Shunichiro Ito
ab and
Kazuo Tanaka
*ab
aDepartment of Polymer Chemistry, Graduate School of Engineering, Kyoto University, Katsura, Nishikyo-ku, Kyoto 615-8510, Japan. E-mail: tanaka@poly.synchem.kyoto-u.ac.jp
bDepartment of Technology and Ecology, Graduate School of Global Environmental Studies, Kyoto University, Katsura, Nishikyo-ku, Kyoto 615-8510, Japan
First published on 17th April 2025
Boron, aluminum, gallium, indium, and thallium are group 13 elements that can induce various electronic properties and unique functions when incorporated into main-chain conjugation through polymers. As vacant p-orbitals in these elements interact with Lewis bases, stimuli responsiveness can be induced. Additionally, the chemical and thermal stability can be enhanced by connecting with extra Lewis bases as supporting ligands. Moreover, superior optoelectronic properties, such as light absorption, emission, and carrier mobility, are often observed from group 13 element-containing π-conjugated systems. The introduction of boron into conjugated systems has been widely applied not only for improving material properties but also for providing new functionalities for conventional polymers. In contrast, there are limited examples of polymers possessing the heavier group 13 elements in their repeating units. According to recent studies, it has been shown that the chemical, physical, and material properties of π-conjugated compounds can be unexpectedly modulated by these heavier group 13 elements. In this review, we mainly explain the synthesis and fundamental photophysical properties of conjugated polymers consisting of the heavier group 13 elements in their main-chains.
B | Al | Ga | In | Tl | |
---|---|---|---|---|---|
a Data taken from ref. 2. | |||||
Ground-state electron configuration | [He]2s22p1 | [Ne]3s23p1 | [Ar]4s24p1 | [Kr]5s25p1 | [Xe]6s26p1 |
Ionization energy/kJ mol−1 | |||||
E → E+ | 800.637 | 577.539 | 578.844 | 558.299 | 589.351 |
E+ → E2+ | 2427.07 | 1816.68 | 1979.41 | 1820.71 | 1971.03 |
E2+ → E3+ | 3659.75 | 2744.78 | 2963 | 2704 | 2878 |
E3+ → E4+ | 25![]() |
11![]() |
6175 | 5210 | (4900) |
Owing to their wide versatility—including semiconductivity, light absorption and emission, non-linear optical properties, and Lewis acidity,—much effort has been devoted to studying the electronic structures of π-conjugated polymers involving boron. By modulating structures, various advantageous chemical and physical properties derived from their polymeric structures, such as excellent processability, elasticity, high local concentrations, environmental sensitivity, and extension of π-conjugation, can be obtained. Thus far, many reviews describing such boron-containing polymeric materials have been published.4–15 In particular, the incorporation of boron-coordinating structures into π-conjugated systems has been revisited because they allowed access to n-type semiconducting molecules and polymers.16 However, the number of polymers containing the heavier congeners of B, Al, Ga, In, and Tl remains limited.17,18 The heavier elements often play a critical role in drastically improving small-molecular material properties, which originate from the chemical structure of a molecule, intermolecular interactions, molecular motions, and other unpredictable factors.
As a representative example, the efficiency of preliminary organic light-emitting diodes (OLEDs) consisting of 8-hydroxylquinolinate complexes is usually higher with aluminum complexes as compared to those with boron complexes.19 Moreover, a device composed of gallium tris(8-hydroxylquinoline) showed higher efficiency as compared to those of aluminum and indium complexes probably because of the superior charge-transporting property of the gallium complex.20 Such striking differences in chemical and physical properties among group 13 elements result from differences in atomic and ionic size, orbital energies, nature of bonds, interactions with neighboring atoms, and intermolecular interactions.
These observations have encouraged the development of a new series of materials composed of these elements and comparison of their chemical and physical properties. Notably, the elemental dependency of optoelectronic properties, especially in the solid state, cannot generally be inferred from boron chemistry. Therefore, it is necessary to discover the appropriate elements that maximize the material properties for each practical application. Furthermore, the adoption of heavier homologues into polymer structures may accentuate the differences between boron and other elements and lead to advanced functionalities, such as room-temperature phosphorescence, stimuli-responsive luminescence, and higher electrical conductivity, as demonstrated in small molecule systems.
There has been a focus on the synthesis of inorganic and organometallic molecular compounds composed of the heavier group 13 elements to clarify their molecular structures, chemical reactivity, and consistency between experimental and theoretical data.21–23 However, the incorporation of those species into π-conjugated polymer backbones has been prevented because it is likely that most compounds consisting of the heavier group 13 elements decompose under common polymerization conditions, such as metal-catalyzed cross-coupling reactions. Thus, alternative synthetic approaches are required to construct π-conjugated polymers composed of these classes of complexes. One successful approach is the post-polymerization functionalization method, where element-containing structures are constructed after polymerization reactions. This method facilitates the development of polymers that incorporate relatively unstable moieties and allows easy control of the functionalization ratio by simply adjusting the reaction stoichiometry.24 Another approach is the kinetic and thermodynamic stabilization of monomers by employing elaborate supporting ligands, which enables us to assess the material properties of these classes of compounds.25,26
This review mainly focuses on recent synthetic advances that can be used to obtain fully characterized π-conjugated polymers and oligomers involving aluminum and gallium. In addition, an indium-containing homocatenate oligomer is also described for stimulating further development in this field. The chemical and physical properties of these materials will be discussed.
![]() | ||
Fig. 1 (a) Chemical structure of Alq3. Conjugated polymers composed of Alq3-type complexes with (b) poly(8-quinolinol) and (c) poly(aryleneethynylene) scaffolds. |
λabs/nm | λPL/nm | PLQY | Condition | Ref. | |
---|---|---|---|---|---|
Alq3 | 384 | 521 | 0.13 | Chloroform solution | 45 |
Alq′2q | 375 | 507 | 0.07 | Chloroform solution | 45 |
1 | 359 | 489 | 0.23 | Chloroform solution | 54 |
564 | Film | ||||
2pre-q(6/4-OEH) | 423 | 472 | 0.34 | Chloroform solution | 55 |
429 | 491, 536 | Film | |||
2pre-r(6/4-OEH) | 413 | 471 | 0.49 | Chloroform solution | |
417 | 486, 522 | Film | |||
2pre-q(8/2-OEH) | 426 | 475 | 0.32 | Chloroform solution | |
428 | 493, 550 | Film | |||
2pre-r(8/2-OEH) | 427 | 476 | 0.34 | Chloroform solution | |
437 | 490, 526 | Film | |||
2-q(6/4-OEH) | 419 | 472 | 0.22 | Chloroform solution | |
425 | 473, 507 | Film | |||
2-r(6/4-OEH) | 413 | 471 | 0.54 | Chloroform solution | |
445 | 474, 504 | Film | |||
2-q(8/2-OEH) | 435 | 474 | 0.25 | Chloroform solution | |
439 | 570 | Film | |||
2-r(8/2-OEH) | 433 | 476 | 0.37 | Chloroform solution | |
442 | 493, 529 | Film |
Aluminum tris(8-quinolinolate), Alq3 (Fig. 1a), and its derivatives are the most important family of aluminum complexes in the field of optoelectronics because Alq3 possesses a superb ability for electron transport accompanied by satisfactory luminescence in solid states. As Alq3 was applied as the emitting and electron-transporting layer in the first organic light-emitting diode,32 tremendous research efforts have been devoted to developing functionalized Alq3-based materials for improved luminescent and electronic properties, processability, and chemical stability. In particular, immobilization by covalent bonds within polymer backbones enhances film formability and device stability. Some examples of Alq3-decorated polymers with non-conjugated backbones are poly(bisphenol-A sulfone),42 polynorbornene,43,44 phenol resin,45 poly(methyl methacrylate),46–52 and polystyrene.53
These studies successfully demonstrated that polymeric complexes can be applied both to solution-processed device fabrication and in preparing luminescent micelles.49 Most polymers in these studies were prepared with post-polymerization complexation to prevent the chemical decomposition of complex moieties. It was also reported that Alq3 is labile and that a ligand exchange reaction often takes place to generate cross-linking, which will result in insoluble polymers, followed by the generation of low-molecular weight polymers with low yields.42 The solubility of polymers can be improved by introducing alkyl chains46 and by directly polymerizing Alq3-decorated monomers using methods, such as ring-opening metathesis43 and radical polymerizations.47
Incorporating Alq3 into the main-chain of π-conjugated polymers has also been attempted for modulating and improving its optoelectronic properties. Post-polymerization complexation of 8-quinolinol-based conjugated polymers was applied to obtain fully characterized Alq3-based polymers in pioneering works.54,55 A soluble poly(8-quinolinol), 1pre, was synthesized by enzyme-catalyzed polymerization and subjected to subsequent complexation with bis(2-methyl-8-quinolinolato)ethylaluminum (Alq′2(Et)) to give 1 (Fig. 1b).54 Horseradish peroxidase (HRP) was used for the polymerization of 8-quinolinol in the presence of hydrogen peroxide as an oxidizing reagent. The number- and weight-average molecular weights (Mn and Mw) were determined to be 2500 and 3400, respectively, by GPC analysis. Importantly, the 1H and 13C NMR spectra of 1pre showed that the polymerization predominantly proceeded regioselectively at the 4 and 7 positions of the 8-quinolinol ring. The subsequent reaction of 1pre with Alq′2(Et) resulted in the complete conversion of the quinolinol repeating units into the aluminum complex of 1. It was reported that 1 exhibits photoluminescence in solution and in the solid state, while 1pre shows little emission. Interestingly, the relative photoluminescence quantum yield (PLQY) of 1 (0.27) is higher than that of Alq′2q (0.07). The longest-wavelength absorption maximum (λabs) of Alq3 and 1 was determined to be 375 and 359 nm, respectively. The hypsochromic shift of λabs might originate from the electron-withdrawing property of the poly(8-quinolinol-4,7-diyl) backbone.
As metal-catalyzed cross-coupling reactions offer more reliable control of regioselectivity than oxidative coupling polymerizations, the palladium-catalyzed Sonogashira–Hagihara reaction was used for preparing poly(aryleneethynylene)-type conjugated polymers consisting of 8-quinolinol units in their main-chain (Fig. 1c).55 The polymerization reactions were carried out using three types of silyl-protected 8-quinolinol monomers with different substitution positions (q–r) and diethynyl- and dibromo–dialkoxyphenylene co-monomers. The subsequent deprotection of the silyl groups using tetrabutylammonium fluoride (TBAF) afforded the precursor polymers 2pre. The Mn values of the silyl-protected polymers ranged from 8500 to 13000, which corresponds to approximately eight monomeric units of monomer q in the polymer chain for the case of 2pre-q(6/4-OEH).
Although treatment of 2pre with Alq′2(Et) was attempted to give the corresponding Alq3-incorporated conjugated polymers, the dodecyloxy-based polymers (the –OC12 series) yielded insoluble products probably due to cross-linking caused by a ligand exchange reaction of the Alq3-type scaffold. However, 2-ethylhexyloxy-based polymers resulted in desired soluble polymers 2 because the branched alkoxy groups retarded the intermolecular ligand exchange. FT-IR spectroscopy and elemental analysis indicated that the conversion of the phenolic hydroxyl group to the aluminum complex was approximately 30% in the case of 2-q(6/4-OEH), suggesting that its single polymer chain could statistically contain approximately two aluminum complex units. Photoluminescence was emitted by both 2pre and 2 in solution and in solid-state films. The photoluminescence spectra of 2pre-q(6/4-OEH) and 2-q(6/4-OEH) were almost the same shape and peaked at 472 nm, which was assignable to the aromatic polymer chain rather than the aluminum complex moiety. In the solid-state thin films, there were additional emission bands attributed to an excimer-like adduct for 2pre-q(6/4-OEH) (536 nm) and an aluminum complex unit for 2-q(6/4-OEH) (507 nm). Hence, it was implied that the intermolecular energy transfer from the aryleneethynylene main-chains to the aluminum complexes could occur in the film of 2-q(6/4-OEH).
Aluminum porphyrins and salen complexes are one of the most widely investigated families of aluminum-based π-conjugated molecules because of their promising catalytic activities. Taking advantage of their microporous structures, it has been revealed in recent years that conjugated microporous polymers (CMPs) consisting of aluminum complexes exhibit outstanding catalytic properties.56–58 Representative aluminum complex-based CMPs 3,56 4,57 and 558 were synthesized by post-polymerization complexation, one-pot polymerization and complexation, and direct polymerization, respectively (Fig. 2).
High-resolution transmission electron microscopy images and gas adsorption analysis of CMP 3 revealed nanometer-sized pores (approximately 0.5 nm) and 798 m2 g−1 of BET surface area. Its CO2 uptake at 298 K and 760 mmHg reached 76.5 mg g−1, which is comparable with that of representative metal–organic frameworks. As a result, cobalt porphyrin CMP 3 exhibited exceptionally high catalytic activity in converting propylene oxide (PO) and CO2 to propylene carbonate (PC) in the presence of tetrabutylammonium bromide as a co-catalyst, achieving 78.2% conversion at atmospheric pressure and room temperature.
Similar highly efficient catalytic activity was also reported for CMP 4 with turnover frequencies (TOF) of up to 443.6 h−1 at room temperature and atmospheric pressure. Furthermore, due to the cationic units in CMP 5, co-catalyst was not required, and remarkably high TOF of up to 2200 h−1 resulted for the conversion of epichlorohydrin and CO2 to the corresponding cyclic carbonate. Further synthetic efforts toward group 13 complex-based polymers involving gallium59 could open a novel avenue for achieving highly efficient and chemically stable catalytic systems.
![]() | ||
Fig. 3 Conjugated polymers composed of Mamx-stabilized (a) gallafluorene and (b) gallium-bridged phenylene. |
λabs/nm | λPL/nm | PLQY | Condition | Ref. | |
---|---|---|---|---|---|
6-C | 326 | 383 | 0.37 | Chloroform solution | 73 |
326 | 385 | 0.05 | Film | ||
6-O | 366 | 411 | 0.34 | Chloroform solution | |
365 | 420 | 0.04 | Film | ||
6-Flu | 383 | 416, 438 | 0.69 | Chloroform solution | |
383 | 430, 454 | 0.06 | Film | ||
6-BTH | 404 | 486 | 0.20 | Chloroform solution | |
407 | 515 | 0.07 | Film | ||
6-BTz | 442 | 479 | 0.53 | Chloroform solution | |
441, 469 | 479, 531 | 0.09 | Film | ||
6-CDT | 481 | 531, 569 | 0.32 | Chloroform solution | |
478 | 572 | 0.004 | Film | ||
6-BTA | 374, 477 | 643 | 0.43 | Chloroform solution | |
377, 485 | 641 | 0.08 | Film | ||
7-Ph | 275 | 378 | 0.32 | Chloroform solution | 74 |
7-PhO | 277, 325 | 392 | 0.23 | ||
Biphenyl | 250 | 314 | <0.01 | ||
7-Model-terphenyl | 266, 318 | 384 | 0.16 | ||
7-Model1 | 263 | 319 | 0.01 | ||
7-Model2 | 269 | 322 | 0.04 | ||
8-B-FL | 399 | 545 | <0.01 | Chloroform solution | 98 |
399 | 545 | 0.07 | Film | ||
8-B-CBZ | 397 | 545 | <0.01 | Chloroform solution | |
404 | 552 | 0.07 | Film | ||
8-B-BT | 404 | 581 | <0.01 | Chloroform solution | |
417 | 575 | 0.07 | Film | ||
8-Ga-FL | 411 | 571 | <0.01 | Chloroform solution | |
415 | 575 | 0.05 | Film | ||
8-Ga-CBZ | 410 | 576 | <0.01 | Chloroform solution | |
414 | 573 | 0.05 | Film | ||
8-Ga-BT | 420 | 610 | <0.01 | Chloroform solution | |
424 | 601 | 0.03 | Film |
The 2,4-di-t-butyl-6-(N,N-dimethylaminomethyl)phenyl group, known as the Mamx ligand (the structure is shown in Fig. 3a),65 thermodynamically and kinetically stabilizes group 13 organoelement compounds, leading to the isolation of gallium-containing polymers such as poly(galla[1]ferrocenophane).66 Thus, the Mamx ligand has opened an avenue for the synthesis and evaluation of π-conjugated compounds consisting of group 13 elements, such as 9-heterofluorenes,67–70 dithienoheteroles,71 and dibenzoheteropins.72 These compounds are stable under atmospheric conditions and can be purified with typical silica-gel column chromatography. It is important to note that the gallafluorene and indafluorene derivatives exhibited fluorescence and phosphorescence in room-temperature solutions, while the borafluorene and alumafluorene derivatives showed only fluorescence. Furthermore, the addition of tris(pentafluorophenyl)borane (B(C6F5)3) to solutions of bora-, aluma-, and gallafluorenes induced phosphorescence from their triplet exciplex of the heterofluorenes and B(C6F5)3.67,68
Mamx-stabilized gallafluorene is sufficiently stable to be directly subjected to the palladium-catalyzed Suzuki–Miyaura, Sonogashira–Hagihara, and Migita–Kosugi–Stille coupling reactions to give a variety of conjugated polymers 6 (Fig. 3a).73 Their absorption and photoluminescence spectra highly relied on the co-monomers, and their emission colors covered the visible region. Notably, the fluorescence lifetime measurements indicated that all their emission bands were assigned to fluorescence, based on their nanosecond-order lifetimes. No apparent phosphorescence was observed even at 77 K. Compared to the spectra in solution states, the redshifted emission spectra of 6-O, 6-Flu, 6-BTH, and 6-BTz suggested intermolecular interactions.
In particular, the film of 6-BTz exhibited two distinct bands in the absorption (441 and 469 nm) and emission (479 and 531 nm) spectra. The lower-energy bands were absent in its dilute solution. The intensity ratio of these two peaks depended on the solution concentration and film-preparation method. Therefore, it was suggested that the intermolecular interactions in concentrated states resulted in those additional absorption and emission bands. Furthermore, the results of cyclic voltammetry showed that the gallafluorene moiety should function as a stronger electron donor than 9,9-dialkylfluorene because the synthesized gallafluorene polymers have higher highest occupied molecular orbital (HOMO) levels than the corresponding fluorene-based polymers.
The effects of gallium atoms on π-conjugation of polymers were evaluated by incorporating the Mamx-stabilized triarylgallium unit into the poly(p-phenylene) scaffolds (Fig. 3c).74 Homopolymerization with nickel-mediated Yamamoto coupling and copolymerization with palladium-catalyzed Suzuki–Miyaura cross-coupling successfully yielded 7-Ph and 7-PhO, respectively. Importantly, compared to the corresponding model compounds, the apparent red shifts of the absorption and emission spectra by polymerization were observed. The dimeric model, 7-model2, exhibited absorption and emission bands between 7-Ph and 7-model1. These observations strongly indicate that main-chain π-conjugation should extend through the gallium atoms. In addition, using theoretical calculations employing periodic boundary conditions for 7-Ph, the bandwidths of the highest occupied and lowest unoccupied crystal orbitals (HOCOs and LUCOs) were calculated to be 0.112 and 0.347 eV, respectively. These values certainly indicate that the electronic conjugation in the LUCO apparently extends over the main-chain involving the gallium atoms.
In the context of material science, a major limitation of typical organic luminophores is that their planar π-conjugated structures, which efficiently emit in dilute solutions but experience luminescence loss in aggregated or crystalline states due to concentration quenching. On the contrary, AIE- and CIE-active materials exhibit efficient luminescence in aggregate and crystalline states, in particular, respectively. Their solutions show only lower efficiency emission.75–78 Therefore, those classes of luminophores are utilized in a wide range of practical applications, such as organic light-emitting diodes, bioimaging,79 and chemosensors.80
Recently, β-diketiminate ligands, which have been utilized for isolating various reactive species,81,82 have enabled us to develop luminescent complexes and demonstrate their aggregation- and crystallization-induced emission (AIE and CIE) properties.36,39,83–92 Furthermore, the incorporation of these complexes into π-conjugated polymer backbones results in emissive film materials with productive responses to external stimuli.84,90,93–97 The mechanisms behind the AIE and CIE properties of β-diketiminate complexes have been proposed: the sterically hindered substituents, e.g., aromatic rings, in the ligands hampered detrimental intermolecular interactions, which caused concentration quenching in concentrated states. The quite weak luminescence in dilute solutions could stem from rapid nonradiative quenching processes through vibronic couplings88 and conical intersections.90
In this context, the post-polymerization complexation method was applied to achieve π-conjugated polymers containing gallium β-diketiminate complexes (Fig. 4).98 The ligand polymers 8pre-FL, 8pre-CBZ, and 8pre-BT were synthesized using the palladium-catalyzed Suzuki–Miyaura coupling reaction. The subsequent treatment of these polymeric ligands with trichlorogallium in the presence of triethylamine afforded the corresponding gallium polymers 8-Ga-FL, 8-Ga-CBZ, and 8-Ga-BT. NMR, FT-IR, and XRF analyses revealed that complexation reactions were completely accomplished at all the β-diketiminate coordination sites. The Suzuki–Miyaura coupling polycondensation employing the diiodo-β-diketiminate boron complex directly yielded the corresponding boron polymers 8-B-FL, 8-B-CBZ, and 8-B-BT. Notably, the direct Suzuki–Miyaura polymerizations of the corresponding gallium-containing monomer likely failed because of a more labile Ga–Cl bond as compared to B–F. Therefore, the post-polymerization method was required for achieving gallium-containing polymers. The enhancement of the PLQY of the film state compared to the solution state clearly indicates their AIE properties derived from the β-diketiminate scaffold. Remarkably, the gallium-based polymers exhibited bathochromic shifts in the absorption and emission spectra compared to the corresponding boron-based polymers.
Cyclic voltammetry measurements and theoretical calculations suggested that the LUMO level was significantly stabilized by replacing the central element boron with gallium, while the HOMO level was not drastically affected. As a result, the HOMO–LUMO gap of the gallium polymers should be lower than that of the boron polymers, leading to the observed bathochromic shifts. Notably, the HOMO–LUMO gap of the small molecules of the gallium and boron complexes is comparable, although the HOMO and LUMO of the gallium complexes are lower than those of the boron complexes. Consequently, the effects of the central atom were likely manifested by incorporating the complexes into the π-conjugated polymers because the relatively electron-rich co-monomers should limit the HOMO level of the polymers.
Except for the series of polymerizations with metal catalysts, electrochemical polymerization has also been applicable for preparing homogenous thin films of conjugated polymers directly onto electrodes. As the obtained polymer films are usually insoluble in any solvent, electropolymerized films can be utilized for further applications, such as transparent electrodes, thin-film sensors, and precursors for preparing inorganic nanoparticles.99–101 As thiophene and its derivatives are easily polymerized on anodes, thiophene-terminated monomers are often applied for this approach.
In 2010, a typical example of electrochemical polymerization with a gallium-containing monomer was reported (Fig. 5).102 To fabricate conducting polymer films as a precursor of wide-bandgap semiconducting Ga2S3 nanoparticles, the monomer was designed. It was composed of a gallium Schiff-base complex decorated with bithiophene, allowing for anode polymerization. First, a five-coordinate neutral monomer possessing a chloride ligand on the gallium atom was used in the preparation. As expected, electrochemical polymerization successfully occurred to form polymeric precursor 9-neutral on various working electrodes, including carbon-coated gold TEM grids, indium tin oxide, stainless steel, and platinum buttons. For the preparation of Ga2S3 nanoparticles, the obtained thin films were repeatedly treated with the following cycle: dichloromethane solution of H2S, methanol solution of Ga(NO3)2, and then the dichloromethane solution of H2S again. This process was similar to that used for the preparation for cadmium sulfide nanoparticles.99 However, the desired Ga2S3 nanoparticles were not obtained due to the low reactivity of 5-coordinated gallium.
![]() | ||
Fig. 5 Electrochemical polymerization of gallium complexes of a Schiff base and subsequent preparation of gallium sulfide nanoparticles. |
The addition of KB(C6F5)4 induced the release of the apical chloride ligands, followed by the generation of cationic species, and the reactivity of gallium was enhanced. This cationic 2-coordinated monomer was also polymerized in a similar manner to give polymer 9-cation, which was further subjected to the nanoparticle preparation process. TEM observations supported the generation of Ga2S3 nanoparticles on polymer films. Moreover, the size of the nanoparticles was tunable and controllable by changing the number of preparation cycles. The average sizes of nanoparticles were 2.99 and 3.41 nm for two and four cycles, respectively.
β-Diketiminate ligands have been widely applied for separating such unstable species because of the facile tunability of their steric and electronic demands. Careful screening of the steric hindrance of β-diketiminate ligands for indium revealed that the 3,5-dimethylphenyl groups were suitable for obtaining a homocatenated hexamer, 10, in the solid state (Fig. 6). The formal oxidation state of the internal four indium atoms was +1, while the terminal two indium atoms were divalent because of the iodine substituents. Such low oxidation states of group 13 atoms should be more easily accessible for the heavier atoms, such as indium and thallium, because of the inert pair effect.
Indeed, a similar trimeric structure was observed in the crystal of the thallium β-diketiminate complex.105 These homocatenated polymers/oligomers can be applied as synthetic precursors of metal nanoparticles. The single-crystal X-ray analysis revealed that the In–In bond lengths increased as the distance from the iodine terminals increased, and they were within the range previously reported for In(II) complexes. In addition, the UV–Vis absorption spectrum of this complex in a hexane solution showed an absorption maximum at 349 nm, which could be attributed to the σ–σ* transition and implied the existence of such oligomeric structures even in solution.
As veiled structure–property relationships remain in the field of group 13 element-containing polymers that have not been experimentally accessible, new synthetic strategies must be developed. Moreover, the unique chemical and physical functionalities, such as large coordination numbers, low-valent species, and redox activity, must be exploited. For instance, it was theoretically predicted that polyborole could possess metal-like properties due to its quite small bandgap.106 If a heavier element is incorporated into the polyheterole backbone instead of boron, systematic change of the electronic properties ranging from metallic to semiconductor-like can occur. Although the chemical instability of group 13 heteroles prevents the isolation of polyborole and its heavier congeners, it might be achieved by insulation of polymer backbones.107 In addition, we witnessed the emergence and development of iminoborane (BN)-containing conjugated polymers, which show the extension of π-conjugation through the B
N bonds.108,109 The heavier congeners of B
N, involving Ga
P and In
P, might provide novel electronic states for conjugated polymers, such as weakly aromatic Ga2P and In2P rings,110 possibly resulting in higher hyperpolarizability.111 Furthermore, the inert pair effect in the heavier atoms could allow us to utilize high and low oxidation states, expanding the possibilities for redox applications, such as transition metal-like catalysts.112 In addition, the heavier congeners often improve the material properties involving charge carrier mobility,113 phosphorescence,114 and stimuli-responsiveness.115
Further improvements in the synthetic approach for isolation of many unstable structures and polymerization methods applicable to a wider range of monomers will expand the frontiers of the heavier group 13 elements by involving thallium chemistry, which is not able to be extrapolated from the chemistry of boron. In this context, the incorporation of these species into polymeric scaffolds could provide enhanced chemical stability because the hydrophobic polymeric structures could expel polar-reactive species, such as water. In addition, we witnessed that photochemical and electrochemical reactions and their cooperative usages have opened new horizons in the field of organic chemistry. The application of these powerful synthetic methodologies in inorganic and organometallic chemistries can play pivotal roles for accessing novel chemical structures.
This journal is © The Royal Society of Chemistry 2025 |