Next Article in Journal
Multinuclear MRI in Drug Discovery
Next Article in Special Issue
Au-TiO2/Ti Hybrid Coating as a Liquid and Gas Diffusion Layer with Improved Performance and Stability in Proton Exchange Membrane Water Electrolyzer
Previous Article in Journal
Synthesis, Docking Studies and Pharmacological Evaluation of Serotoninergic Ligands Containing a 5-Norbornene-2-Carboxamide Nucleus
Previous Article in Special Issue
Controlled Thermal Release of L-Menthol with Cellulose-Acetate-Fiber-Shelled Metal-Organic Framework
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Post-Functionalization of Organometallic Complexes via Click-Reaction

by
Stanislav Petrovskii
,
Viktoria Khistiaeva
,
Aleksandra Paderina
,
Evgenia Abramova
and
Elena Grachova
*
Institute of Chemistry, St. Petersburg University, 198504 St. Petersburg, Russia
*
Author to whom correspondence should be addressed.
Molecules 2022, 27(19), 6494; https://doi.org/10.3390/molecules27196494
Submission received: 30 August 2022 / Revised: 21 September 2022 / Accepted: 26 September 2022 / Published: 1 October 2022
(This article belongs to the Special Issue Feature Papers in Applied Chemistry)

Abstract

:
CuAAC (Cu catalyzed azide-alkyne cycloaddition) click-reaction is a simple and powerful method for the post-synthetic modification of organometallic complexes of transition metals. This approach allows the selective introduction of additional donor sites or functional groups to the periphery of the ligand environment. This is especially important if a metalloligand with free donor sites, which are of the same nature as the primary site for the coordination of the primary metal, has to be created. The concept of post-synthetic modification of organometallic complexes by click-reaction is relatively recent and the currently available experimental material does not yet allow us to identify trends and formulate recommendations to address specific problems. In the present study, we have applied the CuAAC reaction for the post-synthetic modification of diimine mononuclear complexes Re(I), Pt(II) and Ir(III) with C≡C bonds at the periphery of the ligand environment and demonstrated that click-chemistry is a powerful tool for the tunable chemical post-synthetic modification of coordination compounds.

Graphical Abstract

1. Introduction

Transition metal complexes are interesting objects from both fundamental and application points of view and their practically useful properties are determined by the combination of the metal center and ligand environment. To date, there are a large number of applications requiring the incorporation of a transition metal coordination compound into a complex molecular system by covalent conjugation without loss of practically useful properties. This problem can potentially be addressed by two synthetic pathways: (a) utilization of a ready-to-use donor site for metal coordination and (b) introducing a functional group on the periphery of the ligand environment for conjugation to the complementary substrate. The latter approach is a post-synthetic modification/functionalization of the complex and can be realized with high selectivity.
Triazoles are convenient scaffolds [1] that can have various substituents independent of nature (organic, inorganic, polymeric, nanoscale, biological, etc.). Two different fragments can be combined to form a triazole ring by the CuAAC (Cu catalyzed azide-alkyne cycloaddition) click-reaction [2,3], the best known and the most widely used form of the click processes. Despite the generally accepted success and undoubted potential in the construction of complex molecular ensembles, this approach is relatively rarely used in the field of coordination chemistry. However, the philosophy and methodology of click-chemistry are entering deeper into daily synthetic coordination chemistry, in particular, it is a convenient synthetic approach for creating new ligands [4,5]. A natural expansion of the idea of click-chemistry in the field of coordination chemistry is the creation of complex molecular objects by introducing click-groups (in particular, terminal C≡C bond) at the periphery of the ligand sphere for further click-conjugation (Scheme 1).
To date, a number of complexes containing C≡C bond on periphery of ligand environment to be involved in click-reaction have been reported in the literature for Mo(I), Tc(I, V) and Re(I, V) [6,7,8,9,10,11], Ir(III) [12,13,14,15], Pt(II) [16,17,18,19,20,21,22,23], Pt(IV) [24], Ru(II, III) [25,26,27,28,29,30,31,32,33,34], Fe(II) [35,36], ferrocene derivatives [37,38,39,40,41], Mn(I) [42], Ni(II) [43], Zn(II) porphyrins and analogues [44,45,46,47], Cr(0) and W(0) [48], Au(I) [49], Ln(III) [50,51,52,53,54,55,56], borane derivatives [57,58], polyoxometalates (POM) [59,60]. The click-chemistry approach allows the creation of multifunctional luminescent complexes [61], construction of conjugates for precision bio-imaging and diagnostics [62], attachment of vectors for transport in biological systems [63], creation of DNA-based materials, attachment of long hydrocarbon chains, attachment of electronic reservoirs for photocatalysis, construction of homo- and heterometallic bridging complexes, functionalization of polymers [64,65], and modification of MOFs [66] (representative examples are summarized in Table A1.
However, any evaluations of the metallocentre nature influence on the efficiency of the click-reaction have not received sufficient attention, although there is mention that post-modification can depend on the nature of the metal [67]. Thereby, the understanding of fundamental aspects of the click-chemistry approach for post-functionalization of transition metal complexes is still an open problem, and it is critically important for the rational molecular design of smart functional materials.
In the present work, we report a proof-of-principle solution for click-conjugation via CuAAC reaction of 2-azidomethylpyridine with the most popular luminescent diimine mononuclear complexes of Re(I), Pt(II) and Ir(III) bearing terminal C≡C bond on a periphery of the ligand environment and provide support that the click-chemistry is a powerful tool for the chemical post-synthetic modification of coordination compounds. The work presented herein is a contribution to the data bank to support the development of experimental approaches to the post-functional modification of transition metal complexes.

2. Results

2.1. Synthesis of Functionalized NN Compounds 1P−3P

The substituted bipyridines 1P and 2P were synthesized based on the slightly modified literature methods (Scheme 2) [68]. The other method based on previously published work was developed for the synthesis of compound 3P [69]. It should be noted that the yield of compound 3P is low and the ratio “cost of the synthesis/production of the target compound in pure form” is not justified for this potentially interesting compound. The main problem using the literature methods for the synthesis of NN compounds with C≡CH functional groups was that the products were contaminated with phosphine oxide and traces of Cu(II). Both of these impurities could affect the further production of the complexes, so a purification technique was specifically developed to remove these impurities completely. Since unprotected terminal acetylene is able to polymerize in the presence of transition metals, compounds XP (X = 1–3) with TMS protected acetylene units were directly used for the synthesis of transition metal complexes.

2.2. Synthesis and Post-Functionalization of Re(I) Complexes

The archetypal Re(I) heteroleptic neutral complexes ReXP (X = 1−3) decorated by protected NN ligand were obtained in the canonical reaction between Re(CO)5Cl and XP compound in refluxing toluene and in good yields (Scheme 3). Deprotection of the C≡C bond occurs in very mild conditions in methanol solution in the presence of K2CO3 at room temperature and takes 4 h to be completed resulting in complexes ReX (X = 1−3) formation. The following CuAAC reaction of post-functionalization of ReX complexes by 2-azidomethylpyridine also does not request any hard conditions and carries out in an acetone/water mixture in the presence of CuSO4 and NaAsc (sodium ascorbate) at room temperature overnight, resulting in ReXC (X = 1−3) formation with good yield. Complexes of ReXC are microcrystalline yellow solids that are moisture and air-stable, and readily soluble in common polar solvents such as chloroform, dichloromethane, acetone, and methanol.
The composition and structures of ReXP, ReX and ReXC (X = 1−3) were established by proton NMR experiments and ESI MS (in positive mode). On the other side, 1H NMR spectroscopy allowed us to control the transformation of the Re(I) complexes. The number of the signals in 1H NMR spectra, their relative integral intensities and multiplicities clearly indicate the presence of chelating and symmetrically substituted NN ligand in the coordination environment of Re(I) complexes (Figures S2–S4).
The complete disappearance of resonance that was assigned to protons of protective SiMe3 groups in ReXP spectra ca. 0.3 ppm and appearance of signal responsible for acetylene protons ca. 4 ppm in proton NMR spectra provide support for complete deprotection of C≡C bond in coordination environment of ReXP complexes and formation of ReX complexes. The successful click-reaction and ReXC complexes formation are verified by a set of resonances of triazole and pyridine rings in the aromatic part of proton NMR spectra.
ESI+ MS spectra of the Re(I) complexes display the signals of [M – Cl]+, [M + Na]+ and [M – Cl + N2]+ molecular ions with expected m/z values and isotopic patterns matching the calculated ones (Figures S5–S7).

2.3. Synthesis and Post-Functionalization of Ir(III) Complexes

The archetypal Ir(III) heteroleptic cationic complexes decorated by two cyclometallating ligands and ‘clickable’ diimine ligands were obtained in the canonical reaction between dichloro-bridged dimer complexes and XP compounds under mild conditions (Scheme 4). The cyclometallating ligands ppy and dfppy were chosen because of the difference in their donor-acceptor properties. The composition and structures of the Ir(III) complexes under study were established by proton NMR experiments (Figures S8–S11) and ESI+ MS (Figures S12–S15). Likewise, for all complexes studied herein, spectroscopy data provide support that the composition and structure of Ir(III) complexes are as expected.
It is important to note that, unlike all other complexes, the deprotection of Ir(III) complexes occurs by simple dissolution in methanol. The exception is Ir(dfppy)2P, for which deprotection unexpectedly does not complete under these conditions even for a long reaction time and in the presence of KF. Moreover, the addition of K2CO3 or KOH leads to the destruction of the complex. 1H NMR data for complex Ir(dfppy)2 were obtained by subtracting the spectrum of complex Ir(dfppy)2P from the spectrum of the reaction mixture (Figure S11).
The click-reaction of Ir(ppy)1, Ir(ppy)2 and Ir(dfppy)1 with 2-azidomethylpyridine takes a longer time than the one for Re(I) complexes described above and results in the formation of complexes Ir(ppy)1C, Ir(ppy)2C and Ir(dfppy)1C with relatively lower yields. There is a dependence of the reaction yield on the nature of the cyclometallating ligand, and this value is higher for the more acceptor C^N ligand (72% for Ir(dfppy)1C vs. 45% and 48% for Ir(ppy)1C and Ir(ppy)2C, respectively).

2.4. Synthesis and Post-functionalization of Pt(II) Complexes

Pt(II) compounds PtXP (X = 1, 2) were prepared from dichloride starting compounds according to the published protocol [19,70] through attaching of alkynyl ligands by Sonogashira coupling reaction that is the general method used for similar Pt(II) derivatives (Scheme 5). As for the Re(I) complexes described above, the deprotection of the C≡C bond occurs under very mild conditions in methanol solution at room temperature but using KF instead of carbonate gives higher PtX (X = 1, 2) yields. However, in contrast to our expectations and published data [19,21], the click-reactions of PtXP or PtX complexes with 2-azidomethylpyridine did not result in the formation of PtXC. We tested various conditions (solvent, time and Cu(I) catalyst), but only an unidentified mixture of compounds was obtained.
Supposing that the σ-alkynyl ligands can be non-innocent, we replaced the C≡C substituents in the NN ligand with a tBu group and performed a test click-reaction with 2-azidomethylpyridine. In this case, the reaction takes a slight heating and 48 h to be completed and resulting in the formation of only one product, namely Pt(dtbpy)C complex (Scheme 6).
The structures of Pt(dtbpy)C were established by 1H NMR experiment, ESI MS (positive mode), and single crystal X-ray diffraction (XRD) analyses.
The number of signals in 1H NMR spectrum, their relative integral intensities and multiplicities (Figures S18 and S19) clearly indicate the presence of NN ligand and two triazol rings substituted of methylpyridine and phenyl that confirm that both C≡C bonds have reacted with 2-azidomethylpyridine in contrast to the early reported unsymmetrical product obtained in a similar process [20]. ESI+ MS spectra of Pt(dtbpy)C complex demonstrate the signals of [M + H]+ and [M + Na]+ molecular ions with expected m/z values and isotopic patterns matching the calculated ones (Figure S20).
The molecular structure of Pt(dtbpy)C is shown in Figure 1 and Figure S21; crystallographic data are given in Table S1, and selected structural parameters are listed in Table S2. The molecular arrangement of Pt(II) center found in the crystals correlates well with spectroscopic data obtained for Pt(dtbpy)C in solution. According to the XRD data, the platinum metallocentre possesses a square-planar environment, with a chelating ring formed after the coordination of dtbpy ligand, and two monodentate triazolyl rings. To the best of our knowledge, there is only one example of platinum triazolyl-bipyridine complexes, already mentioned [20], but it is unsymmetrical. There are also three more examples of triazolyl Pt(II) complexes, bearing ancillary tridentate ligands. [71] Being the very first example of bis-triazolyl Pt(II) complexes, Pt(dtbpy)C, however, can be compared to the above-mentioned four compounds, as they possess close structural units. Selected structural parameters for the complex described are listed in Table S2.
The bond lengths Pt1–N1 and Pt1–N2 in Pt(dtbpy)C fall into the literature range. On the contrary, the bond Pt1–C19 is slightly longer than described in the literature (Table S2). Bond lengths of triazolyl rings also fall in the literature range. The plane of the triazolyl ring is almost perpendicular to the Pt-dtbpy plane. Pyridine rings are placed in opposite directions, presumably due to steric hindrances. Complex Pt(dtbpy)C is packed in a layered fashion with no π-π interactions.

2.5. Photophysical Properties of Functionalized Complexes

Diimine complexes of Re(I),Pt(II) and Ir(III) are luminescent, and the electronic states of the NN ligand are usually involved in the transitions responsible for photoemission. In order to understand how the post-synthetic modification of the NN ligand affects the luminescence properties of the complex, we investigated their basic photophysical characteristics (Table S3).
The UV-vis absorption spectra of “clicked” complexes (Figure 2) are typical for carbonyl Re(I), alkynyl Pt(II) and cyclometallated Ir(III) complexes with NN ligands, and exhibit strong bands in UV range from ca. 250 to ca. 400 nm due to spin-allowed ππ* intraligand IL transitions located at the aromatic system of ligands. The low energy edge from ca. 400 to ca. 500 nm range may include the contribution of ligand-to-ligand charge transfer LLCT as well as the MLCT transitions from metal dπ-orbitals to π*-orbital of ligands.
All complexes studied exhibit emission in the visible region of the spectrum when excited by UV light. The large Stokes shift, oxygen sensitivity and afterglow lifetime in the microsecond domain clearly indicate a triplet origin of the luminescence, which is typical for this kind of compound. The post-synthetic modification of the NN ligand has a minor effect on the emission energy for Ir(III) complexes, which undergoes a non-significant bathochromic shift ca. 20 nm compared to [Ir(ppy)2(bpy)]+em ca. 600 nm) [72,73] and [Ir(dfppy)2(bpy)]+em ca. 520 nm) [74]. At the same time, the emission energy of the Re(I) complexes undergoes a significant bathochromic shift up to 80 nm compared to Re(CO)3X(bpy) and Re(CO)3X(phen), X = Cl, Br [75].
The post-synthetic conversion of Pt(dtbpy) in Pt(dtbpy)C involves not NN ligand but σ-coordinated C≡C bond. In contrast to the complexes described above, the emission energy of Pt(dtbpy)C undergoes a hypsochromic shift compared to the luminescence energy of Pt(dtbpy) [76] and the closest analog Pt(II) complex with one clicked C≡C bond [20].

3. Discussion

The application of the “chelate, then click” strategy [8] allows post-synthetic modification of the ligand sphere of transition metal complexes with functional groups reactive to azide. In our study, 2-azidomethylpyridine was used to demonstrate a proof-of-principle solution for the selective creation of an additional donor site on the periphery of a complex, and the complementary part for the triazole ring assembly was a terminal C≡C bond attached to the aromatic system of NN ligand (Scheme 1). Thus, the efficiency of the synthetic approach for different metallocentres has been illustrated. We do not pretend to a comprehensive conclusion and consider the results of our study as a contribution to the data bank to support the development of experimental approaches to the post-functional modification of transition metal complexes.
For all reaction systems studied here, a mixture of copper sulfate pentahydrate and sodium ascorbate was used as the Cu(I) source to catalyze the click-reaction. This choice is traditional, however other Cu(I,II) salts or complexes can also be used (see Table A1).
Based on the data obtained herein, Re(I) carbonyl complexes (Scheme 3) are the most convenient platform for post-functionalization by click-reaction. All three steps (preparation of the complex with protected C≡C bond on the periphery, deprotection, and click-reaction) are simple and occur under mild conditions with high yields. It is interesting to note that Re(I) complexes are also popular in the literature as convenient platforms for post-functionalization by click-reaction [6,10,11]. An important factor is that the information on the chemical behavior of rhenium complexes can be applied to similar technetium complexes, which is of fundamental importance for obtaining agents for radio-diagnostics [7,8,9].
Cyclometallated heteroleptic Ir(III) complexes with the composition [Ir(C^N)(NN)]+ represent one of the most popular molecular platforms for many applications due to their chemical stability and variability of luminescent properties. These complexes are ideal candidates for post-synthetic modification, including modification by click-reactions. However, the range of Ir(III) complexes with a terminal C≡C bond at the periphery of the ligand environment that are used for CuCAAC reaction is relatively limited [12,13,14,15]. The literature mainly describes the use of click-reaction to obtain a variety of triazole-based ligands for Ir(III) complexes, the range of which is huge.
In the present study, two cyclometallating C^N ligands with different donor-acceptor properties were selected and four Ir(III) complexes with protected C≡C bonds at the periphery of the ligand environment were synthesized (Scheme 4). It has been found that the combination of two factors significantly influences deprotection as well as following click-reaction. These factors are the nature of the C^N ligand and the position of the C≡C bond in the aromatic ring of the NN ligand. In particular, deprotection of complex Ir(dfppy)2P (acceptor C^N ligand and C≡C bond in the para-position to the nitrogen atom) does not proceed smoothly and the click-reaction is very ineffective. Perhaps, the location of the triple bond in the para-position to the nitrogen atom in combination with the strong acceptor properties of the {Ir(dfppy)2} metallocentre leads to sufficient lowering of electron density on the triple bond, which results in its deactivation toward to the CuAAC reaction.
In contrast to Re(I) (Scheme 3) and Ir(III) (Scheme 4) complexes, diimine Pt(II) complexes with σ-alkynyl ligands (Scheme 5) are objects of dual functionality. The presence of a terminal and σ-coordinated C≡C group increases the variability of the system since both types of triple carbon–carbon bonds can be click-reactive [20]. The literature examples demonstrate that the terminal C≡C group attached to the NN ligand has an advantage in the CuAAC reaction over the σ-alkynyl ligand [19,21,22] or the −C≡C− bond within the oligomer chain [16,17,18]. Our results are in contrast to this picture because no individual product was obtained in the click-reaction of PtX complexes with 2-azidomethylpyridine (Scheme 5). At the same time, click-reaction of the complex Pt(dtbpy) with the same substrate results in only one product (Scheme 6) with an acceptable yield. It is important to note that despite the external simplicity of the CuAAC reaction, this process is sensitive to many parameters including the nature of the azido-derivative. In particular, in the literature examples reviewed, this participant of the reaction does not have an additional donor site with high nucleophilicity (i.e., pyridine).

4. Materials and Methods

4.1. General Comments

The 1H and 1H1H COSY NMR spectra were recorded on a Bruker 400 MHz Avance spectrometer (Daltonics, Bremen, Germany); chemical shifts were referenced to the residual solvent signals. Mass spectra were measured on a Bruker Daltonik MaXis HRMS-ESI-QTOF instrument (Daltonics, Bremen, Germany) operating in positive mode. Preparation of protected NN compounds XP (X = 1–3) [68,69] and Re1P [77] was carried on using the adapted protocol described in the literature, compounds PtXP and PtX (X = 1, 2) were prepared according to the published protocol [19,70]. [Pt(dtbpy)(C≡CPh)2] [78], 2-azidomethylpyridine [39], and cyclometalated Ir(III) dimers [Ir(C^N)2Cl]2 [79] were synthetized according to the published procedure. All other reagents and solvents were purchased from Merck (St. Louis, MO, USA), Alfa Aesar (Ward Hill, MA, USA), Fluka (Germany) and Vekton (St. Petersburg, Russia)) and used without further purification.

4.2. Synthetic Methods

5,5’-trimethylsilylethynyl-2,2’-bipyridine, 1P. Schlenk flask was charged with 75 mg Pd(PPh3)2Cl2, 37,5 mg CuI, 375mg 4,4’-dibromo-2,2’-bipyridine, 25 mL of absolute THF, 580 µL trimethylsilylacetylene and 4 mL of diisopropylamine, then evacuated and filled with argon. The mixture was stirred at 45°C for 40 h. After the process was finished, the solvent was evaporated. Then the residue was dissolved in 25 mL of dichloromethane, and a small amount of activated carbon and water solution containing an excess of NaCN were added. The resulting mixture was filtered through Celites. The organic phase was washed with two portions of water and dried over Na2SO4. The product was purified by column chromatography (Silica gel, eluted with DCM, Rf = 0.35). Off-white powder, yield 332 mg (79%). 1H NMR (400 MHz, acetone-d6): δ 8.74 (d, J = 2.1 Hz, 2H), 8.47 (d, J = 8.3 Hz, 2H), 8.00 (dd, J = 8.2, 2.2 Hz, 2H), 0.30 (s, 18H). ESI HRMS [M+H]+ calcd. C20H25N2Si2 349.1551, found 349.1552, [M+Na]+ calcd. C20H24N2Si2Na 371.1370, found 371.1392.
4,4’-trimethylsilylethynyl-2,2’-bipyridine, 2P. The same procedure as for 1P. Product was purified by column chromatography (Silica gel, eluted with DCM, then EtOAc/hexane/DCM. Rf (EtOAc/hexane/DCM) = 0.75). Off-white powder, yield 296 mg (71%). 1H NMR (400 MHz, acetone-d6): δ 8.71 (d, J = 4.9 Hz, 2H), 8.46 (m, 2H), 7.46 (dd, J = 4.8, 1.7 Hz, 2H), 0.31 (s, 18H). ESI HRMS (m/z): calcd. For [M+H]+ calcd. C20H25N2Si2 349.1551, found 349.1539.
3,8-trimethylsilylethynyl-1,10-phenanthroline, 3P. The same procedure as for 1P. Product was purified by column chromatography (Silica gel, eluted with DCM/hexane 1:1, then EtOAc/hexane/DCM. Rf (EtOAc/hexane/DCM) = 0.85). White solid, yield 86 mg (21 %). 1H NMR (400 MHz, acetone-d6) δ 9.11 (m, 2H), 8.55 (d, J = 2.1 Hz, 2H), 8.02 (s, 2H), 0.33 (s, 18H). ESI HRMS [M+H]+ calcd. C22H25N2Si2 373.1551, found 373.1464, [M+Na]+ calcd. C22H24N2Si2Na 395.1370, found 395.1359.
Re1P. Pentacarbonylrhenium(I) chloride (50 mg, 0.14 mmol) and 1P (1 eq) were suspended in toluene (15 mL) and degassed by purging nitrogen for 15 min upon stirring. Then, the reaction mixture was refluxed for 3 h under an argon atmosphere to give a dark orange solution. The solvent was removed by rotary evaporation to give an oil. The product was dissolved in dichloromethane and precipitated in hexane, the solid obtained was washed with pentane and dried. Orange solid, yield 78 mg (85%). 1H NMR (400 MHz, chloroform-d): δ 9.08 (dd, J = 1.8, 0.8 Hz, 2H), 8.09 – 8.01 (m, 4H), 0.31 (s, 18H). ESI+ MS (m/z) 619.08 [M – Cl]+ (calcd. C23H24N2O3ReSi2 619.08), 647.09 [M – Cl + N2]+ (calcd. C23H24N4O3ReSi2 647.09), 677.02 [M + Na]+ (calcd. C23H24ClN2NaO3ReSi2 677.04).
Re2P. Pentacarbonylrhenium(I) chloride (50 mg, 0.14 mmol) and 2P (50 mg, 0.14 mmol) were suspended in toluene (15 mL) and degassed by purging nitrogen for 15 min upon stirring. Then the reaction mixture was refluxed for 3 h under an argon atmosphere to give an orange suspension. The precipitate was collected, washed with diethyl ether, and dried. Orange solid, yield 86 mg (96%). 1H NMR (400 MHz, acetone-d6): δ 9.44 (d, J = 1.8 Hz, 2H), 8.99 (d, J = 1.8 Hz, 2H), 8.32 (s, 2H), 0.35(s, 18H). ESI+ MS (m/z) 619.08 [M – Cl]+ (calcd. C23H24N2O3ReSi2 619.08), 647.09 [M – Cl + N2]+ (calcd. C23H24N4O3ReSi2 647.09), 677.03 [M + Na]+ (calcd. C23H24ClN2NaO3ReSi2 677.04).
Re3P. The same procedure as for Re1P. Orange solid, yield 88 mg (94%). 1H NMR (400 MHz, chloroform-d): δ 8.95 (d, J = 1.8 Hz, 2H), 8.14 (bs, 2H), 8.47 (dd, J = 5.7, 1.6 Hz, 2H), 0.35 (s, 18H). ESI+ MS (m/z) 643.07 [M – Cl]+ (calcd. C25H24N2O3ReSi2 643.08), 671.09 [M – Cl + N2]+ (calcd. C25H24N4O3ReSi2 671.09), 701.03 [M + Na]+ (calcd. C25H24ClN2NaO3ReSi2 701.04).
General procedure for the synthesis of complexes ReX. To a stirred suspension of ReXP (40 mg, 0.06 mmol) in MeOH (15 mL) K2CO3 (25 mg, 0.018 mmol) was added. The reaction mixture was stirred for 4 h, and then the solvent was removed by rotary evaporation. The product was dissolved in dichloromethane and washed with water, the organic layer was separated and dried over MgSO4. The solvent was removed to a minimum by rotary evaporation and the residue was precipitated in hexane. The precipitate was collected, washed with diethyl ether and dried.
Re1, orange solid, yield 40 mg (80%) 1H NMR (400 MHz, acetone-d6): δ 9.16 (d, J = 1.9 Hz, 2H), 8.76 (d, J = 8.5 Hz, 2H), 8.42 (dd, J = 8.5, 2.0 Hz, 2H), 4.34 (s, 2H). ESI+ MS (m/z) 475.00 [M – Cl]+ (calcd. C17H8N2O3Re 475.01), 503.01 [M – Cl + N2]+ (calcd. C17H8N4O3Re 503.01), 532.96 [M + Na]+ (calcd. C17H8ClN2NaO3Re 532.96).
Re2, orange solid, yield 66 mg (84%).1H NMR (400 MHz, chloroform-d): δ 9.04 (d, J = 5.8, 2H), 8.22 (s, 2H), 7.58 (dd, J = 5.7, 1.6 Hz, 2H), 3.67 (s, 2H). ESI+ MS (m/z) 475.00 [M – Cl]+ (calcd. C17H8N2O3Re 475.01), 503.01 [M – Cl + N2]+ (calcd. C17H8N4O3Re 503.01), 532.95 [M + Na]+ (calcd. C17H8ClN2NaO3Re 532.96).
Re3, orange solid, yield 55 mg (87%). 1H NMR (400 MHz, acetone-d6) δ 9.50 (s, 2H), 9.07 (s, 2H), 8.36 (s, 2H), 4.39 (s, 2H). ESI+ MS (m/z) 499.01 [M – Cl]+ (calcd. C19H8N2O3Re 499.01), 527.01 [M – Cl + N2]+ (calcd. C19H8N4O3Re 527.01), 556.96 [M + Na]+ (calcd. C19H8ClN2NaO3Re 556.96).
General procedure for the synthesis of complexes ReXC. ReX (15 mg, 0.03 mmol) was dissolved in acetone (8 mL) and 2-azidomethylpyridine was added. To the stirring mixture, an aqueous solution (2 mL) of CuSO4×5H2O (8 mg, 0.03 mmol) и NaAsc (12 mg, 0.06 mmol) was added. The reaction mixture was stirred for 12 h and then the solvent was removed. The crude product was dissolved in dichloromethane (30 mL) and washed with NH4OH/EDTA solution. The organic layer was separated and dried over MgSO4 and the solvent was evaporated. The product was dissolved in dichloromethane and precipitated in hexane. The precipitate was collected, washed with diethyl ether and dried.
Re1C, yellow solid, 19 mg (85%). 1H NMR (400 MHz, acetone-d6): δ 9.61 (s, 1H), 8.93 (s, 1H), 8.75 (d, J = 2.5 Hz, 2H), 8.59 (d, J = 4.6 Hz, 1H), 7.86 (d, J = 1.8 Hz, 1H), 7.44 (d, J = 7.8 Hz, 1H), 7.41 – 7.35 (m, 1H), 5.89 (s, 2H). ESI+ MS (m/z) 743.12 [M – Cl]+ (calcd. C29H20N10O3Re 743.13), 771.13 [M – Cl + N2]+ (calcd. C29H20N12O3Re 771.13), 801.08 [M + Na]+ (calcd. C29H20ClN10NaO3Re 801.08).
Re2C, yellow solid, 21 mg (70%). 1H NMR (400 MHz, chloroform-d): δ 8.92 (s, 2H), 8.73 (d, J = 5.8 Hz, 1H), 8.69 (d, J = 3.7 Hz, 1H), 8.61 (s, 2H), 7.81 (td, J = 7.8, 1.8 Hz, 2H), 7.71 (d, J = 5.8 Hz, 2H), 7.41 (d, J = 7.9 Hz, 2H), 7.34 (dd, J = 7.6, 4.9 Hz, 2H), 5.89 – 5.74 (m, 4H). ESI+ MS (m/z) 743.12 [M – Cl]+ (calcd. C29H20N10O3Re 743.12), 771.12 [M – Cl + N2]+ (calcd. C29H20N12O3Re 771.13), 801.07 [M + Na]+ (calcd. C29H20ClN10NaO3Re 801.08).
Re3C. Yellow solid, yield 53 mg (70%). 1H NMR (400 MHz, acetone-d6) δ 9.97 (d, J = 1.7 Hz, 1H), 9.30 (s, 1H), 9.04 (s, 1H), 8.61 (d, J = 4.9 Hz, 1H), 8.32 (s, 1H), 7.87 (td, J = 7.7, 1.8 Hz, 1H), 7.47 (d, J = 7.9 Hz, 1H), 7.39 (dd, J = 7.4, 4.9 Hz, 1H), 5.92 (s, 2H). ESI+ MS (m/z) 808.15 [M – Cl + CH3CN]+ (calcd. C33H23N11O3Re 808.15), 820.13 [M – Cl + CH2ONa]+ (calcd. C32H22N10NaO4Re 820.13).
General procedure for the synthesis of complexes Ir(C^N)XP. 2.1 eqv. 1P or 2P was added to [Ir(C^N)2Cl]2 and suspended in a mixture of CH2Cl2/MeOH (3:1), followed by the addition of an excess of KPF6. The reaction mixture was stirred for 24h at room temperature. After filtration through the Celite the resulting solution was evaporated to dryness. A small amount of CH2Cl2 was added, then a solid was precipitated with hexane. The precipitate was washed with diethyl ether twice.
Ir(ppy)1P. Dark-red powder. Yield: 50.6 mg (64%). 1H NMR (400 MHz, Acetone-d6): δ 8.87 (d, J = 8.5 Hz, 2H), 8.34 – 8.24 (m, 4H), 8.07 – 7.90 (m, 10H), 7.23 – 7.15 (m, 2H), 7.08 (td, J = 7.5, 1.3 Hz, 2H), 6.97 (td, J = 7.4, 1.3 Hz, 2H), 6.35 (d, J = 7.2 Hz, 2H), 0.20 (s, 18H). ESI+ MS (m/z) 849.24 [M – PF6]+ (calcd. C42H40IrN4Si2 849.24).
Ir(dfppy)1P. Dark-yellow powder. Yield: 54.3 mg (54%). 1H NMR (400 MHz, Acetone-d6): δ 8.91 (d, J = 8.4 Hz, 2H), 8.45 – 8.39 (m, 2H), 8.38 – 8.33 (m, 2H), 8.14 – 8.03 (m, 6H), 7.31 – 7.24 (m, 2H), 6.85 – 6.75 (m, 2H), 5.80 (dd, J = 8.5, 2.3 Hz, 2H), 0.21 (s, 18H). ESI+ MS (m/z) 921.20 [M – PF6]+ (calcd. C42H36F4IrN4Si2 921.20).
Ir(ppy)2P. Orange powder. Yield: 41.8 mg (53%). 1H NMR (400 MHz, Acetone-d6): δ 8.97 (s, 2H), 8.26 (d, J = 8.1 Hz, 2H), 8.09 (d, J = 5.7 Hz, 2H), 7.99 (t, J = 7.7 Hz, 2H), 7.91 (d, J = 6.8 Hz, 4H), 7.71 (d, J = 5.6 Hz, 2H), 7.19 (t, J = 6.5 Hz, 2H), 7.06 (t, J = 7.5 Hz, 2H), 6.94 (t, J = 7.4 Hz, 2H), 6.34 (d, J = 7.5 Hz, 2H), 0.30 (s, 18H).
Ir(dfppy)2P. Orange powder. Yield: 60.5 mg (61%). 1H NMR (400 MHz, Acetone-d6): δ 9.00 (d, J = 1.0 Hz, 2H), 8.47 – 8.36 (m, 2H), 8.23 – 8.17 (m, 2H), 8.14 – 8.06 (m, 2H), 8.05 – 7.96 (m, 2H), 7.73 (dd, J = 5.8, 1.7 Hz, 2H), 7.28 (ddd, J = 7.5, 5.8, 1.4 Hz, 2H), 6.78 (ddd, J = 12.8, 9.3, 2.4 Hz, 2H), 5.79 (dd, J = 8.5, 2.4 Hz, 2H), 0.30 (s, 18H). ESI+ MS (m/z) 921.20 [M – PF6]+ (calcd. C42H36F4IrN4Si2 921.20).
General procedure for the synthesis of complexes Ir(C^N)X. Ir(C^N)XP (except Ir(dfppy)2P where 1.2 eqv. KOH was added) was dissolved in MeOH and stirred overnight, then filtered through a pad of Celite and the solvent was evaporated to obtain the powder product.
Ir(ppy)1. Dark-orange powder. Yield: 45 mg (89%). 1H NMR (400 MHz, Acetone-d6): δ 8.90 (d, J = 8.5 Hz, 2H), 8.37 (d, J = 8.5 Hz, 2H), 8.27 (d, J = 8.2 Hz, 2H), 8.11 (s, 2H), 8.04 – 7.96 (m, 4H), 7.94 (d, J = 7.7 Hz, 2H), 7.18 (t, J = 6.7 Hz, 2H), 7.08 (t, J = 7.6 Hz, 2H), 6.97 (t, J = 7.5 Hz, 2H), 6.37 (d, J = 7.5 Hz, 2H), 4.20 (s, 2H). ESI+ MS (m/z) 705.16 [M – PF6]+ (calcd. C36H24IrN4 705.16).
Ir(dfppy)1. Orange powder. Yield: 30 mg (64%). 1H NMR (400 MHz, Acetone-d6): δ 8.94 (d, J = 8.5 Hz, 2H), 8.42 (dd, J = 8.5, 2.1 Hz, 4H), 8.20 (d, J = 1.9 Hz, 2H), 8.13 – 8.05 (m, 4H), 7.26 (ddd, J = 7.4, 5.8, 1.4 Hz, 2H), 6.80 (ddd, J = 12.8, 9.4, 2.4 Hz, 2H), 5.81 (dd, J = 8.5, 2.4 Hz, 2H), 4.23 (s, 2H). ESI MS (m/z) 777.12 [M – PF6]+ (calcd. C36H20F4IrN4 777.13).
Ir(ppy)2. Orange powder. Yield: 34.4 mg (82%). 1H NMR (400 MHz, Acetone-d6): δ 9.04 (s, 2H), 8.26 (d, J = 8.2 Hz, 2H), 8.12 (d, J = 5.6 Hz, 2H), 7.99 (td, J = 7.8, 1.5 Hz, 2H), 7.96 – 7.89 (m, 4H), 7.77 (dd, J = 5.7, 1.7 Hz, 2H), 7.23 – 7.14 (m, 2H), 7.06 (td, J = 7.5, 1.3 Hz, 2H), 6.94 (td, J = 7.4, 1.3 Hz, 2H), 6.34 (dd, J = 7.5, 1.2 Hz, 2H), 4.47 (s, 2H). ESI+ MS (m/z) 705.16 [M – PF6]+ (calcd. C36H24IrN4 705.16).
Ir(dfppy)2. Dark-yellow powder. Yield: 28.4 mg (47%). 1H NMR (400 MHz, Acetone-d6): δ 8.77 (d, J = 1.7 Hz, 2H), 8.41 (d, J = 8.3 Hz, 2H), 8.09 (d, J = 5.9 Hz, 4H), 7.95 – 7.83 (m, 2H), 7.67 (dd, J = 5.7, 1.7 Hz, 2H), 7.26 (ddd, J = 7.4, 5.8, 1.4 Hz, 2H), 6.84 – 6.69 (m, 2H), 5.87 – 5.73 (m, 2H), 4.73 (t, J = 5.3 Hz, 2H).
General procedure for the synthesis of complexes Ir(C^N)XC. Ir(C^N)X (1 eqv.) was dissolved in 12 mL of acetone and 2-(azidomethyl)pyridine (2.3 eqv.) was added. CuSO4×5H2O (2 eqv.) and sodium ascorbate (4 eqv.) were suspended in 3 mL of water and added to the solution. The reaction mixture was stirred for 48h at room temperature. Water and an aqueous solution of NH4OH/EDTA were added to the resulting solution and stirred for 30 min. The organic phase was collected, dried over MgSO4, and evaporated to give an oil. It was dissolved in a small amount of DCM and hexane was added to precipitate a solid. The product was washed with diethyl ether and dried in vacuo.
Ir(ppy)1C. Orange powder. Yield: 15.1 mg (45%). 1H NMR (400 MHz, Acetone-d6): δ 8.91 (d, J = 8.5 Hz, 2H), 8.68 (d, J = 2.0 Hz, 2H), 8.65 (dd, J = 8.4, 2.1 Hz, 2H), 8.57 (d, J = 4.5 Hz, 2H), 8.46 (s, 2H), 8.24 (d, J = 8.2 Hz, 2H), 8.04 (d, J = 5.9 Hz, 2H), 8.00 – 7.89 (m, 4H), 7.85 (td, J = 7.8, 1.9 Hz, 2H), 7.43 (d, J = 7.8 Hz, 2H), 7.38 (dd, J = 7.7, 4.9 Hz, 2H), 7.19 – 7.12 (m, 2H), 7.11 – 7.05 (m, 2H), 6.97 (td, J = 7.4, 1.1 Hz, 2H), 6.42 (d, J = 7.4 Hz, 2H), 5.78 (s, 4H), ESI+ MS (m/z) 973.29 [M – PF6]+ (calcd. C48H36IrN12 973.28).
Ir(dfppy)1C. Orange powder. Yield: 23.2 mg (72%). 1H NMR (400 MHz, Acetone-d6): δ 8.98 – 8.93 (m, 2H), 8.73 – 8.68 (m, 4H), 8.59 (s, 2H), 8.56 (d, J = 4.4 Hz, 2H), 8.42 – 8.38 (m, 2H), 8.13 (d, J = 5.2 Hz, 2H), 8.06 (t, J = 7.8 Hz, 3H), 7.85 (td, J = 7.7, 1.7 Hz, 2H), 7.43 (d, J = 7.8 Hz, 2H), 7.40 – 7.35 (m, 2H), 7.24 (t, J = 6.1 Hz, 2H), 6.85 – 6.78 (m, 2H), 5.87 (dd, J = 8.5, 2.3 Hz, 2H), 5.78 (s, 4H). ESI+ MS (m/z) 1045.24 [M – PF6]+ (calcd. C48H32F4IrN12 1045.24).
Ir(ppy)2C. Orange powder. Yield: 9.1 mg (48%). 1H NMR (400 MHz, Acetone-d6): δ 9.38 (s, 2H), 9.00 (s, 2H), 8.60 – 8.55 (m, 2H), 8.26 (d, J = 8.1 Hz, 2H), 8.18 – 8.14 (m, 2H), 8.14 – 8.11 (m, 2H), 8.02 – 7.96 (m, 4H), 7.94 – 7.91 (m, 2H), 7.87 (td, J = 7.7, 1.7 Hz, 2H), 7.48 (d, J = 7.8 Hz, 2H), 7.41 – 7.36 (m, 2H), 7.17 (t, J = 6.6 Hz, 2H), 7.06 (t, J = 7.5 Hz, 2H), 6.95 (t, J = 6.9 Hz, 2H), 6.39 (d, J = 7.5 Hz, 2H), 5.88 (s, 4H). ESI+ MS (m/z) 973.28 [M – PF6]+ (calcd. C48H36IrN12 973.28).
Pt1P. This complex was prepared according to the literature methods [19]. Brownish powder, yield 22.7 mg (62%). 1H NMR (400 MHz, Chloroform-d) δ 9.82 (m, 2H, dtbpy), 8.13 (t, J = 8.0 Hz, 2H, dtbpy), 8.05 (d, J = 8.5 Hz, 2H, dtbpy), 7.56 (m, 6H, Ph), 7.32 (m, 6H, Ph), 7.20 (t, J = 7.5 Hz, 4H, Ph), 3.75 (s, 2H, alkynyl-H). ESI+ MS (m/z) 640.07 [M + K]+ (calcd. C30H18KN2Pt 640.07).
Pt(dtbpy)C. Pt(dtbpy)(C≡CPh)2 (12.6 mg, 0.019 mmol) and 2-azidomethylpyridine (5 mg, 0.047 mmol) were dissolved in 20 mL of CH2Cl2. 3 eq CuSO4×5H2O (14 mg, 0.057 mmol) and NaAsc (22.5 mg, 0.114 mmol) were dissolved in 5 mL of MeOH, sonicated for 3 min and added to the acetone solution. The reaction mixture was flushed with argon and stirred for 48 h at 40 °C. The solvents were removed in vacuo, and 20 mL of CHCl3, 10 mL of water and 2 mL of 5% EDTA solution were added to the residue. The resulting mixture was vigorously stirred for 30 min, then the organic layer was separated, and the solvent was removed in vacuo to a minimal volume. Then, diethyl ether was added to obtain a crude product as a yellow solid. The resulting powder was washed with diethyl ether, dried in vacuo and recrystallized from an acetone solution. Yellow crystals, yield 12 mg (69%). 1H NMR (400 MHz, DMSO-d6): δ 8.54 (d, J = 1.8 Hz, 2H, dtbpy), 8.36 (m, 4H, i-click (o-Ph)), 8.11 (m, 2H, i-click (py)), 7.89 (d, J = 6.0 Hz, 2H, dtbpy (H3, H4)), 7.49 (dd, J = 6.0 Hz, 1.8 Hz, 2H, dtbpy (H2, H5)), 7.44 (td, J = 7.7 Hz, 1.8 Hz, 2H, i-click (py)), 7.12 (m, 4H, i-click (m-Ph)), 7.07 (m, 2H, i-click (p-Ph)), 7.03 (m, 2H, i-click (py)), 6.64 (d, J = 7.9 Hz, 2H, i-click (py)), 5.36 (d, J = 15.8 Hz, 2H, CH2), 5.36 (d, J = 15.8 Hz, 2H, CH2), 1.35 (s, 21H, tBu). ESI+ MS (m/z) 934.36 [M + H]+ (calcd. C46H46N10Pt 934.36). Single crystals of Pt(dtbpy)C were obtained by slow evaporation of an acetone solution at room temperature.

4.3. X-ray Crystal Structure Determination

The crystal structure of Pt(dtbpy)C was determined by the means of single crystal XRD analysis using a Rigaku Oxford Diffraction XtaLAB HyPix-3000 diffractometer (Rigaku Oxford Diffraction, Oxford, UK) for the data collection at a temperature of 170K. Diffraction data were processed in the CrysAlisPro program (Agilent Technologies, Version 1.171.39.35a, Stockport, UK) [80]. The unit-cell and refinement parameters are listed in Table S1. The structures were solved by the dual-space algorithm and refined using the SHELX programs [81,82] incorporated in the OLEX2 program package [83]. Supplementary crystallographic data for this paper have been deposited at Cambridge Crystallographic Data Centre and available online: https://www.ccdc.cam.ac.uk/structures/ (accessed on 14 August 2022). Pt(dtpby)c: (P21/n (14), a = 11.56500(10), b = 20.6552(2), c = 18.3102(2) Å; α = 90, β = 99.1890(10), γ = 90; V = 4317.76(7) Å3; Z = 4; R1 = 0.0247%; CCDC 2204302).

5. Conclusions

In the present study, we applied the CuAAC reaction for the post-synthetic modification of diimine mononuclear complexes Re(I), Pt(II) and Ir(III) with terminal C≡C bonds at the periphery of the ligand environment and demonstrated the efficiency of the synthetic approach for different metallocentres. All functionalized complexes obtained were characterized by spectroscopic methods. It was found that the position of C≡C bonds in the aromatic ring of the diimine ligand and the composition of the ligand environment forming the first coordination sphere can influence the post-synthetic modification of Ir(III) complexes. The implementation of the standard click protocol for Pt(II) bis-alkynyl complexes bearing the same diimine ligands did not result in the formation of the corresponding click products clearly demonstrating competition between the terminal and organoplatinum acetylene units in the click process. While replacing the diinine with “click-innocent” one, the reaction occurs as a platinum iclick process and bis-triazolyl derivative was obtained in good yield. It was found that post-synthetic modification of the far periphery of the diimino ligand in Re(I) and Ir(III) complexes by click-reaction can change the luminescence properties of the metallocentre, which should be considered when designing molecular systems that exploit the emission properties of the complex. Thus, the data obtained in the present study, together with literature data, show that click-reaction has a high potential for the construction of complex molecular systems involving transition metal complexes and can be used as a powerful tool for the rational molecular design of smart functional materials.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules27196494/s1, X-ray structure determination; Table S1: Crystallographic data for compound Pt(dtbpy)C; Figure S1: 1H NMR spectra of NN compounds XP (X = 1−3); Figure S2: 1H NMR spectra of Re1X complexes; Figure S3: 1H NMR spectra of Re2X complexes; Figure S4: 1H NMR spectra of Re3X complexes; Figure S5: ESI+ MS spectra of Re1X complexes with isotope pattern of key signals; Figure S6: ESI+ MS spectra of Re2X complexes with isotope pattern of key signals; Figure S7: ESI+ MS spectra of Re3X complexes with isotope pattern of key signals; Figure S8: 1H NMR spectra of Ir(ppy)1X complexes; Figure S9: 1H NMR spectra of Ir(ppy)2X complexes; Figure S10: 1H NMR spectra of Ir(dfppy)1X complexes; Figure S11: 1H NMR spectra of Ir(dfppy)2X complexes; Figure S12: ESI+ MS spectra of Ir(ppy)1X complexes with isotope pattern of key signals; Figure S13: ESI+ MS spectra of Ir(ppy)2X complexes with isotope pattern of key signals; Figure S14: ESI+ MS spectra of Ir(dfppy)1X complexes with isotope pattern of key signals; Figure S15: ESI+ MS spectra of Ir(dfppy)2P complexes with isotope pattern of key signals; Figure S16: 1H NMR spectra of Pt1X complexes; Figure S17: 1H NMR spectra of Pt2X complexes; Figure S18: 1H NMR spectrum of Pt(btbpy)C complexes in aromatic region, DMSO-d6, RT; Figure S19: 1H1H COSY NMR spectrum of Pt(btbpy)C complexes in aromatic region, DMSO-d6, RT; Figure S20: ESI+ MS spectrum of Pt(dtbpy)C complex with isotope pattern of key signals; Figure S21: Molecular structure of Pt(dtbpy)C complex with key atoms numeration; Table S2: Selected structural parameters for compound Pt(dtbpy)C and literature data for Pt(II) triazolyl derivatives; Table S3: Photophysical properties of “clicked” complexes ReXC and IrXC.

Author Contributions

Conceptualization, E.G.; Data curation, A.P. and E.G.; Funding acquisition, E.G.; Investigation, S.P., V.K., A.P. and E.A.; Project administration, E.G.; Resources, S.P. and E.G.; Supervision, E.G.; Writing—original draft, S.P., V.K. and E.G.; Writing—review & editing, E.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Russian Science Foundation, grant number 16-13-10064.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data reported herein are accompanying the present article.

Acknowledgments

This work was carried out using the equipment of St Petersburg University Research Park (Centres of Magnetic Resonance, Optical and Laser Materials Research, Chemical Analysis and Materials Research, X-ray Diffraction).

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of all compounds described in the article are available from the authors.

Appendix A

Table A1. Some literature examples of C≡C peripherally functionalized metallocenters, azide substrates and click-reaction conditions.
Table A1. Some literature examples of C≡C peripherally functionalized metallocenters, azide substrates and click-reaction conditions.
##MetalSubstrateConditions; yieldReference
1 Re(I), Mo(I)Molecules 27 06494 i001Cu(SO4)2 × 5H2O, NaAsc, THF/H2O, r.t., 120 h; up to 88%[6]
2 Re(I), Tc(I)Molecules 27 06494 i002Cu(MeCOO)2, PBS, NaAsc, tBuOH/H2O, r.t., 1.5 h; 2–84%[7]
3 Molecules 27 06494 i003Cu(MeCOO)2, NaAsc, tBuOH/H2O, 70 °C, 1.5 h; 81%[8]
4 Re(V), Tc(V)Molecules 27 06494 i004
azido-modified angiotensin-II peptide
Cu(SO4)2 × 5H2O, NaAsc, CH2Cl2/MeOH/H2O, r.t., 5 h; 51%[9]
5 Re(I)Molecules 27 06494 i005Cu(SO4)2 × 5H2O, NaAsc, DMF/H2O, r.t., 24 h; 84%[10]
6 Molecules 27 06494 i006Cu(SO4)2 × 5H2O, NaAsc, THF/H2O, 60 °C, 18 h; 93%[11]
7 Ir(III)Molecules 27 06494 i007Cu(SO4)2 × 5H2O, NaAsc, DMF/H2O, r.t., 12 h; 71–88%[12]
8 Au(PR3)N3DIPEA, DMF, r.t., 48 h; 20%[13]
9 Molecules 27 06494 i008CuBr, MeCN, 70 °C, 12 h[14]
10 Molecules 27 06494 i009CuI, iPr2EtN, MeCN/MeOH/HCl, r.t., 12 h[15]
11 Pt(II)Molecules 27 06494 i010Cu(SO4)2 ×5H2O, NaAsc, DMF/H2O, r.t., up to 16 h; 62–81%[16,17]
12 Molecules 27 06494 i011Cu(SO4)2 × 5H2O, Hasc, DMF/H2O, r.t., 15 h; 62–54%[18]
13 Molecules 27 06494 i012Cu(MeCN)4PF6, Cu, DIPEA, CH2Cl2/MeOH, r.t., 72 h; 73–84%[19]
14 Molecules 27 06494 i013Cu(SO4)2 × 5H2O, Hasc, CH2Cl2/H2O, reflux, 48 h; 75%[20]
15 Molecules 27 06494 i014Cu(MeCN)4PF6, Cu, DIPEA, CH2Cl2/MeOH, r.t., up to 192 h; 17-84%[21]
16 Molecules 27 06494 i015Cu(MeCN)4PF6, Cu, Et(iPr)2N, CH2Cl2/MeOH, reflux, 312 h; 15%[22]
17 Pt(IV)azido-functionalized enterobactinCu(MeCN)4PF6, TBTA, DMF/H2O, r.t., 3 h[24]
18 Ru(II, III)Molecules 27 06494 i016Cu(SO4)2 × 5H2O, NaAsc, tBuOH/H2O, r.t., 12 h; 63%[25]
19 azido-containing poly(benzylether) dendronsCu(SO4)2 × 5H2O, NaAsc, THF/H2O, r.t., 12 h; ca. 50%[26,27]
20 Ru(II)Molecules 27 06494 i017Cu(SO4)2 × 5H2O, NaAsc, (iPr)2EtN, tBuOH/MeOH/H2O, r.t., 18 h; 80%[28]
21 Molecules 27 06494 i018Cu(SO4)2 × 5H2O, NaAsc, CH2Cl2/H2O, r.t., 20 h; 58–81%[29]
22 Molecules 27 06494 i019Cu(SO4)2 × 5H2O, NaAsc, CH2Cl2/H2O, r.t., 12 h; 67–69%[30]
23 Molecules 27 06494 i020Cu(SO4)2 × 5H2O, NaAsc, CH2Cl2/H2O, r.t., 20 h; 61%[31]
24 Molecules 27 06494 i021Cu(SO4)2 × 5H2O, NaAsc, acetone/H2O, r.t., 1h; quantitative[32]
25 Molecules 27 06494 i022Cu(SO4)2 × 5H2O, NaAsc, CH2Cl2/H2O, r.t., 48 h; 55%[33]
26 Fe(II)Molecules 27 06494 i023Cu(SO4)2 × 5H2O, NaAsc, CH2Cl2/NCMe/H2O, r.t., 12 h; 79%[35]
27 Molecules 27 06494 i024Cu(MeCOO)2, toluene, 70 °C, 2 h; 67%[36]
28 ferroceneMolecules 27 06494 i025Cu(SO4)2 × 5H2O, NaAsc, THF/H2O, r.t, 12 h; up to 84%[37]
29 azido-functionalized siloxanesvarious[38]
30 Molecules 27 06494 i026various[39]
31 Molecules 27 06494 i027Cu(SO4)2×5H2O, NaAsc, MeOH/H2O, TTTA, r.t., 20 h; 17-86%[40]
32 Molecules 27 06494 i028Cu(SO4)2×5H2O, NaAsc, THF/H2O, 60°C, 6 h; 69%[41]
Molecules 27 06494 i029Cu(SO4)2×5H2O, NaAsc, THF/H2O, r.t., 24 h; 75%
33 Mn(I)Molecules 27 06494 i030Cu(SO4)2×5H2O, NaAsc, tBuOH/ H2O, r.t., 24 h; 61%[42]
34 Ni(II)Molecules 27 06494 i031CuI, Et3N, DMSO, 70°C, 5 h; 31-92%[43]
35 Zn-porphyrinMolecules 27 06494 i032CuI, DBU, toluene, 70°C, 4 h; 90%[44]
36 Molecules 27 06494 i033CuI, DIPEA, CHCl3, reflux, 12 h; 87%[45]
37 Cr(0), W(0)various various[48]
38 Au(I)azido-amphiphilic oligopeptideCu(SO4)2×5H2O, NaAsc, TBTA, THF/H2O, 24 h, 45°C; up to 87%[49]
39 Tb(III)Molecules 27 06494 i034Cu(SO4)2×5H2O, NaAsc, THPTA, H2O/MeCN, r.t., 12 h, 37%[50]
40 Ln(III)various, including azido-functionalized Ln(III) complexesvarious[51,52,53,54,55,56]
41 {MC3B7H9}, M = Fe, RuMolecules 27 06494 i035Cu(SO4)2×5H2O, NaAsc, tBuOH/H2O, r.t., up to 19 h; up to 82%[57]
42 M-bis(1,2-dicarbollide); M = Fe, Coazido-cholesterolCuI, DIPEA, EtOH, reflux, 12 h[58]
43 POMMolecules 27 06494 i036CuI, DIPEA, MeCN, r.t., 3 h; up to 97%[59]
44 Au(PR3)N3DIPEA, DMF, 35°C, 50 h; 60%[60]
NaAsc = sodium ascorbate. DIPEA = diisopropylethylamin. DBU = 1,8-diazabicyclo[5 .4.0]undec-7-ene. TTTA = tris(1-tert-butyl-1H-1,2,3-triazolyl)methyl amine. POM = polyoxometalate. TBTA = tris(benzyltriazolylmethyl)amine. THPTA = tris(hydroxypropyltriazolylmethyl)amine.

References

  1. Neto, J.S.S.; Zeni, G. A decade of advances in the reaction of nitrogen sources and alkynes for the synthesis of triazoles. Coord. Chem. Rev. 2020, 409, 213217. [Google Scholar] [CrossRef]
  2. Meldal, M.; Tomøe, C.W. Cu-catalyzed azide—Alkyne cycloaddition. Chem. Rev. 2008, 108, 2952–3015. [Google Scholar] [CrossRef] [PubMed]
  3. Liang, L.; Astruc, D. The copper(I)-catalyzed alkyne-azide cycloaddition (CuAAC) “click” reaction and its applications. An overview. Coord. Chem. Rev. 2011, 255, 2933–2945. [Google Scholar] [CrossRef]
  4. Scattergood, P.A.; Sinopoli, A.; Elliott, P.I.P. Photophysics and photochemistry of 1,2,3-triazole-based complexes. Coord. Chem. Rev. 2017, 350, 136–154. [Google Scholar] [CrossRef]
  5. Elliott, P.I.P. (Ed.) Organometallic complexes with 1,2,3-triazole-derived ligands. In Organometallic Chemistry: Volume 39; The Royal Society of Chemistry: Cambridge, UK, 2014; pp. 1–25. [Google Scholar]
  6. Álvarez, C.M.; García-Escudero, L.A.; García-Rodríguez, R.; Miguel, D. Schiff plus click: One-pot preparation of triazole-substituted iminopyridines and ring opening of the triazole ring. Dalt. Trans. 2013, 42, 2556–2561. [Google Scholar] [CrossRef]
  7. Moore, A.L.; Bučar, D.-K.; MacGillivray, L.R.; Benny, P.D. “Click” labeling strategy for M(CO)3 (M = Re, 99mTc) prostate cancer targeted Flutamide agents. Dalt. Trans. 2010, 39, 1926. [Google Scholar] [CrossRef]
  8. Bottorff, S.C.; Moore, A.L.; Wemple, A.R.; Bučar, D.-K.; MacGillivray, L.R.; Benny, P.D. pH-Controlled Coordination Mode Rearrangements of “Clickable” Huisgen-Based Multidentate Ligands with [M(CO)3]+ (M = Re, 99mTc). Inorg. Chem. 2013, 52, 2939–2950. [Google Scholar] [CrossRef]
  9. Castillo Gomez, J.D.; Hagenbach, A.; Abram, U. Propargyl-Substituted Thiocarbamoylbenzamidines of Technetium and Rhenium: Steps towards Bioconjugation with Use of Click Chemistry. Eur. J. Inorg. Chem. 2016, 2016, 5427–5434. [Google Scholar] [CrossRef]
  10. Gillam, T.A.; Caporale, C.; Brooks, R.D.; Bader, C.A.; Sorvina, A.; Werrett, M.V.; Wright, P.J.; Morrison, J.L.; Massi, M.; Brooks, D.A.; et al. Neutral Re(I) Complex Platform for Live Intracellular Imaging. Inorg. Chem. 2021, 60, 10173–10185. [Google Scholar] [CrossRef]
  11. Koenig, J.D.B.; Dubrawski, Z.S.; Rao, K.R.; Willkomm, J.; Gelfand, B.S.; Risko, C.; Piers, W.E.; Welch, G.C. Lowering Electrocatalytic CO2 Reduction Overpotential Using N-Annulated Perylene Diimide Rhenium Bipyridine Dyads with Variable Tether Length. J. Am. Chem. Soc. 2021, 143, 16849–16864. [Google Scholar] [CrossRef]
  12. Passays, J.; Rubay, C.; Marcelis, L.; Elias, B. Synthesis and Photophysical Properties of Triazolyl Ir III Nucleosides. Eur. J. Inorg. Chem. 2017, 2017, 623–629. [Google Scholar] [CrossRef]
  13. Luengo, A.; Marzo, I.; Reback, M.; Daubit, I.M.; Fernández-Moreira, V.; Metzler-Nolte, N.; Gimeno, M.C. Luminescent Bimetallic Ir III/Au I Peptide Bioconjugates as Potential Theranostic Agents. Chem.–A Eur. J. 2020, 26, 12158–12167. [Google Scholar] [CrossRef] [PubMed]
  14. Mandal, S.; Das, R.; Gupta, P.; Mukhopadhyay, B. Synthesis of a sugar-functionalized iridium complex and its application as a fluorescent lectin sensor. Tetrahedron Lett. 2012, 53, 3915–3918. [Google Scholar] [CrossRef]
  15. Wang, J.; Xue, J.; Yan, Z.; Zhang, S.; Qiao, J.; Zhang, X. Photoluminescence Lifetime Imaging of Synthesized Proteins in Living Cells Using an Iridium-Alkyne Probe. Angew. Chem. Int. Ed. 2017, 56, 14928–14932. [Google Scholar] [CrossRef]
  16. Gauthier, S.; Weisbach, N.; Bhuvanesh, N.; Gladysz, J.A. “Click” Chemistry in Metal Coordination Spheres: Copper(I)-Catalyzed 3+2 Cycloadditions of Benzyl Azide and Platinum Polyynyl Complexes trans -(C6F5)(p-tol3P)2Pt(C≡C)nH (n = 2−6). Organometallics 2009, 28, 5597–5599. [Google Scholar] [CrossRef]
  17. Amini, H.; Weisbach, N.; Gauthier, S.; Kuhn, H.; Bhuvanesh, N.; Hampel, F.; Reibenspies, J.H.; Gladysz, J.A. Trapping of Terminal Platinapolyynes by Copper(I) Catalyzed Click Cycloadditions; Probes of Labile Intermediates in Syntheses of Complexes with Extended sp Carbon Chains, and Crystallographic Studies. Chem.–A Eur. J. 2021, 27, 12619–12634. [Google Scholar] [CrossRef]
  18. Clough, M.C.; Zeits, P.D.; Bhuvanesh, N.; Gladysz, J.A. Toward Permetalated Alkyne/Azide 3+2 or “Click” Cycloadducts. Organometallics 2012, 31, 5231–5234. [Google Scholar] [CrossRef]
  19. Stengel, I.; Strassert, C.A.; Plummer, E.A.; Chien, C.-H.; De Cola, L.; Bäuerle, P. Postfunctionalization of Luminescent Bipyridine PtII Bisacetylides by Click Chemistry. Eur. J. Inorg. Chem. 2012, 2012, 1795–1809. [Google Scholar] [CrossRef]
  20. Li, Y.; Tsang, D.P.; Chan, C.K.; Wong, K.M.; Chan, M.; Yam, V.W. Synthesis of Unsymmetric Bipyridine–Pt II –Alkynyl Complexes through Post-Click Reaction with Emission Enhancement Characteristics and Their Applications as Phosphorescent Organic Light-Emitting Diodes. Chem.–A Eur. J. 2014, 20, 13710–13715. [Google Scholar] [CrossRef]
  21. Stengel, I.; Strassert, C.A.; De Cola, L.; Bäuerle, P. Tracking Intramolecular Interactions in Flexibly Linked Binuclear Platinum(II) Complexes. Organometallics 2014, 33, 1345–1355. [Google Scholar] [CrossRef]
  22. Stengel, I.; Götz, G.; Weil, M.; Bäuerle, P. A Dinuclear (bpy)PtII-Decorated Crownophane. Eur. J. Org. Chem. 2015, 2015, 3887–3893. [Google Scholar] [CrossRef]
  23. Sutton, E.C.; McDevitt, C.E.; Yglesias, M.V.; Cunningham, R.M.; DeRose, V.J. Tracking the cellular targets of platinum anticancer drugs: Current tools and emergent methods. Inorg. Chim. Acta 2019, 498, 118984. [Google Scholar] [CrossRef]
  24. Guo, C.; Nolan, E.M. Heavy-Metal Trojan Horse: Enterobactin-Directed Delivery of Platinum(IV) Prodrugs to Escherichia coli. J. Am. Chem. Soc. 2022, 144, 12756–12768. [Google Scholar] [CrossRef]
  25. Xu, G.-L.; Ren, T. Covalent Modification of Diruthenium Alkynyl Compounds: Novel Application of Click Reactions in Organometallic Chemistry. Organometallics 2005, 24, 2564–2566. [Google Scholar] [CrossRef]
  26. Chen, W.-Z.; Fanwick, P.E.; Ren, T. Dendronized Diruthenium Compounds via the Copper(I)-Catalyzed Click Reaction. Inorg. Chem. 2007, 46, 3429–3431. [Google Scholar] [CrossRef]
  27. Forrest, W.P.; Cao, Z.; Chen, W.-Z.; Hassell, K.M.; Kharlamova, A.; Jakstonyte, G.; Ren, T. Decorating Diruthenium Compounds with Fréchet Dendrons via the Click Reaction. Inorg. Chem. 2011, 50, 9345–9353. [Google Scholar] [CrossRef]
  28. Chitre, K.P.; Guillén, E.; Yoon, A.S.; Galoppini, E. Synthesis of Homoleptic Ruthenium “Star” Complexes by Click Reaction for TiO2 Sensitization. Eur. J. Inorg. Chem. 2012, 2012, 5461–5464. [Google Scholar] [CrossRef]
  29. Baron, A.; Herrero, C.; Quaranta, A.; Charlot, M.-F.; Leibl, W.; Vauzeilles, B.; Aukauloo, A. Click Chemistry on a Ruthenium Polypyridine Complex. An Efficient and Versatile Synthetic Route for the Synthesis of Photoactive Modular Assemblies. Inorg. Chem. 2012, 51, 5985–5987. [Google Scholar] [CrossRef]
  30. Wintergerst, P.; Witas, K.; Nauroozi, D.; Schmid, M.; Dikmen, E.; Tschierlei, S.; Rau, S. Minimizing Side Product Formation in Alkyne Functionalization of Ruthenium Complexes by Introduction of Protecting Groups. Z. Anorg. Allg. Chem. 2020, 646, 842–848. [Google Scholar] [CrossRef]
  31. Gotico, P.; Tran, T.; Baron, A.; Vauzeilles, B.; Lefumeux, C.; Ha-Thi, M.; Pino, T.; Halime, Z.; Quaranta, A.; Leibl, W.; et al. Tracking Charge Accumulation in a Functional Triazole-Linked Ruthenium-Rhenium Dyad Towards Photocatalytic Carbon Dioxide Reduction. ChemPhotoChem 2021, 5, 654–664. [Google Scholar] [CrossRef]
  32. Busemann, A.; Araman, C.; Flaspohler, I.; Pratesi, A.; Zhou, X.-Q.; van Rixel, V.H.S.; Siegler, M.A.; Messori, L.; van Kasteren, S.I.; Bonnet, S. Alkyne Functionalization of a Photoactivated Ruthenium Polypyridyl Complex for Click-Enabled Serum Albumin Interaction Studies. Inorg. Chem. 2020, 59, 7710–7720. [Google Scholar] [CrossRef] [PubMed]
  33. Herrero, C.; Batchelor, L.; Baron, A.; El Ghachtouli, S.; Sheth, S.; Guillot, R.; Vauzeilles, B.; Sircoglou, M.; Mallah, T.; Leibl, W.; et al. Click Chemistry as a Convenient Tool for the Incorporation of a Ruthenium Chromophore and a Nickel–Salen Monomer into a Visible-Light-Active Assembly. Eur. J. Inorg. Chem. 2013, 2013, 494–499. [Google Scholar] [CrossRef]
  34. Mede, T.; Jäger, M.; Schubert, U.S. “Chemistry-on-the-complex”: Functional Ru II polypyridyl-type sensitizers as divergent building blocks. Chem. Soc. Rev. 2018, 47, 7577–7627. [Google Scholar] [CrossRef] [PubMed]
  35. Belov, A.S.; Prikhod’ko, A.I.; Novikov, V.V.; Vologzhanina, A.V.; Bubnov, Y.N.; Voloshin, Y.Z. First “Click” Synthesis of the Ribbed-Functionalized Metal Clathrochelates: Cycloaddition of Benzyl Azide to Propargylamine Iron(II) Macrobicycle and the Unexpected Transformations of the Resulting Cage Complex. Eur. J. Inorg. Chem. 2012, 2012, 4507–4514. [Google Scholar] [CrossRef]
  36. Zelinskii, G.E.; Belov, A.S.; Vologzhanina, A.V.; Limarev, I.P.; Pavlov, A.A.; Olshevskaya, V.A.; Makarenkov, A.V.; Dorovatovskii, P.V.; Lebed, E.G.; Voloshin, Y.Z. Iron(II) Clathrochelate with Terminal Triple C≡C Bond and Its Carboranoclathrochelate Derivative with a Flexible Linker between the Polyhedral Cages: Synthesis and X-ray Structure. ChemistrySelect 2019, 4, 11572–11577. [Google Scholar] [CrossRef]
  37. Rapakousiou, A.; Mouche, C.; Duttine, M.; Ruiz, J.; Astruc, D. Click Synthesis and Redox Chemistry of Mono- and Heterobimetallic Triazolyl and Triazolium-Ferrocene and Cobalticinium Complexes. Eur. J. Inorg. Chem. 2012, 2012, 5071–5077. [Google Scholar] [CrossRef]
  38. Yu, G.; Suzaki, Y.; Abe, T.; Osakada, K. Introduction of ferrocene-containing [2]rotaxanes onto siloxane, silsesquioxane and polysiloxanes via click chemistry. Dalt. Trans. 2013, 42, 1476–1482. [Google Scholar] [CrossRef]
  39. Sun, R.; Wang, H.; Hu, J.; Zhao, J.; Zhang, H. Pyridine-phosphinimine ligand-accelerated Cu(i)-catalyzed azide–alkyne cycloaddition for preparation of 1-(pyridin-2-yl)-1,2,3-triazole derivatives. Org. Biomol. Chem. 2014, 12, 5954. [Google Scholar] [CrossRef]
  40. Wieczorek, A.; Błauż, A.; Makal, A.; Rychlik, B.; Plażuk, D. Synthesis and evaluation of biological properties of ferrocenyl–podophyllotoxin conjugates. Dalt. Trans. 2017, 46, 10847–10858. [Google Scholar] [CrossRef]
  41. Biegański, P.; Kovalski, E.; Israel, N.; Dmitrieva, E.; Trzybiński, D.; Woźniak, K.; Vrček, V.; Godel, M.; Riganti, C.; Kopecka, J.; et al. Electronic Coupling in 1,2,3-Triazole Bridged Ferrocenes and Its Impact on Reactive Oxygen Species Generation and Deleterious Activity in Cancer Cells. Inorg. Chem. 2022, 61, 9650–9666. [Google Scholar] [CrossRef]
  42. Aucott, B.J.; Ward, J.S.; Andrew, S.G.; Milani, J.; Whitwood, A.C.; Lynam, J.M.; Parkin, A.; Fairlamb, I.J.S. Redox-Tagged Carbon Monoxide-Releasing Molecules (CORMs): Ferrocene-Containing [Mn(C^N)(CO)4] Complexes as a Promising New CORM Class. Inorg. Chem. 2017, 56, 5431–5440. [Google Scholar] [CrossRef] [PubMed]
  43. Larionov, V.A.; Adonts, H.V.; Gugkaeva, Z.T.; Smol’yakov, A.F.; Saghyan, A.S.; Miftakhov, M.S.; Kuznetsova, S.A.; Maleev, V.I.; Belokon, Y.N. The Elaboration of a General Approach to the Asymmetric Synthesis of 1,4-Substituted 1,2,3-Triazole Containing Amino Acids via Ni(II) Complexes. ChemistrySelect 2018, 3, 3107–3110. [Google Scholar] [CrossRef]
  44. Li, Y.; Zhao, Y.; Flood, A.H.; Liu, C.; Liu, H.; Li, Y. Nanometer-Sized Reactor-A Porphyrin-Based Model System for Anion Species. Chem.—A Eur. J. 2011, 17, 7499–7505. [Google Scholar] [CrossRef]
  45. Husain, A.; Ganesan, A.; Sebastian, M.; Makhseed, S. Large ultrafast nonlinear optical response and excellent optical limiting behaviour in pyrene-conjugated zinc(II) phthalocyanines at a near-infrared wavelength. Dye. Pigment. 2021, 184, 108787. [Google Scholar] [CrossRef]
  46. Ladomenou, K.; Nikolaou, V.; Charalambidis, G.; Coutsolelos, A.G. “Click”-reaction: An alternative tool for new architectures of porphyrin based derivatives. Coord. Chem. Rev. 2016, 306, 1–42. [Google Scholar] [CrossRef]
  47. Araújo, A.R.L.; Tomé, A.C.; Santos, C.I.M.; Faustino, M.A.F.; Neves, M.G.P.M.S.; Simões, M.M.Q.; Moura, N.M.M.; Abu-Orabi, S.T.; Cavaleiro, J.A.S. Azides and Porphyrinoids: Synthetic Approaches and Applications. Part 1—Azides, Porphyrins and Corroles. Molecules 2020, 25, 1662. [Google Scholar] [CrossRef]
  48. Baeza, B.; Casarrubios, L.; Sierra, M.A. A Click Approach to Polymetallic Chromium(0) and Tungsten(0) Fischer-Carbene Complexes and Their Use in the Synthesis of Functionalized Polymetallic Metal ðCarbene Complexes. Chem. —A Eur. J. 2013, 19, 1429–1435. [Google Scholar] [CrossRef] [PubMed]
  49. Lewe, V.; Preuss, M.; Woźnica, E.A.; Spitzer, D.; Otter, R.; Besenius, P. A clickable NHC–Au(i)-complex for the preparation of stimulus-responsive metallopeptide amphiphiles. Chem. Commun. 2018, 54, 9498–9501. [Google Scholar] [CrossRef]
  50. O’Malley, W.I.; Abdelkader, E.H.; Aulsebrook, M.L.; Rubbiani, R.; Loh, C.-T.; Grace, M.R.; Spiccia, L.; Gasser, G.; Otting, G.; Tuck, K.L.; et al. Luminescent Alkyne-Bearing Terbium(III) Complexes and Their Application to Bioorthogonal Protein Labeling. Inorg. Chem. 2016, 55, 1674–1682. [Google Scholar] [CrossRef]
  51. Szíjjártó, C.; Pershagen, E.; Borbas, K.E. Functionalisation of lanthanide complexes via microwave-enhanced Cu(i)-catalysed azide–alkyne cycloaddition. Dalt. Trans. 2012, 41, 7660. [Google Scholar] [CrossRef]
  52. Goswami, L.N.; Ma, L.; Chakravarty, S.; Cai, Q.; Jalisatgi, S.S.; Hawthorne, M.F. Discrete Nanomolecular Polyhedral Borane Scaffold Supporting Multiple Gadolinium(III) Complexes as a High Performance MRI Contrast Agent. Inorg. Chem. 2013, 52, 1694–1700. [Google Scholar] [CrossRef] [PubMed]
  53. Tropiano, M.; Kenwright, A.M.; Faulkner, S. Lanthanide Complexes of Azidophenacyl-DO3A as New Synthons for Click Chemistry and the Synthesis of Heterometallic Lanthanide Arrays. Chem. —A Eur. J. 2015, 21, 5697–5699. [Google Scholar] [CrossRef] [PubMed]
  54. Tropiano, M.; Kilah, N.L.; Morten, M.; Rahman, H.; Davis, J.J.; Beer, P.D.; Faulkner, S. Reversible Luminescence Switching of a Redox-Active Ferrocene–Europium Dyad. J. Am. Chem. Soc. 2011, 133, 11847–11849. [Google Scholar] [CrossRef] [PubMed]
  55. Tropiano, M.; Record, C.J.; Morris, E.; Rai, H.S.; Allain, C.; Faulkner, S. Synthesis and Spectroscopic Study of d–f Hybrid Lanthanide Complexes Derived from triazolylDO3A. Organometallics 2012, 31, 5673–5676. [Google Scholar] [CrossRef]
  56. Sørensen, T.J.; Tropiano, M.; Blackburn, O.A.; Tilney, J.A.; Kenwright, A.M.; Faulkner, S. Preparation and study of an f,f,f′,f′′ covalently linked tetranuclear hetero-trimetallic complex—A europium, terbium, dysprosium triad. Chem. Commun. 2013, 49, 783–785. [Google Scholar] [CrossRef]
  57. Perez-Gavilan, A.; Carroll, P.J.; Sneddon, L.G. Efficient and Systematic Click-Based Synthetic Routes to Amino Acid Functionalized Metallatricarbadecaboranes. Organometallics 2012, 31, 2741–2748. [Google Scholar] [CrossRef]
  58. Druzina, A.A.; Kosenko, I.D.; Zhidkova, O.B.; Ananyev, I.V.; Timofeev, S.V.; Bregadze, V.I. Novel Cobalt Bis(dicarbollide) Based on Terminal Alkynes and Their Click-Reactions. Eur. J. Inorg. Chem. 2020, 2020, 2658–2665. [Google Scholar] [CrossRef]
  59. Linnenberg, O.; Kondinski, A.; Stöcker, C.; Monakhov, K.Y. The Cu(i)-catalysed Huisgen 1,3-dipolar cycloaddition route to (bio-)organic functionalisation of polyoxovanadates. Dalt. Trans. 2017, 46, 15636–15640. [Google Scholar] [CrossRef]
  60. Petrovskii, S.K.; Khistiaeva, V.V.; Sizova, A.A.; Sizov, V.V.; Paderina, A.V.; Koshevoy, I.O.; Monakhov, K.Y.; Grachova, E.V. Hexavanadate-Organogold(I) Hybrid Compounds: Synthesis by the Azide-Alkyne Cycloaddition and Density Functional Theory Study of an Intriguing Electron Density Distribution. Inorg. Chem. 2020, 59, 16122–16126. [Google Scholar] [CrossRef]
  61. Zabarska, N.; Stumper, A.; Rau, S. CuAAC click reactions for the design of multifunctional luminescent ruthenium complexes. Dalt. Trans. 2016, 45, 2338–2351. [Google Scholar] [CrossRef]
  62. Wirth, R.; White, J.D.; Moghaddam, A.D.; Ginzburg, A.L.; Zakharov, L.N.; Haley, M.M.; DeRose, V.J. Azide vs Alkyne Functionalization in Pt(II) Complexes for Post-treatment Click Modification: Solid-State Structure, Fluorescent Labeling, and Cellular Fate. J. Am. Chem. Soc. 2015, 137, 15169–15175. [Google Scholar] [CrossRef] [PubMed]
  63. Salmain, M.; Fischer-Durand, N.; Rudolf, B. Bioorthogonal Conjugation of Transition Organometallic Complexes to Peptides and Proteins: Strategies and Applications. Eur. J. Inorg. Chem. 2020, 2020, 21–35. [Google Scholar] [CrossRef]
  64. Wang, X.-Y.; Kimyonok, A.; Weck, M. Functionalization of polymers with phosphorescent iridium complexes via click chemistry. Chem. Commun. 2006, 37, 3933. [Google Scholar] [CrossRef] [PubMed]
  65. Baranovskii, E.M.; Khistiaeva, V.V.; Deriabin, K.V.; Petrovskii, S.K.; Koshevoy, I.O.; Kolesnikov, I.E.; Grachova, E.V.; Islamova, R.M. Re(I) Complexes as Backbone Substituents and Cross-Linking Agents for Hybrid Luminescent Polysiloxanes and Silicone Rubbers. Molecules 2021, 26, 6866. [Google Scholar] [CrossRef]
  66. Li, P.-Z.; Wang, X.-J.; Zhao, Y. Click chemistry as a versatile reaction for construction and modification of metal-organic frameworks. Coord. Chem. Rev. 2019, 380, 484–518. [Google Scholar] [CrossRef]
  67. González Cabrera, D.; Koivisto, B.D.; Leigh, D.A. A metal-complex-tolerant CuAAC ‘click’ protocol exemplified through the preparation of homo- and mixed-metal-coordinated [2]rotaxanes. Chem. Commun. 2007, 41, 4218. [Google Scholar] [CrossRef]
  68. Grosshenny, V.; Romero, F.M.; Ziessel, R. Construction of Preorganized Polytopic Ligands via Palladium-Promoted Cross-Coupling Reactions. J. Org. Chem. 1997, 62, 1491–1500. [Google Scholar] [CrossRef]
  69. Ponce, J.; Aragó, J.; Vayá, I.; Magenti, J.G.; Tatay, S.; Ortí, E.; Coronado, E. Photophysical Properties of Oligo (phenylene ethynylene) Iridium(III) Complexes Functionalized with Metal-Anchoring Groups. Eur. J. Inorg. Chem. 2016, 2016, 1851–1859. [Google Scholar] [CrossRef]
  70. Ni, J.; Wu, Y.-H.; Zhang, X.; Li, B.; Zhang, L.-Y.; Chen, Z.-N. Luminescence Vapochromism of a Platinum(II) Complex for Detection of Low Molecular Weight Halohydrocarbon. Inorg. Chem. 2009, 48, 10202–10210. [Google Scholar] [CrossRef]
  71. Guha, R.; Defayay, D.; Hepp, A.; Müller, J. Targeting Guanine Quadruplexes with Luminescent Platinum(II) Complexes Bearing a Pendant Nucleobase. Chempluschem 2021, 86, 662–673. [Google Scholar] [CrossRef]
  72. Ma, D.; Qiu, Y.; Duan, L. New Insights into Tunable Volatility of Ionic Materials through Counter-Ion Control. Adv. Funct. Mater. 2016, 26, 3438–3445. [Google Scholar] [CrossRef]
  73. Suhr, K.J.; Bastatas, L.D.; Shen, Y.; Mitchell, L.A.; Frazier, G.A.; Taylor, D.W.; Slinker, J.D.; Holliday, B.J. Phenyl substitution of cationic bis-cyclometalated iridium(iii) complexes for iTMC-LEECs. Dalt. Trans. 2016, 45, 17807–17823. [Google Scholar] [CrossRef] [PubMed]
  74. Li, X.; Tong, X.; Yin, Y.; Yan, H.; Lu, C.; Huang, W.; Zhao, Q. Using highly emissive and environmentally sensitive o-carborane-functionalized metallophosphors to monitor mitochondrial polarity. Chem. Sci. 2017, 8, 5930–5940. [Google Scholar] [CrossRef] [PubMed]
  75. Kurz, P.; Probst, B.; Spingler, B.; Alberto, R. Ligand Variations in [ReX(diimine)(CO)3] Complexes: Effects on Photocatalytic CO2 Reduction. Eur. J. Inorg. Chem. 2006, 2006, 2966–2974. [Google Scholar] [CrossRef]
  76. Chan, S.-C.; Chan, M.C.W.; Wang, Y.; Che, C.-M.; Cheung, K.-K.; Zhu, N. Organic Light-Emitting Materials Based on Bis(arylacetylide)platinum(II) Complexes Bearing Substituted Bipyridine and Phenanthroline Ligands: Photo- and Electroluminescence from 3MLCT Excited States. Chem. —A Eur. J. 2001, 7, 4180–4190. [Google Scholar] [CrossRef]
  77. Wei, Q.-H.; Yin, G.-Q.; Zhang, L.-Y.; Chen, Z.-N. Luminescent Hepta- and Tetradecanuclear Complexes of 5,5′-Diethynyl-2,2′-bipyridine Capped with Triangular Trinuclear Cu3/Ag3 Cluster Units. Inorg. Chem. 2006, 45, 10371–10377. [Google Scholar] [CrossRef] [PubMed]
  78. Hissler, M.; Connick, W.B.; Geiger, D.K.; McGarrah, J.E.; Lipa, D.; Lachicotte, R.J.; Eisenberg, R. Platinum Diimine Bis(acetylide) Complexes: Synthesis, Characterization, and Luminescence Properties. Inorg. Chem. 2000, 39, 447–457. [Google Scholar] [CrossRef]
  79. Lamansky, S.; Djurovich, P.; Murphy, D.; Abdel-Razzaq, F.; Lee, H.E.; Adachi, C.; Burrows, P.E.; Forrest, S.R.; Thompson, M.E. Highly phosphorescent bis-cyclometalated iridium complexes: Synthesis, photophysical characterization, and use in organic light emitting diodes. J. Am. Chem. Soc. 2001, 123, 4304–4312. [Google Scholar] [CrossRef]
  80. CrysAlisPro, Rigaku Oxford Diffraction; Version 1.171.39.35a; Agilent Technologies: Santa Clara, CA, USA, 2017.
  81. Sheldrick, G.M. SHELXT—Integrated space-group and crystal-structure determination. Acta Crystallogr. Sect. A Found. Crystallogr. 2015, 71, 3–8. [Google Scholar] [CrossRef]
  82. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Crystallogr. Sect. C 2015, 71, 3–8. [Google Scholar] [CrossRef]
  83. Dolomanov, O.V.; Bourhis, L.J.; Gildea, R.J.; Howard, J.A.K.; Puschmann, H. OLEX2: A complete structure solution, refinement and analysis program. J. Appl. Crystallogr. 2009, 42, 339–341. [Google Scholar] [CrossRef]
Scheme 1. Schematic representation of CuAAC reaction between functionalized metal complex and substituted organic azide.
Scheme 1. Schematic representation of CuAAC reaction between functionalized metal complex and substituted organic azide.
Molecules 27 06494 sch001
Scheme 2. Synthesis of functionalized NN compounds XP (X = 1–3). i = Pd(PPh3)2Cl2, CuI, HN(iPr)2, (CH3)3SiC2H; ii = THF, RT, 24 h; iii = THF, 45 °C, 72 h.
Scheme 2. Synthesis of functionalized NN compounds XP (X = 1–3). i = Pd(PPh3)2Cl2, CuI, HN(iPr)2, (CH3)3SiC2H; ii = THF, RT, 24 h; iii = THF, 45 °C, 72 h.
Molecules 27 06494 sch002
Scheme 3. Synthesis and post-functionalization of Re(I) complexes. i = Re(CO)5Cl, refluxing toluene, 3 h; ii = CuSO4×5H2O, NaAsc, acetone/H2O, r. t., 12 h.
Scheme 3. Synthesis and post-functionalization of Re(I) complexes. i = Re(CO)5Cl, refluxing toluene, 3 h; ii = CuSO4×5H2O, NaAsc, acetone/H2O, r. t., 12 h.
Molecules 27 06494 sch003
Scheme 4. Synthesis and post-functionalization of Ir(III) complexes. i = [Ir(C^N)2Cl]2, CH2Cl2/MeOH, excess of KPF6, r.t., 24 h; ii = CuSO4×5H2O, NaAsc, acetone/H2O, r.t., 48 h. # Reaction uncompleted.
Scheme 4. Synthesis and post-functionalization of Ir(III) complexes. i = [Ir(C^N)2Cl]2, CH2Cl2/MeOH, excess of KPF6, r.t., 24 h; ii = CuSO4×5H2O, NaAsc, acetone/H2O, r.t., 48 h. # Reaction uncompleted.
Molecules 27 06494 sch004
Scheme 5. Synthesis of Pt(II) complexes. i = (DMSO)2PtCl2, MeCN, r.t.; ii = C2Ph, CuI, iPr2NH, CH2Cl2, r.t., 12 h.
Scheme 5. Synthesis of Pt(II) complexes. i = (DMSO)2PtCl2, MeCN, r.t.; ii = C2Ph, CuI, iPr2NH, CH2Cl2, r.t., 12 h.
Molecules 27 06494 sch005
Scheme 6. Post-functionalization of Pt(dtbpy) complex. i = CuSO4 × 5H2O, NaAsc, CH2Cl2/MeOH, 40 °C, 48 h.
Scheme 6. Post-functionalization of Pt(dtbpy) complex. i = CuSO4 × 5H2O, NaAsc, CH2Cl2/MeOH, 40 °C, 48 h.
Molecules 27 06494 sch006
Figure 1. Molecular structure of Pt(dtbpy)C complex.
Figure 1. Molecular structure of Pt(dtbpy)C complex.
Molecules 27 06494 g001
Figure 2. Absorption spectra and normalized emission (λext 365 nm) spectra of Re(I), Pt(II), and Ir(III) “clicked” complexes, DCE solution, 10−5 M, r.t.
Figure 2. Absorption spectra and normalized emission (λext 365 nm) spectra of Re(I), Pt(II), and Ir(III) “clicked” complexes, DCE solution, 10−5 M, r.t.
Molecules 27 06494 g002
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Petrovskii, S.; Khistiaeva, V.; Paderina, A.; Abramova, E.; Grachova, E. Post-Functionalization of Organometallic Complexes via Click-Reaction. Molecules 2022, 27, 6494. https://doi.org/10.3390/molecules27196494

AMA Style

Petrovskii S, Khistiaeva V, Paderina A, Abramova E, Grachova E. Post-Functionalization of Organometallic Complexes via Click-Reaction. Molecules. 2022; 27(19):6494. https://doi.org/10.3390/molecules27196494

Chicago/Turabian Style

Petrovskii, Stanislav, Viktoria Khistiaeva, Aleksandra Paderina, Evgenia Abramova, and Elena Grachova. 2022. "Post-Functionalization of Organometallic Complexes via Click-Reaction" Molecules 27, no. 19: 6494. https://doi.org/10.3390/molecules27196494

Article Metrics

Back to TopTop