduces the cost for manufactur-
ing [120]. Based on these considerations, organoplatinum(II)-based
human telomere DNA [116], protamine [117], heparin [118], and nanostructures display promising prospects for emitting applica-
etc. Readers could refer to a nice review for more detailed informa- tions. De Cola and coworkers have reported a neutral organoplat-
tion [20]. inum(II) compound 33 (Fig. 18a), which forms nanofibers and
In addition, an interesting visual chiral recognition system has gels in hexane with aggregation-induced green emission character
been reported by Tu and co-workers [119]. Briefly, 32 (Fig. 17) is [PLQY (photoluminescence quantum yield) = 90% at the gelation
designed with a chloro alkynylplatinum(II) core. The compound state, and 87% in thin film] [121]. When 10 wt% of 33 is doped in
is capable of forming supramolecular organogels, by entrapping the emissive layer, the maximum current efficiency (gc,max) is
various aprotic solvents such as chloroform, dichloroethane, 15.6 cd A1 (4.5 Lm W1) at a brightness of 203 cd m2. The max-
dimethylacetamide, and benzonitrile. The chloride unit on 32 imum brightness of OLEDs is determined to be 11360 cd m2.
undergoes ligand exchange with the bulky phosphine ligands. It Hence, the ‘‘turn-on” luminescent character of organoplatinum
gives rise the huge steric hindrance, and thereby inducing the (II)-based nanostructures is attractive for OLEDs applications.

Fig. 18. (a) Structures of 33 and 34, together with graphic representation of aggregation-induced electrochemiluminescence behavior of 34. (b) Circularly polarized
luminescent behaviours of organoplatinum(II)-based enantiomers (+)-35 and ()-35. Reproduced with permission of 2017 American Chemical Society from [122], and 2015
American Chemical Society from [124].
16 Y. Han et al. / Coordination Chemistry Reviews 414 (2020) 213300

Later on, the same group has also investigated aggregation- The resulting high-molecular-weight supramolecular polymers
induced electrochemiluminescence (AIECL) behaviors of are capable of forming electrospun microfibers with uniform
organoplatinum(II) complexes 34 (Fig. 18a) on the surface of elec- geometry and smooth surface. It enables light propagation with
trode [122]. Although molecularly dissolved state dominates for 34 high anitrosopy and extremely low scattering loss (0.008 dB lm1).
at the ambient state, gentle grinding of its powder facilitates the Hence, the study provides a novel avenue toward high-
conversion to the aggregated structure. It is accompanied by the performance optical waveguiding materials, by combining
emission color change from blue to orange. Remarkably, the aggre- bottom-up supramolecular polymerization and top-down electro-
gated sample of 34 shows 20-fold enhancement in ECL emission spinning techniques in an elaborate manner. Giving that chirality
intensity, when comparing to the blue-emissive sample prior to signals emerges at the self-assembled state, it is expected that
the grinding process (Fig. 18a). Unambiguously, supramolecular flourescent supramolecular polymers derived from 2b also possess
assembly is of particularly importance toward emissive materials CPL signal, which is now underway in our labotorary.
with the high ECL intensity.
Circularly polarized luminescent (CPL) materials have also 3.3. Semi-conductive materials
aroused significant interests in recent years, because of their
potential applications in 3D display and optical identification sen- Organic semiconductors have aroused tremendous interest
sor [123]. The prerequisite of CPL emitters requires the embed- because of their potential application in field-effect transistors
ment of emission and chirality characters in a concurrent (FET) and photovoltaic cells [127]. Supramolecular assembly of
manner. In this regard, You’s group has reported the p-conjugated molecules represent a plausible route to provide bulk
organoplatinum(II)-based enantiomers (+)-35 and ()-35 semi-conducting materials. In this context, Che and coworkers
(Fig. 18b), featuring the asymmetric pinene on the tridentate have reported a series of alkynylplatinum(II) complexes with the
cyclometalated ligand [124]. The chiral pinene units benefit for tridentate 1,5-bis(20 -pyridine)benzene ligands [128]. For one of
the formation of helical supramolecular aggregates in water. the synthetic compound 36 (Fig. 19a), it is capable of assembling
Accordingly, CPL signals emerge for both enantiomers with the into quasi-two-dimensional nanosheets. Depending on X-ray mea-
opposite helicity. The glum values around the maximum emission surement, infinite Pt–Pt chains exist with a short intermetal con-
wavelength are calculated to be 0.0012 and –0.0018 for (+)-35 tact of 3.383 Å (Fig. 19a). Neighboring molecules are arranged in
and ()-35 at 295 K, respectively (Fig. 18b). Organoplatinum(II)- a staggered fashion with a torsion angle of 49.4° around the Pt–
based CPL materials are also developed by Haino and coworkers Pt axis. The hole-transporting properties are further studied for
[125,126]. When tris(phenylisoxazole)-linked point-chiral alkyl the drop-casting nanosheets of 36. Upon switching on and off the
chains are attached to the Pt(II) phenylbipyridine core, it shows white light (intensity: 40 mW cm2) for every 5 s, a transient chan-
aggregation-induced emission enhancement character. Accord- nel current is detected when drain-source voltage (VDS) and gate
ingly, the resulting assemblies feature a pronounced glum value voltage (VG) are ±40 V. More intriguingly, the field-effect transistor
(~0.01), despite the fact that neither CD nor CPL signal is observed device shows an increase in the density of forward and backward
at the monomeric state. current after photoexcitation (Fig. 19a).
As discussed above, our research group has successfully assem- Notably, the charge-transporting properties can be optimized
bling platinum(II) acetylides 2b (Fig. 2b) into fluorescent via the rational choice of ligand unit on organoplatinum(II) com-
supramolecular polymers via the cooperative mechanism [38]. plexes. For another compound 37 (Fig. 19b) designed by the same

Fig. 19. (a) Crystalline packing structure of 36 and the output characteristics of FET device fabricated from 36 (VDS: drain–source voltage, VG: gate voltage). (b) Crystal packing
structure and p-channel characteristic of the 37 micro-crystal. Reproduced with permission from 2009 Wiley-VCH Verlag GmbH from ref. [128], and 2011 Royal Society of
Chemistry from ref. [129].
Y. Han et al. / Coordination Chemistry Reviews 414 (2020) 213300 17

Accordingly, the electron mobility (le) value reaches to the


remarkable value of 20 cm2 V1 s1 for FET device measurements
(Fig. 19b).

3.4. Biomedical materials

As widely documented, cisplatin and carboplatin display high


chemotherapeutic efficiency toward cancer cells. It stimulates the
development of novel organoplatinum(II)-based anti-tumor drugs.
Che’s group has reported some promising candidates in this field
[130,131]. For example, supramolecular hydrogel derived from
38 (Fig. 20) shows sustained and pH-dependent releasing behav-
iors. Under the acidic conditions, the survival percentage for HeLa
cells is determined to be ca. 30% after incubation for 6 h (Fig. 20).
However, it takes 72 h of incubation treatment to reach 29% cell
survival under neutralized condition (Fig. 20). Accordingly, the
acidic condition is favorable for the enhanced cytotoxicity, which
is associated with pH-dependent assembling capability of 38 in
water. In detail, 38 is prone to accumulate in the acidic lysosomes
of HeLa cells, leading to the increase in lysosomal membrane
Fig. 20. Supramolecular self-assembly and gelation behaviours of 38, together with permeability. With the further combination of strong emission
time-dependent HeLa cell survival upon treating with the hydrogel or cisplatin and chemotherapeutic characters, organoplatinum(II)-based
under the neutral or acidic incubation conditions. Reproduced with permission of supramolecular nanostructures have also demonstrated promising
2015 Royal Society of Chemistry from [131].
prospects in the theranostics fields.

group [129], two pyrazolyl units are incorporated into the main 3.5. Other potential applications
ligand, which enable the formation of intermolecular hydrogen
bonds. In the crystal structure (Fig. 19b), such non-covalent forces Organoplatinum(II)-based supramolecular nanostructures can
are harmonized with extended p–p and Pt(II)–Pt(II) metal–metal be also endowed with other functionalities. For example, Yam
interactions to align the planar cations into a long-range ordered and co-workers have designed a homotritopic monomer 39
quasi-1-D columnar structure. The co-facial configuration of 37 (Fig. 21a), which tends to form hexagonal honeycomb-like porous
enhances the charge-hopping process inside the single crystal. structure [132]. Intriguingly, the resulting polycationic

Fig. 21. (a) Hexagonal honeycomb-like porous networks assembled from 39, together with the adsorption performance toward the diverse cations and anions. (b) Hybrid
composites ZJU-28/40 and ZJU-28/41, together with the TON values for photo-catalyzed dehydrogenation reaction of 1-phenylethanol. Reproduced with permission of 2019
American Chemistry Society from [132], and 2018 Royal Society of Chemistry from [133].
18 Y. Han et al. / Coordination Chemistry Reviews 414 (2020) 213300

two-dimensional supramolecular polymers selectivity recognize organoplatinum(II)-based systems. A variety of possible orienta-
anionic platinum(II) complexes such as PtCl2 4 from the guest mix- tion modes can be generated, including segregated stacks, block
ture. The high affinity toward PtCl24 is correlated with the strong co-assembly, randomly mixing co-assembly, and alternately mix-
Pt–Pt interactions. The supramolecular polymeric material can be ing co-assembly. It is mainly ascribed to the inter- and intra-
processed into composite membrane on the surface of polycarbon- chain exchange of monomers, because of the kinetic lability of
ate filter support, which also exhibits encapsulation preference for non-covalent forces involved in the co-assembly processes. In this
the anionic platinum(II) complexes (the adsorption capacity: regard, domain control is a challenging task for supramolecular
33 mg g1, Fig. 21a). chemists in this field. Although we and other research group have
Besides, Che and co-workers have reported an interesting successfully directed the formation of strictly alternate co-
hybrid co-assembly system for photo-catalytic application assemblies [101–103], other types of precise orientation modes
(Fig. 21b) [133]. For metal organic frameworks (ZJU-28), their such as multi-block copolymers have yet to be explored.
one-dimensional channels provide ‘‘solid solution” environments. Thirdly, the current reported organoplatinum(II)-based
As a result, it facilitates to encapsulate 40 and 41 (Fig. 21b) in supramolecular assemblies are mainly in the ‘‘static” equilibrium
the supramolecular assembled form. The hybrid composites ZJU- state. In comparison, out-of-equilibrium supramolecular systems,
28/40 and ZJU-28/41 show good performance for the photo- denoting the supply and consumption of energy during the assem-
catalyzed C–H dehydrogenation reactions. Upon photo-irradiation bling processes, are seldom exploited. The out-of-equilibrium dis-
at room temperature, 1-phenylenthanol transforms into acetophe- sipative state could not reach to the global minima, according to
none with high turn-over numbers [TONs: 363 for ZJU-28/40 the energy landscapes of supramolecular systems. Nevertheless,
(1.36% loading); 210 for ZJU-28/40 (1.50% loading); 252 for ZJU- it is highly mimicking of natural assembled systems, facilitating
28/41 (1.32% loading); 160 for ZJU-28/41 (2.41% loading)]. Note- to interconvert rapidly and reversibly between various
worthily, trace amount or even no products can be detected for supramolecular structures. Overall, in-depth understanding of
either organoplatinum(II) photocatalysts or ZJU-28 at the individ- supramolecular assembling behaviors would benefit for the excit-
ual form. It is rationalized that 3MMLCT excited states derived from ing new applications of such organoplatinum(II)-based systems in
the 40 (or 41) self-aggregation process is crucial for the photo- the future.
catalytic reactions, considering that the diradical features of 3-
MMLCT excited states facilitate to hydrogen atom extraction from
C–H bonds. Conflict of interest

4. Conclusion The authors declare no conflict of interest.

In conclusion, the state-of-art progress of long-range ordered Acknowledgments


organoplatinum(II)-based nanostructures has been summarized.
Square-planar organoplatinum(II) complexes are regarded as ideal This work is supported by the National Natural Science Founda-
supramolecular building blocks, which facilitate to incorporate a tion of China (Grants 21922110, 21871245, and 21674106), the
variety of non-covalent forces such as p–p stacking, metal–metal, Fundamental Research Funds for the Central Universities (Grant
and hydrogen bonding interactions. Delicate manipulation of these WK3450000004), the CAS Youth Innovation Promotion Association
non-covalent bonds give rise to diverse assembling modes. With (Grant 2015365).
the deeper understanding of the assembling mechanism and path-
way, it is capable of constructing supramolecular nanostructures
with the tailored properties. By combining the advantageous from Appendix A. Supplementary data
organoplatinum(II) complexes (such as fascinating photo-physical
and therapeutic properties) and long-range ordered nanostruc- Supplementary data to this article can be found online at
tures (such as excellent solution processability and the collective https://doi.org/10.1016/j.ccr.2020.213300.
signal amplification), the resulting systems have displayed promis-
ing prospects in sensing, optoelectronics, biomedical, and environ- References
mental fields.
Despite all of these achievements, there are still some issues to [1] I.A. Haque, L. Xu, R.A. Al-Balushi, M.K. Al-Suti, R. Ilmi, Z. Guo, M.S. Khan, W.-Y.
be solved from the perspective of supramolecular chemistry. Wong, P.R. Raithby, Cyclometallated tridentate platinum(II) arylacetylide
complexes: old wine in new bottles, Chem. Soc. Rev. 48 (2019) 5547–5563.
Firstly, monodispersity is in keen pursuit for long-range ordered [2] W. Wang, H.-B. Yang, Linear neutral platinum-acetylide moiety: beyond the
organoplatinum(II) assemblies, since polydisperse nanostructures links, Chem. Commun. 50 (2014) 5171–5186.
are obtained for the vast majority of the reported examples. In this [3] Y. Chi, P.-T. Chou, Transition-metal phosphors with cyclometalating ligands:
fundamentals and applications, Chem. Soc. Rev. 39 (2010) 638–655.
respect, ‘‘seeded growth approach” proposed by Manners repre- [4] W. Yang, J. Zhao, Photophysical properties of visible-light-harvesting PtII bis
sents a promising strategy to tackle the problem. To attain this (acetylide) complexes, Eur. J. Inorg. Chem. 34 (2016) 5283–5299.
objective, it requires the adoption of nucleation–elongation coop- [5] L. Kai, G.S.M. Tong, Q. Wan, G. Cheng, W.-Y. Tong, W.-H. Ang, W.-L. Kwong, C.-
M. Che, Highly phosphorescent platinum(II) emitters: photophysics,
erative mechanism for organoplatinum(II)-based assemblies, materials and biological applications, Chem. Sci. 7 (2016) 1653–1673.
meanwhile the supramolecular assembly process should be kinet- [6] J.A.G. Williams, S. Develay, D.L. Rochester, L. Murphy, Optimising the
ically trapped to retard spontaneous nucleation. Accordingly, luminescence of platinum(II) complexes and their application in organic
light emitting devices (OLEDs), Coord. Chem. Rev. 252 (2008) 2596–2611.
incorporation of monomeric species takes place only from the [7] C. Liao, A.H. Shelton, K.-Y. Kim, K.S. Schanze, Organoplatinum chromophores
seed’s ends to provide defined aggregation length, the phenomena for application in high-performance nonlinear absorption materials, ACS
of which are similar to that of covalent living polymerization. Appl. Mater. Interfaces 3 (2011) 3225–3238.
[8] T.C. Johnstone, K. Suntharalingam, S.J. Lippard, The next generation of
Although Manners and co-workers have demonstrated one fasci-
platinum drugs: targeted Pt(II) agents, nanoparticle delivery, and Pt(IV)
nating example in this field [94], adaptability of the strategy to prodrugs, Chem. Rev. 116 (2016) 3436–3486.
other organoplatinum(II)-based assembled systems should be [9] K. Ariga, J.P. Hill, M.V. Lee, A. Vinu, R. Charvet, S. Acharya, Challenges and
examined. breakthroughs in recent research on self-assembly, Sci. Technol. Adv. Mater. 9
(2008) 014109.
Secondly, in addition to mono-component supramolecular sys- [10] E. Busseron, Y. Ruff, E. Moulin, N. Giuseppone, Supramolecular self-
tem, more and more attentions should be paid to the co-assembled assemblies as functional nanomaterials, Nanoscale 5 (2013) 7098–7140.
Y. Han et al. / Coordination Chemistry Reviews 414 (2020) 213300 19

[11] L. Yang, X. Tan, Z. Wang, X. Zhang, Supramolecular polymers: historical [41] M. Chen, C. Wei, J. Tao, X. Wu, N. Huang, G. Zhang, L. Li, Supramolecular
development, preparation, characterization, and functions, Chem. Rev. 115 polymers self-assembled from trans-bis(pyridine) dichloropalladium and
(2015) 7196–7239. platinum complexes, Chem. Eur. J. 20 (2014) 2812–2818.
[12] R. Hu, D.-H. Xin, A.-J. Qin, B.Z. Tang, Polymers with aggregation-induced [42] M. Chen, C. Wei, X. Wu, M. Khan, N. Huang, G. Zhang, L. Li, Metallogels self-
emission characteristics, Act. Polym. Sin. 49 (2018) 132–144. assembled from linear rod-like platinum complexes: influence of the linkage,
[13] K.M.-C. Wong, V.K.-M. Au, V.W.W. Yam, Non-covalent metal-metal Chem. Eur. J. 21 (2015) 4213–4217.
interactions, Compr. Inorg. Chem. II (Second Ed.) 8 (2013) 59–130. [43] N. Bämer, K.K. Kartha, N.K.N. Allampally, S. Yagai, R.Q. Albuquerque, G.
[14] M. Yoshida, M. Kato, Regulation of metal-metal interactions and chromic Fernández, Exploiting coordination isomerism for controlled self-assembly,
phenomena of multi-decker platinum complexes having p-systems, Coord. Angew. Chem. Int. Ed. 58 (2019) 15626–15630.
Chem. Rev. 355 (2018) 101–115. [44] N.K. Allampally, M.J. Mayoral, S. Chansai, M.C. Lagunas, C. Hardacre, V.
[15] A. Aliprandi, D. Genovese, M. Mauro, L. de Cola, Recent advances in Stepanenko, R.Q. Albuquerque, G. Fernández, Control over the self-assembly
phosphorescent Pt(II) complexes featuring metallophilic interactions: modes of PtII complexes by alkyl chain variation: from slipped to parallel p-
properties and applications, Chem. Lett. 44 (2015) 1152–1169. stacks, Chem. Eur. J. 22 (2016) 7810–7816.
[16] J. Schill, A.P.H.J. Schenning, L. Brunsveld, Self-assembled fluorescent [45] C. Rest, M.J. Mayoral, K. Fucke, J. Schellheimer, V. Stepanenko, G. Fernández,
nanoparticles from p-conjugated small molecules: en route to biological Self-assembly and (hydro)gelation triggered by cooperative p-p and
applications, Macromol. Rapid. Comm. 36 (2015) 1306–1321. unconventional CH-X hydrogen bonding interactions, Angew. Chem. Int. Ed.
[17] K.M.-C. Wong, V.W.-W. Yam, Self-assembly of luminescent alkynylplatinum 53 (2014) 700–705.
(II) terpyridine complexes: modulation of photophysical properties through [46] V.W.-W. Yam, K.M.-C. Wong, N. Zhu, Solvent-induced aggregation through
aggregation behavior, Acc. Chem. Res. 44 (2011) 424–434. metal-metal/p-p interactions: large solvatochromism of luminescent
[18] T.-F. Fu, L. Ao, Z.-C. Gao, X.-L. Zhang, F. Wang, Advances on supramolecular organoplatinum(II) terpyridine complexes, J. Am. Chem. Soc. 124 (2002)
assembly of cyclometalated platinum(II) complexes, Chin. Chem. Lett. 27 6506–6507.
(2016) 1147–1154. [47] A.Y.-Y. Tam, K.M.-C. Wong, N. Zhu, G. Wang, V.W.-W. Yam, Luminescent
[19] L. Herkert, A. Sampedro, G. Fernández, Cooperative self-assembly of discrete alkynylplatinum(II) terpyridine metallogels stabilized by Pt-Pt, p-p, and
metal complexes, CrystEngComm. 18 (2016) 8813–8822. hydrophobic-hydrophobic interactions, Langmuir 25 (2009) 8685–8695.
[20] K.M.-C. Wong, M.M.-Y. Chan, V.W.-W. Yam, Supramolecular assembly of [48] A.Y.-Y. Tam, K.M.-C. Wong, G. Wang, V.W.-W. Yam, Luminescent metallogels
metal-ligand chromophores for sensing and phosphorescent OLED of platinum(II) terpyridine complexes: interplay of Pt-Pt, p-p, and
applications, Adv. Mater. 26 (2014) 5558–5568. hydrophobic-hydrophobic interactions on gel formation, Chem. Commun.
[21] V.W.-W. Yam, V.K.-M. Au, S.Y.-L. Leung, Light-emitting self-assembled (2007) 2028–2030.
materials based on d8 and d10 transition metal complexes, Chem. Rev. 115 [49] A.Y.-Y. Tam, V.W.-W. Yam, Recent advances in metallogels, Chem. Soc. Rev.
(2015) 7589–7728. 42 (2013) 1540–1567.
[22] M. Mauro, A. Aliprandi, D. Septiadi, N.S. Kehra, L. De Cola, When self-assembly [50] K. Liu, L. Meng, S. Mo, M. Zhang, Y. Mao, X. Cao, C. Huang, T. Yi, Colour change
meets biology: luminescent platinum complexes for imaging applications, and luminescence enhancement in as cholesterol-based terpyridine platinum
Chem. Soc. Rev. 43 (2014) 4144–4166. metallogel via sonication, J. Mater. Chem. C 1 (2013) 1753–1762.
[23] D. Zhao, J.S. Moore, Nucleation-elongation: a mechanism for cooperative [51] F. Camerel, R. Ziessel, B. Donnio, C. Bourgogne, D. Guillon, M. Schmutz, C.
supramolecular polymerization, Org. Biomol. Chem. 1 (2003) 3471–3491. Iacovita, J.-P. Bucher, Formation of gels and liquid crystals induced by Pt-Pt
[24] T.F.A. de Greef, M.M.J. Smulders, M. Wolffs, A.P.H.J. Schenning, R.P. Sijbesma, and p-p* interactions in luminescent r-alkynyl platinum(II) terpyridine
E.W. Meijer, Supramolecular polymerization, Chem. Rev. 109 (2009) 5687– complexes, Angew. Chem. Int. Ed. 119 (2007) 2659–2662.
5754. [52] C. Po, Z. Ke, A.Y.-Y. Tam, H.-F. Chow, V.W.-W. Yam, A platinum(II) terpyridine
[25] C. Kulkarni, S. Balasubramanian, S.J. George, What molecular features govern metallogel with an L-valine-modified alkynyl ligand: interplay of Pt-Pt, p-p,
the mechanism of supramolecular polymerization?, ChemPhysChem 14 and hydrogen-bonding interactions, Chem. Eur. J. 19 (2013) 15735–15744.
(2013) 661–673. [53] A.Y.-Y. Tam, K.M.-C. Wong, V.W.-W. Yam, Unusual luminescence
[26] P.A. Korevaar, T.F.A. de Greef, E.W. Meijer, Pathway complexity in p- enhancement of metallogels of alkynylplatinum(II) 2, 6-bis(N-
conjugated materials, Chem. Mater. 26 (2014) 576–586. alkylbenzimidazol-2’-yl)pyridine complexes upon a gel-to-sol phase
[27] A. Sorrenti, J. Leira-lglesias, A.J. Markvoort, T.F.A. de Greef, T.M. Hermans, transition at elevated temperatures, J. Am. Chem. Soc. 131 (2009) 6253–6260.
Non-equilibrium supramolecular polymerization, Chem. Soc. Rev. 46 (2017) [54] K.-C. Chang, J.-L. Lin, Y.-T. Shen, C.-Y. Hung, C.-Y. Chen, S.-S. Sun, Synthesis
5476–5490. and photophysical properties of self-assembled metallogels of platinum(II)
[28] J. Matern, Y. Dorca, L. Sánchez, G. Fernández, Revising complex acetylide complexes with elaborate long-chain pyridine-2,6-dicarboxamides,
supramolecular polymerization under kinetic and thermodynamic control, Chem. Eur. J. 18 (2012) 1312–1321.
Angew. Chem. Int. Ed. 58 (2019) 16730–16740. [55] L. Li, N. Zhou, H. Kong, X. He, Controlling the supramolecular polymerization
[29] T. Cardolaccia, Y. Li, K.S. Schanze, Phosphorescent platinum acetylide and metallogel formation of Pt(II) complexes via delicate tuning of non-
organogelators, J. Am. Chem. Soc. 130 (2008) 2535–2545. covalent interactions, Polym. Chem. 10 (2019) 5465–5472.
[30] L.-J. Chen, J. Zhang, J. He, X.-D. Xu, N.-W. Wu, D.-X. Wang, Z. Abliz, H.-B. Yang, [56] W. Lu, Y. Chen, V.A.L. Roy, S.S.-Y. Chui, C.-M. Che, Supramolecular polymers
Synthesis of platinum acetylide derivatives with different shapes and their and chromonic mesophases self-organized from phosphorescent cationic
gel formation behavior, Organometallics 30 (2011) 5590–5594. organoplatinum(II) complexes in water, Angew. Chem. Int. Ed. 48 (2009)
[31] X.-D. Xu, J. Zhang, L.-J. Chen, X.-L. Zhao, D.-X. Wang, H.-B. Yang, Large-scale 7621–7625.
honeycomb microstructures constructed by platinum-acetylide gelators [57] Y. Ai, Y. Li, H.L.-K. Fu, A.K.-W. Chan, V.W.-W. Yam, Aggregation and tunable
through supramolecular self-assembly, Chem. Eur. J. 18 (2012) 1659–1667. color emission behaviors of L-glutamine-derived platinum(II) bipyridine
[32] X.-D. Xu, J. Zhang, X. Yu, L.-J. Chen, D.-X. Wang, T. Yi, F. Li, H.-B. Yang, Design complexes by hydrogen-bonding, p-p stacking and metal-metal interactions,
and preparation of platinum-acetylide organogelators containing ethynyl- Chem. Eur. J. 25 (2019) 5251–5258.
pyrene moieties as the main skeleton, Chem. Eur. J. 18 (2012) 16000–16013. [58] J. Chen, L. Ao, C. Wei, C. Wang, F. Wang, Self-assembly of platinum(II) 6-
[33] B. Jiang, L.-J. Chen, L. Xu, S.-Y. Liu, H.-B. Yang, A series of new star-shaped or phenyl-2,2’-bipyridine complexes with solvato- and iono-chromic
branched platinum-acetylide derivatives: synthesis, characterization, and phenomena, Chem. Commun. 55 (2018) 229–232.
their aggregation behavior, Chem. Commun. 49 (2013) 6977–6979. [59] N.K. Allampally, M. Bredol, C.A. Strassert, L. De Cola, Highly phosphorescent
[34] Y.-Y. Ren, N.-W. Wu, J. Huang, Z. Xu, D.-D. Sun, C.-H. Wang, L. Xu, A neutral supramolecular hydrogels based on platinum(II) emitters, Chem. Eur. J. 20
branched platinum-acetylide complex possessing a tetraphenylethylene (2014) 16863–16868.
core: preparation of a luminescent organometallic gelator and its [60] A. Colombo, F. Fiorini, D. Sepeiadi, C. Dragonetti, F. Nisic, A. Valore, D. Roberto,
unexpected spectroscopic behavior during sol-to-gel transition, Chem. M. Mauro, L. De Cola, Neutral N^C^N terdentate luminescent Pt(II)
Commun. 51 (2015) 15153–15156. complexes: their synthesis, photophysical proeprties, and bioimaging
[35] J. Zhang, N.-W. Wu, X.-D. Xu, Q.-J. Li, C.-H. Wang, H. Tan, L. Xu, Branched applications, Dalton Trans. 44 (2015) 8478–8487.
platinum-acetylide complexes: synthesis, properties, and their aggregation [61] L.-B. Xing, S. Yu, X.-J. Wang, G.-X. Wang, B. Chen, L.-P. Zhang, C.-H. Tung, L.-Z.
behavior, RSC Adv. 4 (2014) 16047–16054. Wu, Reversible multistimuli-responsive vesicles formed by an amphiphilic
[36] Y.-J. Tian, E.W. Meijer, F. Wang, Cooperative self-assembly of platinum(II) cationic platinum terpyridine complex with a ferrocene unit in water, Chem.
complexes, Chem. Commun. 49 (2013) 9197–9199. Commun. 48 (2012) 10886–10888.
[37] Z. Gao, J. Zhu, Y. Han, X. Lv, X. Zhang, F. Wang, Ligand effects on cooperative [62] Y. Hu, K.H.-Y. Chan, C.Y.-S. Chung, V.W.-W. Yam, Reversible thermos-
supramolecular polymerization of platinum(II) acetylide complexes, Ploym. responsive luminescent metallo-supramolecular triblock copolymers based
Chem. 7 (2016) 5763–5767. on platinum(II) terpyridine chromophores with unusual aggregation
[38] X. Wang, Y. Han, Y. Liu, G. Zou, F. Wang, Cooperative supramolecular behavior and red-near-infrared (NIR) emission upon heating, Dalton Trans.
polymerization of fluorescent platinum acetylide for optical waveguide 40 (2011) 12228–12234.
applications, Angew. Chem. Int. Ed. 56 (2017) 12466–12470. [63] S.Y.-L. Leung, K.M.-C. Wong, V.W.-W. Yam, Self-assembly of alkynylplatinum
[39] Z. Gao, Y. Han, F. Wang, Cooperative supramolecular polymers with (II) terpyridine amphiphiles into nanostructures via steric control and metal-
anthracene-endoperoxide photo-switching for fluorescent anti- metal interactions, Proc. Natl. Acad. Sci. 113 (2016) 2845–2850.
counterfeiting, Nat. Commun. 9 (2018) 3977. [64] M.H.-Y. Chan, S.Y.-L. Leung, V.W.-W. Yam, Controlling self-assembly
[40] Y. Han, M. Liu, R. Zhong, Z. Chen, M. Zhang, F. Wang, Photo-responsiveness of mechanisms through rational molecular design in oligo(p-
anthracene-based supramolecular polymers regulated via a r-platinated 4,4- phenyleneethynylene)-containing alkynylplatinum(II) 2,6-bis(N-
difluo-4-bora-3a,4a-diaza-s-indacene photosensitizer, Inorg. Chem. 58 alkylbenzimidazole-2’-yl)pyridine amphiphiles, J. Am. Chem. Soc. 140
(2019) 12407–12414. (2018) 7637–7646.
20 Y. Han et al. / Coordination Chemistry Reviews 414 (2020) 213300

[65] C.Y.-S. Chung, V.W.-W. Yam, Dual pH- and temperature-responsive [88] W. Meng, Q. He, M. Yu, Y. Zhou, C. Wang, B. Yu, B. Zhang, W. Bu, Telechelic
metallosupramolecular block copolymers with tunable critical micelle amphiphilic metallopolymers end-functionalized with platinum(II)
temperature by modulation of the self-assembly of NIR-emissive complexes: synthesis, luminescence enhancement, and their self-assembly
alkynylplatinum(II) complexes induced by changes in hydrophilicity and into flowerlike vesicles and giant flowerlike vesicles, Polym. Chem. 10 (2019)
electrostatic effects, Chem. Eur. J. 19 (2013) 13182–13192. 4477–4484.
[66] N. Komiya, T. Muraoka, M. Iida, M. Miyanaga, K. Takahashi, T. Naota, [89] F. Qu, B. Yang, Q. He, W. Bu, Synthesis of platinum(II) complex end
Ultrasound-induced emission enhancement based on structure-dependent functionalized star polymers: luminescence enhancements and
homo- and hetero-chiral aggregations of chiral binuclear platinum unimolecular micelles in solvents of weakened quality, Polym. Chem. 8
complexes, J. Am. Chem. Soc. 133 (2011) 16054–16061. (2017) 4716–4728.
[67] M. Ikeshita, M. Ito, T. Naota, Variations in the solid-state emissions of [90] Y. Tidhar, H. Weissman, S.G. Wolf, A. Gulino, B. Rybtchinski, Pathway-
clothespin-shaped binuclear trans-bis(salicylaldiminato)platinum(II) with dependent self-asembly of perylene diimide/peptide conjugate in aqueous
halogen functionalities, Eur. J. Inorg. Chem. 31 (2019) 3561–3571. medium, Chem. Eur. J. 17 (2011) 6068–6075.
[68] L.-J. Chen, H.-B. Yang, Construction of stimuli-responsive functional materials [91] A. Aliprandi, A. Mauro, L. De Cola, Controlling and imaging biomimetic self-
via hierarchical self-assembly involving coordination interactions, Acc. Chem. assembly, Nat. Chem. 8 (2015) 10–15.
Res. 51 (2018) 2699–2710. [92] A. Aliprandi, L. Capaldo, C. Bobica, S. Silvestrini, L. De Cola, Effects of the
[69] S.Y.-L. Leung, A.Y.-Y. Tam, C.-H. Tao, H.S. Chow, V.W.-W. Yam, Single-turn molecular design on the supramolecular organization of luminescent Pt(II)
helix-coil strands stabilized by metal-metal and p-p interactions of the complexes, Isr. J. Chem. 59 (2019) 892–897.
alkynylplatinum(II) terpyridine moieties in meta-phenylene ethynylene [93] M.E. Robinson, D.J. Lunn, A. Nazemi, G.R. Whittell, L. de Cola, I. Manners,
foldamers, J. Am. Chem. Soc. 134 (2012) 1047–1056. Length control of supramolecular polymeric nanofibers based on stacked
[70] M.H.-Y. Chan, S.Y.-L. Leung, V.W.-W. Yam, Rational design of multi-stimuli- planar platinum(II) complexes by seeded-growth, Chem. Commun. 51 (2015)
responsive scaffolds: synthesis of luminescent oligo(ethynylpyridine)- 15921–15924.
containing alkynylplatinum(II) polypyridine foldamers stabilized by PtPt [94] M.E. Robinson, A. Nazemi, D.J. Lunn, D.W. Hayward, C.E. Boott, M.-S. Hsiao, R.
interactions, J. Am. Chem. Soc. 141 (2019) 12312–12321. L. Harniman, S.A. Davis, G.R. Whittell, R.M. Richardson, L. de Cola, I. Manners,
[71] M.H.-Y. Chan, M. Ng, S.Y.-L. Leung, W.H. Lam, V.W.-W. Yam, Synthesis of Dimensional control and morphological transformations of supramolecular
luminescent platinum(II) 2,6-bis(N-dodecylbenzimidazol-2’-yl)pyridine polymeric nanofibers based on cofacially-stacked planar amphiphilic
foldamers and their supramolecular assembly and metallogel formation, J. platinum(II) complexes, ACS Nano 11 (2017) 9162–9175.
Am. Chem. Soc. 139 (2017) 8639–8645. [95] L. Herkert, J. Dorste, K.K. Kartha, P.A. Korevaar, T.F.A. de Dreef, M.R. Hansen, G.
[72] S.Y.-L. Leung, V.W.-W. Yam, Hierarchical helices of helices directed by Pt-Pt Fernández, Pathway control in cooperative vs. anti-cooperative
and p-p stacking interactions: reciprocal association of multiple helices of supramolecular polymers, Angew. Chem. Int. Ed. 58 (2019) 11344–11349.
multiple helices of dinuclear alkynylplatinum(II) complex with luminescence [96] A. Langenstroer, K.K. Kartha, Y. Dorca, J. Dorste, V. Stepanenko, R.Q.
enhancement behaviors, Chem. Sci. 4 (2013) 4228–4234. Albuquerque, M.R. Hansen, L. Sánchez, G. Fernández, Unravelling
[73] Y. Ai, M.H.-Y. Chan, A.K.-W. Chan, M. Ng, Y. Li, V.W.-W. Yam, A platinum(II) concomitant packing polymorphism in metallosupramoelcular polymers, J.
molecular hinge with motions visualized by phosphorescence changes, Proc. Am. Chem. Soc. 141 (2019) 5192–5200.
Natl. Acad. Sci. 116 (2019) 13856–13861. [97] X.-S. Xiao, W. Lu, C.-M. Che, Phosphorescent nematic hydrogels and
[74] B. Jiang, J. Zhang, W. Zheng, L.-J. Chen, G.-Q. Yin, Y.-X. Wang, B. Sun, X. Li, H.-B. chromonic mesophases driven by intra- and intermolecular interactions of
Yang, Construction of alkynylplatinum(II) bzimpy-functionalized bridged dinuclear cyclometalated platinum(II) complexes, Chem. Sci. 5
metallacycles and their hierarchical self-assembly behavior in solution (2014) 2482–2488.
promoted by PtPt and p–p interactions, Chem. Eur. J. 22 (2016) 14664– [98] M. Atoji, J.W. Richardson, R.E. Rundle, On the crystal structures of the Magnus
14671. salts, Pt(NH3)4PtCl4, J. Am. Chem. Soc. 79 (1957) 3017–3020.
[75] B. Jiang, J. Zhang, J.-Q. Ma, W. Zhang, L.-J. Chen, B. Sun, C. Li, B. Hu, H. Tan, X. Li, [99] W. Lu, S.S.-Y. Chui, K.-M. Ng, C.-M. Che, A Submicrometer wire-to-wheel
H.-B. Yang, Vapochromic behavior of a chair-shaped supramolecular metamorphism of hybrid tridentate cyclometalated platinum(II) complexes,
metallacycle with ultra-stability, J. Am. Chem. Soc. 138 (2016) 738–741. Angew. Chem. Int. Ed. 47 (2008) 4568–4572.
[76] Z.-Y. Li, L. Xu, C.-H. Wang, X.-L. Zhao, H.-B. Yang, Novel platinum-acetylide [100] W. Lu, K.-M. Ng, C.-M. Che, From supramolecular organoplatinum(II)
metaalocycles constructed via a stepwise fragment coupling approach and nanowires to platinum-containing nanocomposites, Chem. Asian J. 4 (2009)
their aggregation behavior, Chem. Commun. 49 (2013) 6194–6196. 830–834.
[77] N.-W. Wu, J. Zhang, X.-D. Xu, H.-B. Yang, Design and preparation of ethynyl- [101] Z. Gao, Z. Li, Z. Gao, F. Wang, Supramolecular alternate donor–acceptor
pyrene modified platinum-acetylide gelators and their application in copolymers mediated by Pt-Pt metal–metal interactions and their
dispersion of graphene, Chem. Commun. 50 (2014) 10269–10272. photocatalytic applications, Nanoscale 10 (2018) 14005–14011.
[78] L. Xu, H.-B. Yang, Our expedition in linear neutral platinum-acetylide [102] Z. Gao, P.A. Korevaar, R. Zhong, Z. Wu, F. Wang, Two-component
complexes: The preparation of micro/nanostructure materials, complicated supramolecular metallogels with the presence of Pt–Pt metal–metal
topologies, and dye-sensitized solar cells, Chem. Rec. 16 (2016) 1274–1297. interactions, Chem. Commun. 54 (2018) 9857–9860.
[79] J. Zhang, X.-D. Xu, L.-J. Chen, Q. Luo, N.-W. Wu, D.-X. Wang, X.-L. Zhao, H.-B. [103] V.C.-H. Wong, C. Po, S.Y.-L. Leung, A.K.-W. Chan, S. Yang, B. Zhu, X. Cui, V.W.-
Yang, Platinum acetylide complexes containing iptycene as cores: a new W. Yam, Formation of 1D infinite chains directed by metal-metal and/or p-p
family of unexpected efficient organometallic gelators, Organometallics 30 stacking interactions of water-soluble platinum(II) 2,6-bis(benzimidazol-2’-
(2011) 4032–4038. yl)pyridine double complex salts, J. Am. Chem. Soc. 140 (2018) 657–666.
[80] Y. Zhang, Q.-F. Zhou, G.-F. Huo, G.-Q. Yin, X.-L. Zhao, B. Jiang, H. Tan, X. Li, H.- [104] Z. Li, Y. Han, F. Wang, Compartmentalization-induced phosphorescent
B. Yang, Hierarchical self-assembly of an alkynylplatinum(II) bzimpy emission enhancement and triplet energy transfer in aqueous medium, Nat.
functionalized metallacage via PtPt and p–p interactions, Inorg. Chem. 57 Commun. 10 (2019) 3735–3742.
(2018) 3516–3520. [105] C. Yu, K.M.-C. Wong, K.H.-Y. Chan, V.W.-W. Yam, Polymer-induced self-
[81] C. Po, A.Y.-Y. Tam, K.M.-C. Wong, V.W.-W. Yam, Supramolecular self- assembly of alkynylplatinum(II) terpyridine complexes by metal-metal/p-p
assembly of amphiphilic anionic platinum(II) complexes: a correlation interactions, Angew. Chem. Int. Ed. 44 (2005) 791–794.
between spectroscopic and morphological properties, J. Am. Chem. Soc. 133 [106] K. Zhang, M.C.-L. Yeung, S.Y.-L. Leung, V.W.-W. Yam, Living supramolecular
(2011) 12136–12143. polymerization achieved by collaborative assembly of platinum(II)
[82] H.-L. Au-Yeung, S.Y.-L. Leung, A.Y.-Y. Tam, V.W.-W. Yam, transformable complexes and block copolymers, Proc. Natl. Acad. Sci. 114 (2017) 11844–
nanostructures of platinum-containing organosilane hybrids: non-covalent 11849.
self-assembly of polyhedral oligomeric silsesquioxanes assisted by Pt-Pt and [107] K. Zhang, M.C.-L. Yeung, S.Y.-L. Leung, V.W.-W. Yam, Manipulation of
p-p stacking interactions of alkynylplatinum(II) terpyridine moieties, J. Am. nanostructures in the co-assembly of platinum(II) complexes and block
Chem. Soc. 136 (2014) 17910–17913. copolymers, Chem 2 (2017) 825–839.
[83] L. Hu, F. Qu, Y. Wang, J. Shen, Q. He, B. Zhang, W. Bu, Phosphorescent and [108] K. Zhang, M.C.-L. Yeung, S.Y.-L. Leung, V.W.-W. Yam, Energy landscape in
semiconductive fiber-like micelles formed by platinum(II) complexes and supramolecular coassembly of platinum(II) complexes and polymers:
block copolymers, J. Mater. Chem. C 5 (2017) 12500–12506. morphological diversity, transformation, and dilution stability of
[84] N. Liu, B. Wang, W. Liu, W. Bu, Luminescent polymeric hybrids formed by nanostructures, J. Am. Chem. Soc. 140 (2018) 9594–9605.
platinum(II) complexes and block copolymers, Chem. Commun. 47 (2011) [109] C. Yong, K.H.-Y. Chan, K.M.-C. Wong, V.W.-W. Yam, Single-stranded nucleic
9336–9338. acid-induced helical self-assembly of alkynylplatinum(II) terpyridine
[85] N. Liu, B. Wang, W. Liu, W. Bu, Reversible luminescent switching complexes, Proc. Natl. Acad. Sci. 103 (2006) 19652–19657.
accompanied by assembly-disassembly of metallosupramolecular [110] S. Sinn, L. Yang, F. Biedermann, D. Wang, C. Kübel, J.J.L.M. Cornelissen, L. De
amphiphiles based on a platinum(II) complex, J. Mater. Chem. C 1 (2013) Cola, Templated formation of luminescent virus-like particles by tailor-made
1130–1136. Pt(II) amphiphiles, J. Am. Chem. Soc. 140 (2018) 2355–2362.
[86] N. Liu, Q. He, Y. Wang, W. Bu, Stepwise self-assembly of a block copolymer- [111] F.C.-M. Leung, S.Y.-L. Leung, C.Y.-S. Chung, V.W.-W. Yam, Metal-metal and p-
platinum(II) complex hybrid in solvents of variable quality: from wormlike p interactions directed end-to-end assembly of gold nanorods, J. Am. Chem.
micelle to free standing sheet to vesiclelike nanostructure, Soft. Mater. 13 Soc. 138 (2016) 2989–2992.
(2017) 4791–4798. [112] Y.-S. Wong, F.C.-M. Leung, M. Ng, H.-K. Cheng, V.W.-W. Yam, Platinum(II)-
[87] N. Liu, Y. Wang, C. Wang, Q. He, W. Bu, Syntheses and controllable self- based supramolecular scaffold-templated side-by-side assembly of gold
assembly of luminescence platinum(II) plane-coil diblock copolymers, nanorods through Pt-Pt and p-p interactions, Angew. Chem. Int. Ed. 57
Macromolecules 50 (2017) 2825–2837. (2018) 15797–15801.
Y. Han et al. / Coordination Chemistry Reviews 414 (2020) 213300 21

[113] C.Y.-S. Chuang, V.W.-W. Yam, Induced self-assembly and Förster resonance [122] S. Carrara, A. Aliprandi, C.F. Hogan, L. De Cola, Aggregation-induced
energy transfer studies of alkynylplatinum(II) terpyridine complex through electrochemiluminescence of platinum(II) complexes, J. Am. Chem. Soc. 139
interaction with water-soluble ploy(phenylene ethynylene sulfonate) and (2017) 14605–14610.
the proof-of-principle demonstration of this two-component ensemble for [123] J.R. Brandt, F. Salerno, M.J. Fuchter, The added value of small-molecule
selective label-free detection of human serum albumin, J. Am. Chem. Soc. 133 chirality in technological applications, Nature Rev. Chem. 1 (2017) 0045.
(2011) 18775–18784. [124] X.-P. Zhang, V.Y. Chang, J. Liu, X.-L. Yang, W. Huang, Y. Li, C.-H. Li, G. Muller,
[114] C.Y.-S. Chung, K.H.-Y. Chan, V.W.-W. Yam, ‘‘Proof-of-principle” concept for X.-Z. You, Potential switchable circularly polarized luminescence from chiral
label-free detection of glucose and a-glucosidase activity through the cyclometaled platinum(II) complexes, Inorg. Chem. 54 (2015) 143–152.
electrostatic assembly of alkynylplatinum(II) terpyridine complexes, Chem. [125] T. Ikeda, K. Hirano, T. Haino, A circularly polarized luminescent organogel
Commun. 47 (2011) 2000–2002. based on a Pt(II) complex possessing phenylisoxazoles, Mater. Chem. Front. 2
[115] M.C.-L. Yeung, K.M.-C. Wong, Y.K.T. Tsang, V.W.-W. Yam, Aptamer-induced (2018) 468–474.
self-assembly of a NIR emissive platinum(II) terpyridine complex for label- [126] T. Ikeda, M. Takayama, J. Kumar, T. Kawai, T. Haino, Novel helical assembly of
and immobilization-free detection of lysozyme and thrombin, Chem. a Pt(II) phenylbipyridine complex directed by metal-metal interaction and
Commun. 46 (2010) 7709–7711. aggregation-induced circularly polarized emission, Dalton Trans. 44 (2015)
[116] C.Y.-S. Chung, V.W.-W. Yam, Selective label-free detection of G-quadruplex 13156–13162.
structure of human telomere by emission spectral changes in visible-and-NIR [127] C. Wang, H. Dong, W. Hu, Y. Liu, D. Zhu, Semiconducting p-conjugated
region under physiological condition through the FRET of a two-component systems in field-effect transistors: a materials odyssey of organic electronics,
PPE-SO3–Pt(II) complex ensemble with Pt-Pt, electrostatic and p-p Chem. Rev. 112 (2012) 2208–2267.
interactions, Chem. Sci. 4 (2013) 377–387. [128] Y. Chen, K. Li, W. Lu, S.S.-Y. Chui, C.-W. Ma, C.-M. Che, Photoresponsive
[117] C.W.-T. Chan, H.-K. Cheng, F.K.-W. Hau, A.K.-W. Chan, V.W.-W. Yam, supramolecular organometallic nanosheets induced by PtII-PtII and CH-p
Protamine-induced supramolecular self-assembly of red-emissive interactions, Angew. Chem. Int. Ed. 48 (2009) 9909–9913.
alkynylplatinum(II) 2,6-bis(benzimidazol-2’-yl)pyridine complex for [129] C.-M. Che, C.-F. Chow, M.-Y. Yuen, V.A.L. Roy, W. Lu, Y. Chen, S.S.-Y. Chui, N.
selective label-free sensing of heparin and real-time monitoring of trypsin Zhu, Single microcrystals of organoplatinum(II) complexes with high charge-
activity, ACS Appl. Mater. Interfaces. 11 (2019) 31585–31593. carrrier mobility, Chem. Sci. 2 (2011) 216–220.
[118] M.C.-L. Chan, V.W.-W. Yam, NIR-emissive alkynylplatinum(II) terpyridine [130] R.W.-Y. Sun, A.L.-F. Chow, X.-H. Li, J.J. Yan, S.S.-Y. Chui, C.-M. Che, Luminescent
complex as a turn-on selective probe for heparin quantification by cyclometaled platinum(II) complexes containing N-heterocyclic carbene
induced helical self-assembly behavior, Chem. Eur. J. 17 (2011) 11987– ligands with potent in vitro and in vivo anti-cancer properties accumulate
11990. in cytoplasmic structures of cancer cells, Chem. Sci. 2 (2011) 728–736.
[119] T. Tu, W. Fang, X. Bao, X. Li, K.H. Dötz, Visual chiral recognition through [131] J.L.-L. Tsai, T. Zou, J. Liu, T. Chen, A.O.-Y. Chan, C. Yang, C.-N. Lok, C.-M. Che,
enantioselective metallogel collapsing: synthesis, characterization, and Luminescent platinum(II) complexes with self-assembly and anti-cancer
application of platinum-steroid low-molecular-mass gelators, Angew. properties: hydrogel, pH dependent emission color and sustained-release
Chem. 123 (2011) 6731–6735. properties under physiological conditions, Chem. Sci. 6 (2015) 3823–3830.
[120] F.J.M. Hoeben, P. Jonkheijm, E.W. Meijer, A.P.H.J. Schenning, About [132] Z. Chen, A.K.-W. Chan, V.C.-H. Wong, V.W.-W. Yam, A supramolecular
supramolecular assemblies of p-conjugated systems, Chem. Rev. 105 strategy toward an efficient and selective capture of platinum(II)
(2005) 1491–1546. complexes, J. Am. Chem. Soc. 141 (2019) 11204–11211.
[121] C.A. Strassert, C.-H. Chien, M.D.G. Lopez, D. Kourkoulos, D. Hertel, K. [133] C.-Y. Sun, W.-P. To, F.-F. Hung, X.-L. Wang, Z.-M. Su, C.-M. Che, Metal-organic
Meergolz, L. De Cola, Switching on luminescence by the self-assembly of a framework composites with luminescent pincer platinum(II) complexes:
3
platinum(II) complex into gelating nanofibers and electroluminescent films, MMLCT emission and photoinduced dehydrogenation catalysis, Chem. Sci. 9
Angew. Chem. Int. Ed. 50 (2011) 946–950. (2018) 2357–2364.

You might also like